1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: \documentclass[eqsecnum,twocolumn,showpacs,amsmath,amssymb]{revtex4}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \usepackage{bm}
5: \usepackage{graphicx,subfigure,afterpage}
6: \usepackage{amsfonts}
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8: %% the paper text follows this prescription, the diagram labels and file names follow the original research notes -- hence the discrepancy
9: %prescription for converting from the parameters of my numerics to more sensible ones
10: %multiply alpha by 10 to get the paper's alpha
11: %multiply beta by 0.8e-4 to get l
12: %mulitly gamm by 4e-4 to get xi
13:
14: \newcommand{\etal}{{\it et al.}}
15:
16: \newcommand{\djs}{{\mathfrak{D}}_{\rm M}}
17: \newcommand{\df}{{\mathfrak{D}}_{\rm F}}
18: \newcommand{\di}{{\mathfrak{D}}_{\rm i}}
19:
20: \newcommand{\tdjs}{{\mathfrak{\tilde{D}}}_{\rm M}}
21: \newcommand{\tdf}{{\mathfrak{\tilde{D}}}_{\rm F}}
22: \newcommand{\tdi}{{\mathfrak{\tilde{D}}}_{\rm i}}
23:
24: \newcommand{\hdjs}{{\mathfrak{\hat{D}}}_{\rm M}}
25:
26: \newcommand{\vect}[1]{\underline{#1}}
27: \newcommand{\tens}[1]{\underline{\underline{#1}}}
28:
29: \newcommand{\vm}{\vect{v}_{\,\rm m}}
30: \newcommand{\vs}{\vect{v}_{\,\rm s}}
31:
32: \newcommand{\vrel}{\vect{v}_{\, \rm rel}}
33: \newcommand{\vv}{\vect{v}}
34:
35: \newcommand{\dx}{\Delta x}
36: \newcommand{\ny}{N_y}
37: \newcommand{\dt}{\Delta t}
38:
39: \newcommand{\etam}{\eta_{\,\rm m}}
40: \newcommand{\etas}{\eta_{\,\rm s}}
41:
42: \newcommand{\Dm}{\tens{D}_{\,\rm m}}
43: \newcommand{\Ds}{\tens{D}_{\,\rm s}}
44:
45: \newcommand{\Drel}{\tens{D}_{\,\rm rel}}
46: \newcommand{\Dv}{\tens{D}}
47:
48: \newcommand{\Omm}{\tens{\Omega}_{\,\rm m}}
49: \newcommand{\Oms}{\tens{\Omega}_{\,\rm s}}
50:
51: \newcommand{\nablu}{\vect{\nabla}}
52:
53: \newcommand{\gdotb}{\bar{\dot{\gamma}}}
54: \newcommand{\phib}{\bar{\phi}}
55:
56: \newcommand{\be}{\begin{equation}}
57: \newcommand{\ee}{\end{equation}}
58: \newcommand{\bea}{\begin{eqnarray}}
59: \newcommand{\eea}{\end{eqnarray}}
60: \newcommand{\lav}{\left\langle}
61: \newcommand{\rav}{\right\rangle}
62: \newcommand{\gdot}{\dot{\gamma}}
63: \newcommand{\bgdot}{\bar{\dot{\gamma}}}
64: \newcommand{\la}{\lambda}
65: \newcommand{\de}{\Delta}
66: \newcommand{\sto}{\rightarrow}
67: \newcommand{\D}{\displaystyle}
68: \newcommand{\taur}{\tau_{\rm r}}
69: \newcommand{\Tf}{T_{\rm f}}
70: \newcommand{\gae}{\stackrel{>}{\scriptstyle\sim}}
71: \newcommand{\lae}{\stackrel{<}{\scriptstyle\sim}}
72: \newcommand{\ie}{{\it i.e.\/}}
73: \newcommand{\eg}{{\it e.g.\/}}
74: \newcommand{\etc}{{\it etc.\/}}
75: \newcommand{\versus}{{\it vs.\/}}
76: \newcommand{\eps}{\epsilon}
77: \newcommand{\Deta}{\Delta{\eta}}
78: \newcommand{\etab}{{\eta}}
79: \newcommand{\etabar}{\bar{\eta}}
80: \newcommand{\bw}{\begin{widetext}}
81: \newcommand{\ew}{\end{widetext}}
82:
83: \newcommand{\bmini}{\begin{minipage}}
84: \newcommand{\emini}{\end{minipage}}
85:
86: \newcommand{\eigenvec}{\vect{{\tt{v}}}_{\vect{k},\alpha}}
87: \newcommand{\eigenvecmax}{\vect{{\tt{v}}}_{\vect{k},\Upsilon}}
88:
89: \newcommand{\phil}{\phi_\ell}
90: \newcommand{\phic}{\bar{\phi}_{\rm c}}
91: \newcommand{\gdotc}{\bar{\gdot}_{\rm c}}
92: \newcommand{\model}{d-JS-$\phi$}
93:
94:
95: \newcommand{\sigmas}{\Sigma_{\rm sel}}
96: \newcommand{\gdoth}{\gdot_{\rm h}}
97: \newcommand{\gdotl}{\gdot_\ell}
98:
99: %\numberwithin{equation}{section}
100:
101: \begin{document}
102:
103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
104: \title{Flow phase diagrams for concentration-coupled shear banding}
105: \author{S. M. Fielding}
106: \email{physf@irc.leeds.ac.uk}
107: \author{P. D. Olmsted}
108: \email{p.d.olmsted@leeds.ac.uk}
109: \affiliation{Polymer IRC and
110: Department of Physics \& Astronomy, University of Leeds, Leeds LS2
111: 9JT, United Kingdom}
112: \date{\today}
113: \begin{abstract}
114: After surveying the experimental evidence for concentration coupling
115: in the shear banding of wormlike micellar surfactant systems, we
116: present flow phase diagrams spanned by shear stress (or strain-rate)
117: and concentration, calculated within the two-fluid, non-local
118: Johnson-Segalman (\model) model. We also give results for the
119: macroscopic flow curves $\Sigma(\gdotb,\phib)$ for a range of
120: (average) concentrations $\phib$. For any concentration that is high
121: enough to give shear banding, the flow curve shows the usual
122: non-analytic kink at the onset of banding, followed by a coexistence
123: ``plateau'' that slopes upwards, $d\Sigma/d\gdotb>0$. As the
124: concentration is reduced, the width of the coexistence regime
125: diminishes and eventually terminates at a non-equilibrium critical
126: point $[\Sigma_{\rm c},\phib_{\rm c},\gdotb_{\rm c}]$. We outline
127: the way in which the flow phase diagram can be reconstructed from a
128: family of such flow curves, $\Sigma(\gdotb,\phib)$, measured for
129: several different values of $\phib$. This reconstruction could be
130: used to check new measurements of concentration differences between
131: the coexisting bands. Our \model\ model contains two different
132: spatial gradient terms that describe the interface between the shear
133: bands. The first is in the viscoelastic constitutive equation, with
134: a characteristic (mesh) length $l$. The second is in the
135: (generalised) Cahn-Hilliard equation, with the characteristic length
136: $\xi$ for equilibrium concentration-fluctuations. We show that the
137: phase diagrams (and so also the flow curves) depend on the ratio
138: $r\equiv l/\xi$, with loss of unique state selection at $r=0$. We
139: also give results for the full shear-banded profiles, and study the
140: divergence of the interfacial width (relative to $l$ and $\xi$) at
141: the critical point.
142:
143: \end{abstract}
144: \pacs{{47.50.+d}{ Non-Newtonian fluid flows}--
145: {47.20.-k}{ Hydrodynamic stability}--
146: {36.20.-r}{ Macromolecules and polymer molecules}
147: }
148: \maketitle
149:
150: \section{Introduction}
151: \label{sec:intro}
152:
153: For many complex fluids, the intrinsic constitutive curve of shear
154: stress $\Sigma$ as a function of shear rate $\gdot$ is non-monotonic,
155: admitting multiple values of shear rate at common stress. For
156: semi-dilute wormlike micelles, theory~\cite{cates90,SpenCate94,SCM93}
157: predicts the form ACEG of Fig.~\ref{fig:schem}. In the range
158: $\gdot_{\rm c1}<\gdot<\gdot_{\rm c2}$ where the stress is decreasing,
159: steady homogeneous flow (Fig.~\ref{fig:picture}a) is
160: unstable~\cite{Yerushalmi70}. For an applied shear rate $\bar{\gdot}$
161: in this unstable range, Spenley, Cates and McLeish~\cite{SCM93}
162: proposed that the system separates into high and low shear rate bands
163: ($\gdoth$ and $\gdotl$; Fig.~\ref{fig:picture}b) and that
164: any change in the applied shear rate then merely adjusts the relative
165: fraction of the bands, while the stress $\sigmas$ (which is common to
166: both) remains constant. The steady state flow curve then has the form
167: ABFG. Several constitutive models augmented with interfacial gradient
168: terms have captured this
169: behaviour~\cite{olmsted99a,lu99,olmstedlu97,spenley96}.
170:
171: %
172: \begin{figure}[h]
173: \begin{center}
174: \includegraphics[scale=1.0]{makefigure1.ps}
175: %\input{schem.pstex_t}
176: %\centerline{\psfig{figure=schem.eps,width=8cm}}
177: \caption{Schematic flow curves for wormlike micelles: the homogeneous
178: constitutive curve is ACEG; the steady shear-banded flow curve is BF
179: (without concentration coupling in planar shear) or B'F' (with
180: concentration coupling, or in a cylindrical Couette device).
181: \label{fig:schem} }
182: \end{center}
183: \end{figure}
184: %
185: %
186: \begin{figure}[h]
187: \includegraphics[scale=0.3]{picture.eps}
188: \caption{(a) Homogeneous shear rate and (b) banded profiles.
189: \label{fig:picture} }
190: \end{figure}
191: %
192:
193: Experimentally, this scenario has been widely observed in semidilute
194: wormlike micelles~\cite{berret94b,Call+96,grand97}. The steady state
195: flow curve (which is often attained only after very long
196: transients~\cite{grand97}) has a well defined, reproducible stress
197: plateau $\sigmas$. Coexistence of high and low viscosity bands has
198: been observed by NMR
199: spectroscopy~\cite{Call+96,MairCall96,MairCall96c,BritCall97}. Further
200: evidence comes from small angle neutron scattering
201: (SANS)~\cite{berret94a,schmitt94,Capp+97,rehage91,berret94b}; and from
202: flow birefringence (FB)~\cite{Decr+95,Makh+95,DCC97,BPD97}, which
203: reveals a (quasi) nematic birefringence band coexisting with an
204: isotropic one (but see~\cite{FisCal01,FisCal00}).
205:
206: In some systems, the coexistence plateau is not perfectly flat, but
207: slopes upward slightly with increasing shear rate (B'F' in
208: Fig.~\ref{fig:schem}). See, for example, Ref.~\cite{LerDecBer00} for
209: CTAB(0.3M)/${\rm NaNO}_3\rm {(1.79M)}$/H$_2$O at micellar volume
210: fraction $\phi=11\%$. This effect is much more pronounced in other,
211: more concentrated systems that are near an underlying (zero-shear)
212: isotropic-nematic (I-N) transition ($\phi\approx
213: 30\%$)~\cite{schmitt94,BRL98,berret94a}.
214:
215: In a cylindrical Couette geometry, this upward slope is qualitatively
216: consistent with the inhomogeneous stress arising from the cell
217: curvature: as the high-shear band at the inner cylinder expands
218: outward with increasing applied shear rate, the applied torque must
219: increase to ensure that the interface between the bands stays at the
220: selected stress $\sigmas$~\cite{olmsted99a}. However a more generic
221: explanation, independent of geometry, is that the shear banding
222: transition is coupled to concentration~\cite{schmitt95,olmstedlu97}.
223: In this case, the properties of each phase must change as the applied
224: shear rate is tracked through the coexistence regime, because material
225: is redistributed between the bands as the high shear band grows to
226: fill the gap.
227:
228: Generically, one expects flow to be coupled to concentration in
229: viscoelastic solutions where the different constituents (polymer and
230: solvent) have widely separated relaxation
231: timescales~\cite{brochdgen77,HelfFred89,DoiOnuk92,milner93,WPD91,beris94,SBJ97,tanak96}.
232: This was explained by Helfand and Fredrickson (HF)~\cite{HelfFred89}
233: as follows. In a sheared solution, the parts of an extended polymer
234: molecule (micelle for our purposes) in regions of lower viscosity
235: will, upon relaxing to equilibrium, move more than the parts mired in
236: a region of high viscosity and concentration. A relaxing molecule
237: therefore on average moves towards the higher concentration region.
238: This provides a positive feedback mechanism whereby micelles can move
239: {\em up} their own concentration gradient, and leads to flow-enhanced
240: concentration fluctuations perpendicular to the shear compression
241: axis.
242: This was observed in steadily sheared polymer solutions in the early
243: 1990's~\cite{WPD91}. In a remarkable paper, Schmitt
244: \etal~\cite{schmitt95} discussed the implications of this feedback
245: mechanism for the onset of flow instabilities. Strongly enhanced
246: concentration fluctuations were subsequently observed in the early
247: time kinetics of the shear banding instability in
248: Ref.~\cite{DecLerBer01}.
249:
250: Recently, therefore, we introduced a model of concentration-coupled
251: shear banding~\cite{FieOlm02,FieOlm02b} by combining the diffusive
252: Johnson Segalman (d-JS) model~\cite{johnson77,olmsted99a} with a
253: two-fluid approach~\cite{brochdgen77,milner91,deGen76,Brochard83} to
254: concentration fluctuations. This ``d-JS-$\phi$" model does not address
255: the microscopics of any particular viscoelastic system, but instead
256: should be regarded as a minimal model that combines (i) a constitutive
257: curve like that of semidilute wormlike micelles (Fig.~\ref{fig:schem})
258: with (ii) the non-local (interfacial) terms required for selection of
259: a unique banded state~\cite{lu99} and (iii) a simple approach to
260: concentration coupling.
261:
262: In Refs.~\cite{FieOlm02,FieOlm02b}, we examined the linear stability
263: of initially homogeneous shear states in this \model\ model with
264: respect to coupled fluctuations in shear rate $\gdot$, micellar strain
265: $\tens{W}$ and concentration $\phi$. We thereby calculated the
266: ``spinodal'', inside which such homogeneous states are unstable. We
267: also calculated the selected length scale at which inhomogeneity first
268: emerges during startup flows in the unstable region. In the limit of
269: zero concentration coupling, the unstable region coincides with that
270: of negative slope in the homogeneous constitutive curve, as expected;
271: but no length scale is selected during startup. Concentration coupling
272: enhances this instability at short length scales. It thereby broadens
273: the region of instability, and selects a length scale at which
274: inhomogeneity must emerge.
275:
276: In the present paper, we compute the corresponding steady-state flow
277: phase diagram (the ``binodals'' and their tie-lines). As far as we are
278: aware, this is the first concrete calculation aimed at qualitatively
279: describing concentration-coupled shear banding for systems such as
280: semi-dilute wormlike micelles. We start in Sec.~\ref{sec:background}
281: by describing the experimental background in more detail. We also
282: compare our present calculation with the only other existing one for
283: concentration-coupled shear banded states, in concentrated solutions of
284: rigid rods~\cite{olmsted99c}. In Sec.~\ref{sec:model} we summarize our
285: \model\ model. We then review our results for the spinodal onset of
286: instability in Sec.~\ref{sec:spinodals}. In Sec.~\ref{sec:numerics} we
287: describe our numerical procedure for computing the banded steady
288: states, with brief discussion of our careful study of mesh and finite
289: size effects. We then (Sec.~\ref{sec:results}) present our results
290: for the flow phase diagrams and shear-banded profiles. We conclude in
291: Sec.~\ref{sec:conclusion}.
292:
293: \section{Experimental background; theoretical context}
294: \label{sec:background}
295:
296: In this section, we discuss in more detail the experimental evidence
297: for concentration coupling in the shear banding of wormlike micelles.
298: We survey both (i) concentrated systems near the zero-shear I-N phase
299: transition and (ii) semidilute systems, in which underlying nematic
300: interactions are likely to be less important. Correspondingly, we
301: compare the present calculation (aimed at the semi-dilute systems)
302: with an earlier calculation of flow phase diagrams in concentrated
303: solutions of rigid rods (near the I-N transition)~\cite{olmsted99c}.
304:
305: The earliest observations of an upwardly sloping stress plateau in
306: wormlike micelles were made by Schmitt \etal\ \cite{schmitt94} and
307: Berret \etal~\cite{berret94a,BRL98}. Schmitt \etal\ \cite{schmitt94}
308: studied CpClO$_3$/NaClO$_3$(0.05M)/H$_2$O at the high micellar volume
309: fraction $\phi\approx 31\%$, just below the onset of the I-N
310: transition at $\phil=34\%$.
311: In the steady-state flow curve, the stress increased smoothly up to
312: the critical shear rate $\gdotl$, where it showed a pronounced
313: downward kink before curving upward again for $\gdot>\gdotl$
314: (qualitatively like B'F' in Fig.~\ref{fig:schem}). SANS measurements
315: confirmed a superposition of nematic and isotropic contributions in
316: this regime $\gdot>\gdotl$, with the nematic contribution rising
317: linearly from zero at $\gdot=\gdotl$.
318:
319: Berret \etal~\cite{berret94a,BRL98} studied CpCl/hexanol/NaCl for
320: several micellar volume fractions, again at a volume fraction just
321: below the onset of the zero-shear I-N transition ($\phil\approx
322: 32\%$). The overall height of the coexistence plateau (which again
323: sloped upwards in $\gdot$) was found to fall with increasing
324: surfactant concentration $\phi\to\phil$, extrapolating to zero at
325: $\phi\gae\phil$, which is already biphasic in zero shear. They also
326: found an increasing nematic contribution to SANS patterns for
327: increasing shear rates above $\gdotl$. They further used the SANS data
328: to show that the nematic (high shear) band was more concentrated than
329: the low shear band.
330:
331: As noted above, the majority of existing calculations of shear-banded
332: states have assumed uniform concentration. An important exception is
333: the calculation of Olmsted and Lu~\cite{olmsted99c}. Although this
334: model was aimed at concentrated solutions of rigid rods, it broadly
335: captured some of the experimental phenomenology for the concentrated
336: ($\phi\approx 30\%$) wormlike
337: micelles~\cite{schmitt94,berret94a,BRL98}. For example, the overall
338: height of the coexistence plateau $\sigmas$ increased from zero as the
339: concentration was reduced below the threshold $\phil$ of the
340: zero-shear I-N biphasic regime. The coexistence plateau sloped upward
341: markedly in shear rate. In further agreement with experiment, the high
342: shear (nematic) phase had a higher volume fraction of rods. It should
343: be noted that model of Ref.~\cite{olmsted99c} was explicitly aimed at
344: concentrated systems, which in zero shear are already close to the I-N
345: transition: hence, the dynamics of the relevant order parameter
346: $\tens{Q}$ was driven by a free energy that already contained a phase
347: transition. In contrast, the simple free energy $F_{\rm e}(\tens{W})$
348: we consider below has no underlying phase transition, and flow-induced
349: efects are driven by convective, rather than dissipative
350: (relaxational) dynamics.
351:
352: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
353: \begin{figure}[h]
354: \includegraphics[scale=0.35]{Berr94PDiagB.eps}
355: \caption{Height of the coexistence plateau in the system
356: CpCl/NaSal/brine. The 5 leftmost points are taken from the data of
357: Ref.~\cite{berret94b}. The righthand point represents the zero-shear
358: biphasic regime of this system, and is in accordance with the
359: extrapolation of $G(\phi)/\Sigma_{\rm sel}$ in Ref.~\cite{berret94b}
360: (see main text for details).
361: \label{fig:coexistence} }
362: \end{figure}
363: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
364:
365: There have also been several experimental studies of concentration
366: dependence in the shear banding of more dilute wormlike micellar
367: solutions. For example, Berret \etal~\cite{berret94b} investigated the
368: non-linear rheology of CpCl/NaSal/brine in the concentration range
369: $5\%-20\%$, well below the I-N$_{\rm c}$ transition at $\phi\approx
370: 36\%$. (N$_{\rm c}$ is ``nematic calamitic''.) In contrast to the more
371: concentrated ``prenematic'' systems, the plateau height $\sigmas$
372: decreased with {\em decreasing} concentration: see the left 5 points
373: in Fig.~\ref{fig:coexistence}. The width of the plateau also
374: decreased so that the difference $\gdot_{\rm h}-\gdot_\ell$ fell to
375: zero at a critical point $\phic,\gdotc,\Sigma_{\rm c}$ (leftmost point
376: in Fig.~\ref{fig:coexistence}). (These trends are the same as those in
377: Fig.~\ref{fig:alpha1.0e-3_gamma0.0:d} below.) In contrast the {\em
378: scaled} plateau height $\sigmas/G(\phi)$ (where $G$ is the plateau
379: modulus) decreased with {\em increasing} concentration and
380: extrapolated to zero in the zero-shear biphasic (I-N) regime at
381: $\phi\gae\phil\approx 33\%$. According to this extrapolation (which is
382: actually well beyond the final data point at $\phi\approx 22\%$), and
383: in the absence of a divergence in $G(\phi)$, the unscaled plateau
384: height $\sigmas$ must itself fall to zero at $\phi\approx 33\%$
385: (rightmost data point in Fig.~\ref{fig:coexistence}), consistent with
386: the behaviour of the concentrated systems discussed above. To
387: summarise, the plateau height $\sigmas$ appears to be a non-monotonic
388: function of concentration, increasing with $\phi$ through the studied
389: regime $5\%<\phi<20\%$ before (probably) falling to zero in the
390: zero-shear biphasic $I-N$ regime $\phi\approx 33\%$.
391:
392: Although this experiment showed that shear banding depends on the
393: {\em overall} concentration of the solution, there is relatively little
394: evidence for concentration-{\em coupling} (\ie\ concentration
395: differences between the bands) in such dilute systems, far from the
396: I-N transition. Indeed, the experiment just described revealed no
397: discernible upward slope in the coexistence plateau. We are not aware
398: of any measurements of concentration differences between the
399: coexisting shear bands in such systems. Nonetheless, recent
400: experiments on CTAB(0.3M)/${\rm NaNO}_3\rm {(1.79M)}$/H$_2$O at
401: $\phi=11\%$~\cite{LerDecBer00}
402: did reveal a stress plateau with slight upward slope. Along with the
403: generic expectation that flow should be coupled to concentration in these
404: viscoelastic solutions, this suggests that an explicit calculation of
405: concentration difference in the shear bands of systems far from an I-N
406: transition might be worthwhile.
407:
408: In this paper, therefore, we present the first such calculation, using
409: our \model\ model~\cite{FieOlm02,FieOlm02b}. In contrast to the work
410: of Olmsted \etal\ \cite{olmsted99c} for concentrated rigid rods,
411: shear banding in the \model\ model is not due to any underlying nematic
412: feature of the elastic free energy $F^{\rm e}(\tens{W})$. Instead the
413: instability results mainly from the non-linear effects of shear (the
414: intrinsic constitutive curve has a region of negative slope), though
415: it can be strongly enhanced by concentration coupling in systems close
416: to an underlying Cahn-Hilliard (CH) demixing instability (governed by
417: the osmotic free energy $F^{\rm o}(\phi)$). Indeed, the \model\ model
418: captures a broad crossover between (i) instabilities that are mainly
419: mechanical (governed by the negative slope of the flow curve) and (ii)
420: instabilities that are essentially CH demixing (governed by $F^{\rm
421: o}(\phi)$), but now triggered by shear. [Likewise, in practice
422: there should be no sharp distinction between concentrated micelles
423: with an underlying nematic feature in $F^{\rm e}(\tens{W})$ on the one
424: hand and ``non-nematic'' (more dilute) systems on the other: any more
425: refined model should allow a smooth crossover between the two cases.
426: This will be the focus of a future publication~\cite{FieOlm02f}.]
427:
428: \section{Model}
429: \label{sec:model}
430:
431: In this section we outline the \model\ model, which couples shear
432: banding instabilities to concentration in a simple way by combining
433: the non-local Johnson-Segalman (d-JS) model~\cite{johnson77} with a
434: 2-fluid framework~\cite{brochdgen77,milner93} for concentration
435: fluctuations. While this description is self-contained, readers
436: are referred to Ref.~\cite{FieOlm02b} for fuller details.
437:
438: \subsection{Free energy}
439:
440: In a sheared fluid, one cannot strictly define a free energy.
441: Nonetheless, for realistic shear rates, many internal degrees of
442: freedom of a polymeric solution relax quickly on the timescale of the
443: moving constraints and are therefore essentially equilibrated.
444: Integrating over these fast variables, one obtains a free energy for a
445: given fixed configuration of the slow variables. For our purposes, the
446: relevant slow variables are the fluid momentum and micellar
447: concentration $\phi$ (which are both conserved and therefore truly
448: slow in the hydrodynamic sense), and the micellar strain $\tens{W}$
449: that would have to be reversed in order to relax the micellar stress
450: (which is slow for all practical purposes):
451: %
452: \begin{equation}
453: {W}_{\alpha\beta}=\frac{\partial{R}'_{\alpha}}{\partial
454: \vect{R}}\cdot\frac{\partial {R}'_{\beta}}{\partial
455: \vect{R}}-\delta_{\alpha\beta}
456: \end{equation}
457: %
458: where $\delta\vect{R}'$ is the deformed vector corresponding to the
459: undeformed vector $\delta\vect{R}$.
460:
461: The resulting free energy is assumed to comprise separate
462: osmotic and elastic components,
463: %
464: \begin{equation}
465: \label{eqn:free_energy_total}
466: %F=F^{\rm K}(\vect{v})+F^{\rm o}(\phi)+F^{\rm e}(\tens{W},\phi).
467: F=F^{\rm o}(\phi)+F^{\rm e}(\tens{W},\phi).
468: \end{equation}
469: %
470: %The kinetic component is
471: %%
472: %\be
473: %\label{eqn:kinetic_fe}
474: %F^{\rm K}(\vect{v})=\tfrac{1}{2}\int d^3x \rho\vect{v}^2,
475: %\end{equation}
476: %
477: %where $\rho$ is the fluid density and $\vect{v}$ the velocity. The
478: The osmotic component is
479: %
480: \bea
481: \label{eqn:free_energy}
482: F^{\rm o}(\phi)&=&\int d^3x \left[f(\phi)+\frac{g}{2}(\nablu \phi)^2\right]\nonumber\\
483: &\approx& \tfrac{1}{2} \int d^3q\, (1+\xi^2 q^2)f''|\phi(q)|^2,
484: \end{eqnarray} %
485: where $f''$ is the osmotic susceptibility and $\xi$ is the
486: equilibrium correlation length for concentration fluctuations.
487: The elastic component is
488: %
489: \begin{equation}
490: \label{eqn:elastic_fe}
491: F^{\rm e}(\tens{W},\phi)=\tfrac{1}{2}\int d^3x\,G(\phi) {\rm tr} \left[\tens{W}-\log(\tens{\delta}+\tens{W})\right]
492: \end{equation}
493: %
494: in which $G(\phi)$ is the micellar stretching modulus.
495:
496: \subsection{Dynamics}
497:
498: The basic assumption of the two-fluid model is a separate force
499: balance for the micelles (velocity $\vm$; volume fraction $\phi$) and
500: the solvent (velocity $\vs$) within any element of solution.
501:
502:
503: The micellar force balance equation is:
504: %
505: \bw
506: \begin{equation}
507: \label{eqn:micelle}
508: \rho_{\rm m}\,\phi\,\left(\partial_t+\vm.\nablu\right)\vm=\nablu.G(\phi)\,\tens{W}-\phi\,\nablu \frac{\delta F(\phi)}{\delta\phi}+2\,\nablu.\,\phi\, \etam\, \Dm^{0} -\zeta(\phi)\,\vrel-\phi\nablu p.
509: \end{equation}
510: %
511: In this equation, $G(\phi)\tens{W}\equiv
512: 2G(\phi)\tens{W}.\tfrac{\delta F}{\delta\tens{W}}$ is the viscoelastic
513: micellar backbone stress due to deformation of the local molecular
514: strain, while the osmotic stress $\tfrac{\delta F^{\rm o}}{\delta
515: \phi}$ results from direct monomeric interaction. The Newtonian
516: stress $2\phi\, \etam\, \Dm^{0}$ describes fast micellar processes
517: (\eg\ Rouse modes) with $\Dm^{0}$ the traceless symmetric micellar
518: strain rate tensor. The force $\zeta\vrel$, where $\vrel=\vm-\vs$,
519: impedes relative motion; $\zeta$ is the drag coefficient
520: (Eq.~\ref{eqn:relative}). Incompressibility determines the pressure
521: $p$.
522:
523: Likewise, the solvent force balance comprises the Newtonian viscous
524: stress, the drag force (equal and opposite to the drag on the
525: micelles), and the hydrostatic pressure:
526: %
527: \begin{equation}
528: \label{eqn:solvent}
529: \rho_{\rm s}(1-\phi)\left(\partial_t+\vs.\nablu\right)\vs=2\nablu.\,(1-\phi)\,
530: \etas\, \Ds^{0}+\zeta(\phi)\vrel-(1-\phi)\nablu p.
531: \end{equation}
532: %
533: %The solvent stress $2(1-\phi)\, \etas\, \Ds^{0}$ is small
534: %for these systems: disregarding it does not change our results.
535:
536: Equations~\ref{eqn:micelle} and \ref{eqn:solvent} contain the basic
537: assumption of ``dynamical asymmetry'', \ie\ that the viscoelastic
538: stress acts only on the micelles and not on the solvent. Adding them,
539: and assuming equal mass densities $\rho_{\rm m}=\rho_{\rm
540: s}\equiv\rho$, we obtain the overall force balance equation for the
541: centre of mass velocity, $\vect{v}=\phi\vm+(1-\phi)\vs$:
542: %
543: \begin{equation}
544: \rho \left(\partial_t+\vect{v}.\nablu\right)\vv\equiv D_t\vect{v}
545: =\nablu.G(\phi)\,\tens{W}-\phi\,\nablu\frac{\delta
546: F(\phi)}{\delta\phi}+2\,\nablu.\,\phi\, \etam\,
547: \Dm^{0}+2\nablu.\,(1-\phi)\, \etas\, \Ds^{0} -\nablu p.
548: \label{eqn:navier}
549: \end{equation}
550: %
551: Subtracting them (with each predivided by its own volume fraction), we
552: obtain an expression for the relative motion $\vrel=\vm-\vs$, which in
553: turn specifies the concentration fluctuations:
554: %
555: \begin{equation}
556: D_t\phi=-\nablu\cdot\phi(1-\phi)
557: \vrel=-\nablu\cdot\frac{\phi^2(1-\phi)^2}{\zeta(\phi)}\left[\frac{\nablu
558: \cdot G(\phi)\tens{W}}{\phi}-\nablu\frac{\delta F}{\delta
559: \phi}+\frac{2\,\nablu\cdot\,\phi\, \etam\,
560: \Dm^{0}}{\phi}-\frac{2\nablu\cdot\,(1-\phi)\, \etas\,
561: \Ds^{0}}{1-\phi}\right]
562: \label{eqn:relative}
563: \end{equation}
564: %
565: which defines the micellar diffusion coefficient $D\equiv
566: f''(\phi)\phi^2(1-\phi)^2/\zeta$. We have omitted negligible inertial
567: corrections to Eqs.~(\ref{eqn:navier}) and (\ref{eqn:relative})
568: \cite{FieOlm02b}.
569:
570: %
571: The essence of the 2-fluid model is that the physically distinct
572: elastic and osmotic stresses appear together in the force-balance
573: equation~(\ref{eqn:navier}) and also in the generalised CH
574: equation~(\ref{eqn:relative}). This allows micellar diffusion in
575: response to gradients in concentration {\em and} in the viscoelastic
576: stress. We will see below that this gives rise to a positive HF
577: feedback between concentration and flow~\cite{HelfFred89}, allowing
578: micelles to diffuse up their own concentration gradient.
579:
580:
581:
582: For the dynamics of the viscoelastic micellar backbone strain we use
583: the phenomenological d-JS model~\cite{johnson77,olmsted99a}:
584: %
585: \begin{equation}
586: \label{eqn:JSd}
587: (\partial_t+\vm.\nablu)\tens{W}=a(\Dm.\tens{W}+\tens{W}.\Dm)+(\tens{W}.\Omm-\Omm.\tens{W})+2\Dm-\frac{\tens{W}}{\tau(\phi)}+\frac{l^2}{\tau(\phi)} \nablu^2 \tens{W},
588: \end{equation}
589: \ew
590: %
591: where $2\Omm=\nablu \vm - (\nablu \vm)^T$ with $(\nablu
592: \vm)_{\alpha\beta}\equiv \partial_{\alpha}(v_{\rm m})_\beta$.
593: $\tau(\phi)$ is the Maxwell time and $l$ is a length scale discussed
594: in Sec.~\ref{sec:interfaces} below. The slip parameter $a$
595: measures the non-affinity of the molecular deformation, \ie\ the
596: fractional stretch of the polymeric material with respect to that of
597: the flow field. For $|a|<1$ (slip) the intrinsic constitutive curve in
598: planar shear is capable of the non-monotonicity of
599: Fig.~\ref{fig:schem}.
600:
601: We use Eqns.~\ref{eqn:navier},~\ref{eqn:relative} and~\ref{eqn:JSd},
602: together with the incompressibility condition, $\nablu.\vect{v}=0$, as
603: our model for the remainder of the paper.
604:
605: \subsection{Flow geometry. Boundary conditions}
606: \label{sec:geometry}
607:
608: We consider idealised planar shear bounded by infinite plates at
609: $y=\{0,L\}$ with $(\vect{v},\nablu v, \nablu \wedge \vect{v})$ in the
610: $(\hat{\vect{x}},\hat{\vect{y}},\hat{\vect{z}})$ directions. We allow
611: variations only in the flow-gradient direction, and therefore set all
612: other derivatives to zero: $\partial_x\ldots=0$, $\partial_z\ldots=0$.
613: In appendix~\ref{app:fulleqns} we give all the relevant components of
614: the model equations~\ref{eqn:navier},~\ref{eqn:relative}
615: and~\ref{eqn:JSd} in this coordinate system.
616:
617: The boundary conditions at the plates are as follows. For the
618: velocity we assume there is no slip. For the concentration we assume
619: %
620: \begin{equation}
621: \label{eqn:BCconc}
622: \partial_y \phi=\partial^3_y \phi=0,
623: \end{equation}
624: %
625: which ensures (in zero shear at least) zero flux of concentration at
626: the boundaries.
627: %
628: Following Ref.~\cite{olmsted99a}, for the micellar strain we assume
629: %
630: \begin{equation}
631: \label{eqn:BCstrain}
632: \partial_y W_{\alpha\beta}=0\;\forall\; \alpha,\beta.
633: \end{equation}
634: %
635: Conditions~\ref{eqn:BCconc} and~\ref{eqn:BCstrain} together ensure
636: zero concentration flux at the boundary even in shear. For the
637: controlled shear rate conditions assumed throughout,
638: %
639: \begin{equation}
640: \label{eqn:constant_strain_rate}
641: \bar{\gdot}=\int_0^L dy \gdot(y)={\rm constant.}
642: \end{equation}
643: %
644:
645: \subsection{The interfacial terms}
646: \label{sec:interfaces}
647:
648: The model contains two different interfacial terms. The first is the
649: gradient term on the RHS of Eqn.~(\ref{eqn:JSd}). The length $l$ in
650: this term could, for example, be set by the mesh size or by the
651: equilibrium correlation length for concentration fluctuations. Here we
652: assume the former, since the dynamics of the micellar conformation are
653: more likely to depend on gradients in molecular conformation than in
654: concentration. Physically, one can interpret the gradient term in
655: equation~\ref{eqn:JSd} as resulting dynamically, from the diffusion of
656: stretched molecules across the interface~\cite{elkareh89}, or
657: statically, from nematic interactions between the micelles, or both.
658: There is, at present, no accepted theory for these gradient terms in
659: semi-dilute solutions. The equilibrium correlation length $\xi$ of
660: course still enters our analysis through our second interfacial term,
661: in the osmotic free energy of Eqn.~\ref{eqn:free_energy}.
662:
663: Together, $l$ and $\xi$ set the length scale of any interfaces.
664: Throughout this paper, we study the physical limit in which $l$ and
665: $\xi$ are small compared to the system size so that we have a sharp
666: interface connecting two bulk homogeneous phases. In this case, the
667: solution to Eqns.~\ref{eqn:navier},~\ref{eqn:relative}
668: and~\ref{eqn:JSd} naturally fits the zero-gradient boundary
669: conditions, and is invariant under $y\to y/2$, $l\to l/2$ and $\xi\to
670: \xi/2$. Therefore, a simultaneous reduction in $l$ and $\xi$ by the
671: same factor only changes the overall length of the interface, and not
672: the values of the order parameters in each phase (which determines
673: the phase diagram). However the phase diagram does depend slightly on
674: the ratio $r=l/\xi$: below we will give results for $r=0,r=\infty$ and
675: $r=O(1)$. This provides a concrete example of the early insight of Lu
676: and co-workers~\cite{lu99}, that the banded state must depend on the
677: nature of the interfacial terms. This contrasts notably with
678: equilibrium phase coexistence, in which the dynamical equations are
679: integrable and therefore insensitive to interfaces.
680:
681: \subsection{Model parameters}
682: \label{sec:parameters}
683:
684: \begin{table}
685: \begin{center}
686: \begin{tabular}{|p{2.85cm}|c|c|c|}
687: \hline
688: {\small\bf Parameter} & {\small \bf Symbol} $Q$ & {\small \bf Value at $\phi=0.11$} & $\frac{d\log Q}{d\log\phi}$ \\
689: \hline\hline
690: {\small Rheometer gap} & $L$ & $0.15\mbox{\,mm}$ & 0 \\
691: \hline
692: {\small Maxwell time} & $\tau$ & $0.17\mbox{\,s}$ & 1.1 \\
693: \hline
694: {\small Plateau modulus} & $G$ & $232\mbox{\,Pa}$ & 2.2 \\
695: \hline
696: {\small Density} & $\rho$ & $10^3\mbox{\,kg\,m}^{-3}$ & 0 \\
697: \hline
698: {\small Solvent viscosity} & $\etas$ & $10^{-3}\mbox{\,kg\,m}^{-1}\mbox{s}^{-1}$ & 0 \\
699: \hline
700: {\small Rouse viscosity} & $\etam$ & $0.4\mbox{\,kg\,m}^{-1}\mbox{s}^{-1}$ & 0 \\
701: \hline
702: {\small Mesh size} & $l$ & 2.6$\times 10^{-8}\mbox{m}$ & -0.73 \\
703: \hline
704: {\small Diffusion coefficient} & $D$ & $3.5 \times 10^{-11}\mbox{m}^2\mbox{s}^{-1}$ &0.77 \\
705: \hline
706: {\small Drag coefficient} & $\zeta$ & $2.4\times10^{12}\mbox{kg\,m}^{-3}\mbox{\,s}^{-1}$ & 1.54 \\
707: \hline
708: {\small Correlation length} & $\xi$ & $6.0\times10^{-7}$m & -0.77 \\
709: \hline
710: {\small Slip parameter} & $a$ & $0.92$ & 0 \\
711: \hline
712: \end{tabular}
713: \end{center}
714: \caption{Experimental values of the model's parameters at volume
715: fraction $\phi=0.11$ (column 3). Scaling laws for the dependence of
716: each parameter upon $\phi$ (column 4). In most calculations we use
717: the reference values of column 3 at $\phi=0.11$, then tune $\phi$
718: using the scaling laws of column 4. Only where stated do we allow
719: the parameters to vary independently. \label{table:parameters}}
720: \end{table}
721:
722: The \model\ model
723: (Eqns.~\ref{eqn:navier},~\ref{eqn:relative},~\ref{eqn:JSd}) has the
724: following parameters: the solvent viscosity $\etas$ and density
725: $\rho$; the plateau modulus $G$; the Maxwell time $\tau$; the Rouse
726: viscosity $\etam$; the mesh size $l$; the osmotic modulus $f''(\phi)$
727: and the equilibrium correlation length $\xi$ (recall
728: Eqn.~\ref{eqn:free_energy}); the drag coefficient $\zeta$ and the slip
729: parameter $a$. We also need to know the typical rheometer gap, $L$.
730: A reference set of parameter values at $\phi=0.11$ is summarised in
731: table~\ref{table:parameters}. These values were taken from experiment
732: or calculated using scaling arguments: see Ref.~\cite{FieOlm02b} for
733: details. Note that explicit data is not available for $f''(\phi)$;
734: however dynamic light scattering gives the diffusion coefficient
735: \begin{equation}
736: \label{eq:1}
737: D\equiv \frac{f''(\phi)\phi^2(1-\phi)^2}{\zeta(\phi)}.
738: \end{equation}
739: In this paper we will be guided by these parameter values, but subject
740: to the following considerations.
741:
742: First, we are only interested in steady states so for convenience can
743: take the limit of zero Reynolds number ($\rho=0$) and rescale the
744: kinetic coefficient $1/\zeta$ so that the diffusion time $L^2/D$ is of
745: order the Maxwell time. These choices have no effect on the steady
746: state, but make our numerical calculation of it much more efficient
747: (by evolving the {\em dynamical}
748: equations~\ref{eqn:navier},~\ref{eqn:relative} and~\ref{eqn:JSd}).
749: Second, realistic interfaces are much narrower than the typical
750: rheometer gap, with $l$ and $\xi$ both of $O(10^{-4}L)$. To resolve
751: such interfaces (allowing a minimal 10 numerical mesh points per
752: interface) would therefore require $O(10^5)$ grid points, while in
753: practice we are limited to $O(10^2)$. We will therefore use
754: artificially large values of $l$ and $\xi$. However this does not
755: affect the phase diagram, provided the interface is still small
756: compared with the gap size: see Sec.~\ref{sec:numerics} for details,
757: and Fig. \ref{fig:finiteSizeBeta} for a typical banded profile.
758: Finally, we artificially increase the Rouse viscosity $\etam$ by a
759: factor $50$ to ensure, again for numerical convenience, that the shear
760: rate of the high shear phase is not too large. This does
761: quantitatively change the phase diagram, but we checked that the
762: qualitative trends are not affected.
763:
764: Exploring this large parameter space is a daunting prospect so we
765: shall not, in general, vary the parameters independently of each
766: other. Instead we simply tune the concentration $\phi$, relying on
767: known semi-dilute scaling laws for the $\phi$-dependence of the other
768: parameters (column 4 of table~\ref{table:parameters}). However we
769: will, in separate $\phi-$sweeps, vary the degree of concentration
770: coupling, which is dictated by ratio of the elastic term
771: $\nablu.G(\phi)\tens{W}$ to the osmotic term $\nablu\frac{\delta
772: F}{\delta \phi}$ and which we encode in the parameter
773: \begin{equation}
774: \label{eq:2}
775: \alpha\equiv \frac{G'(\phi=0.11)}{2f''(\phi=0.11)}
776: \end{equation}
777: (where a prime denotes a derivative). In other $\phi-$sweeps we will
778: vary the characteristic interface widths $l(\phi=0.11)$ and
779: $\xi(\phi=0.11)$, to investigate any dependence of the phase diagram
780: on the ratio $l/\xi$ in the double limit $l/L\to 0$, $\xi/L\to 0$. In
781: what follows, we adopt the convenient shorthand of $l$ for
782: $l(\phi=0.11)$ with the understanding that $l$ does actually varies
783: with $\phi$ according to the scaling given in
784: table~\ref{table:parameters}. We do likewise for $\xi$.
785:
786: Throughout we rescale stress, time and length so that
787: $G(\phi=0.11)=1$, $\tau(\phi=0.11)=1$, and $L=1$.
788:
789: \section{Intrinsic flow curves; spinodals}
790: \label{sec:spinodals}
791:
792: %
793: \begin{figure}[tb]
794: \includegraphics[scale=0.3]{spinodals.eps}
795: \caption{Intrinsic flow curves (dotted lines) for
796: $\phi=0.11,0.10\ldots0.01$ (downwards). Spinodals for concentration
797: couplings $\alpha=10^{-2}, 10^{-3}, 10^{-4}$.
798: \label{fig:spinodals}}
799: \end{figure}
800: %
801:
802: The homogeneous intrinsic steady state flow curves
803: $\Sigma(\gdotb,\phib)=G(\phib)W_{xy}+\eta(\phib)\gdotb$ that satisfy
804: $\partial_t\vv=\partial_t\phi=\partial_t\tens{W}=0$ are shown as
805: dotted lines in Fig.~\ref{fig:spinodals}. (The average viscosity
806: $\eta(\phi)\equiv\phi\etam+(1-\phi)\etas$.) The region of negative
807: slope ends at a ``critical'' point $\phib_{\rm c}\approx 0.015$.
808: CPCl/NaSal in brine~\cite{berret94b} shows the same trend. For
809: completeness, in App.~\ref{app:homogen} we give analytical results for
810: the steady state conditions in homogeneous shear flow.
811:
812: In Ref.~\cite{FieOlm02,FieOlm02b} we linearised in fluctuations about
813: these homogeneous states to find the spinodal region in which the
814: homogeneous states are unstable. The spinodals are shown in
815: Fig.~\ref{fig:spinodals} for different levels concentration coupling,
816: $\alpha$.
817:
818: In the limit of zero concentration coupling $\alpha\to 0$,
819: fluctuations in the ``mechanical variables'', $\tens{W}$ and $\gdot$
820: decouple from those in concentration, and are unstable in the region
821: of negative constitutive slope, as expected. Separately, the
822: concentration could have its own Cahn-Hilliard demixing instability,
823: when the diffusion coefficient $D<0$; however we are interested only
824: in flow-induced instabilities and set $D>0$ throughout. For finite
825: $\alpha>0$, the region of mechanical instability is broadened by
826: coupling to the concentration fluctuations, as seen in
827: Fig.~\ref{fig:spinodals}. This can be understood as follows. Consider
828: the first term in the square brackets of Eqn.(\ref{eqn:relative}).
829: This causes micelles to move up gradients in the viscoelastic stress
830: $\tens{W}$, thereby increasing the concentration in stressed regions.
831: If $G'(\phi)>0$ (assumed here), the increased concentration causes the
832: stress to increase further, closing a positive HF~\cite{HelfFred89}
833: feedback loop whereby the micelles can diffuse {\em up} their own
834: concentration gradient.
835:
836: \section{Numerical details}
837: \label{sec:numerics}
838:
839: In this section, we outline our numerical procedure for solving the
840: dynamical equations~\ref{eqn:navier},~\ref{eqn:relative}
841: and~\ref{eqn:JSd} and discuss our careful study of time-step, mesh
842: size and finite size effects. Readers who are not interested in these
843: issues can skip this section.
844:
845: %
846: \begin{figure}[htb]
847: \includegraphics[scale=0.3]{finiteSizeBeta.eps}
848: \caption{Main figure: steady banded concentration profile at
849: $\alpha=10^{-2}$, $\xi=0$, $\gdotb=7.0$, $\phib=0.15$. Solid lines
850: $l=0.016$, for $(N_y,\dt)=(100,0.05),(200,0.0125),(400,0.0125)$,
851: dotted lines $l=0.008$ for $(N_y,\dt)=(200,0.05),(400,0.003125)$,
852: dashed lines $l=0.004$ for $(N_y,\dt)=(400,0.05),(800,0.05)$.
853: Upper inset, the same data, enlarged in the left hand phase
854: (decreasing $\phi$ with increasing $N_y$). Lower inset:
855: corresponding selected stresses, for the same parameter values and
856: mesh sizes (decreasing stress with increasing $N_y$).
857: \label{fig:finiteSizeBeta} }
858: \end{figure}
859: %
860:
861: We consider variations only in the flow gradient direction, in which we
862: discretise $y\in {0,1}$ on an algebraic grid $y_n=n/N_y$ for
863: $n=0,1\ldots N_y$. We stored $\phi$ and $\tens{W}$ on these grid points.
864: The velocities $\vm$ and $\vs$ were stored on half grid points
865: $y_{n+1/2}$, and we used linear interpolation between the half and
866: full grid points. Likewise we discretized time such that $t_n=n\Delta
867: t$. We evolved the discretized
868: equations~\ref{eqn:navier},~\ref{eqn:relative} and~\ref{eqn:JSd} using
869: the Crank-Nicholson algorithm which is semi-implicit in time, with
870: centred space derivatives
871:
872: %\bw
873: \begin{figure*}
874: \centering \subfigure[Tie lines (solid); spinodal (dashed).]{
875: \includegraphics[scale=0.33]{sigmaVsPhi_alpha1p0e-3_gamma0p0.eps}
876: \label{fig:alpha1.0e-3_gamma0.0:a}
877: }
878: \subfigure[Tie lines (solid); spinodal (dashed); homogeneous
879: constitutive curve (dotted); for $l=0.008$. ]{
880: \includegraphics[scale=0.33]{sigmaVsgdot_alpha1p0e-3_gamma0p0.eps}
881: \label{fig:alpha1.0e-3_gamma0.0:b}
882: } \subfigure[Tie lines (solid); spinodal (dashed) for $l=0.008$.]{
883: \includegraphics[scale=0.33]{gdotVsPhi_alpha1p0e-3_gamma0p0.eps}
884: \label{fig:alpha1.0e-3_gamma0.0:c}
885: } \subfigure[Coexistence regime of the macroscopic flow curves (solid);
886: spinodal (dashed); homogeneous constitutive curves (dotted)]{
887: \includegraphics[scale=0.3]{fc_alpha1p0e-3_gamma0p0_new.eps}
888: \label{fig:alpha1.0e-3_gamma0.0:d}
889: }
890: \caption{Phase diagrams and flow curves for $\alpha=10^{-2}$,
891: $\xi=0.0$ for small $l/L$. (Recall that $l$ is actually a function
892: of $\phi$: we are using the convenient shorthand of $l$ for the
893: value $l(\phi=0.11)$.) (a) Thin (upper) solid lines: tie lines for
894: $l=0.016,\ny=100,\dt=0.05$. Thick (lower) solid lines: tie lines for
895: $l=0.008,\ny=200,\dt=0.05$. As described in the main text, we
896: actually rescaled $l$ in the successive runs of each $\phib-$sweep
897: (\ie\ as $\phib$ was tracked from $0.15$ down to $\phic$) so that
898: the interfacial width remained (approximately) constant throughout
899: the sweep: the value of $l$ in the figure legends refers to the
900: value used in the first run of the sweep, at $\phib=0.15$. (b,c)
901: Solid lines: tie lines repeated in the $(\Sigma,\gdot)$,
902: $(\Sigma,\phi)$ representations for $l=0.008,\ny=200,\dt=0.05$. (d)
903: Solid lines: macroscopic flow curves for
904: $\phib=0.11,0.10,\ldots0.04$ (downward). These flow curves were
905: recontructed from the tie lines of the phase diagrams (using the tie
906: lines shown in this figure, and some additional ones). Because we
907: have only calculated tie lines for discrete values of $\Sigma$, in
908: some cases the reconstructed flow curves stop short of the
909: single-phase region, and have been continued by eye with a dashed
910: line. The inset in (d) shows the same data, but on a log-log plot.
911: The experimentally observed slope $0.3$ is marked as a dot-dashed
912: line for comparison. The spinodal is shown in each of Figs a-d as a
913: dashed line. In (b,d) the thin dotted lines are the intrinsic
914: (homogeneous) constitutive curves for $\phib=0.11,0.1\ldots0.01$
915: (downwards).}
916: \label{fig:alpha1.0e-3_gamma0.0}
917: \end{figure*}
918: %\ew
919:
920: For each run, we seeded an initial profile that was either homogeneous
921: up to a small random contribution, or inhomogeneous according to
922: $\phi=\bar{\phi}[1+\Delta\cos(\pi x)]$ (with $\Delta\approx 0.1$). We
923: then evolved the discretized
924: equations~\ref{eqn:navier},~\ref{eqn:relative} and~\ref{eqn:JSd} under
925: an imposed wall velocity until a steady banded state was reached. We
926: checked that the homogeneous phases between the interfaces were
927: insensitive to the initial conditions. However, for the random
928: initial condition several bands could form (and did not coarsen over
929: any accessible timescale). Therefore in most runs we used the
930: co-sinusoidal initial profile, to conveniently obtain just two bands
931: (as in Fig.~\ref{fig:profiles_compared}).
932:
933: For the dynamics to be independent of time-step, a very small
934: time-step has to be used. However the numerically-attained steady is
935: much less sensitive hence allowing much larger time-steps. A typical
936: steady state changes by less than $10^{-3}\,\%$ for a factor-two
937: reduction in timestep \ref{fig:finiteSizeBeta}. For the special case
938: of $\xi=0$, timesteps $\Delta t \propto \ny^{-2}$ can be used, since
939: the highest spatial derivative is second order. For $\xi\neq 0$, we
940: have a fourth order derivative in Eqn.~\ref{eqn:relative} and much
941: smaller timesteps $\Delta t\propto \ny^{-4}$ must be used.
942:
943: In all our calculations, we are interested in the physical limit where
944: the interface width is much smaller than the rheometer gap. This
945: creates a delicate balance, since narrow interfaces require a very
946: fine grid. Therefore we adopted the following procedure. For any
947: fixed value of the interfacial lengthscales $l$ and $\xi$, we
948: performed several runs with progressively finer meshes (but always
949: with a small enough time-step) until the shear banded profile and
950: selected stress didn't depend on the mesh. This is quite easy to
951: achieve: a typical steady state presented below changes by less than
952: $0.1\,\%$ upon doubling the number of mesh points. We then reduced $l$
953: and $\xi$ (in fixed ratio) until the order parameters in the
954: homogeneous phases changed by less than $0.5\%$ upon further halving
955: of $l$ and $\xi$ (but always ensuring convergence with respect to the
956: number of grid points). A sample study of these issues is presented
957: in Fig.~\ref{fig:finiteSizeBeta} for the special case $\xi=0$.
958:
959:
960: \section{Results}
961: \label{sec:results}
962:
963: We now present our results for the steady-state flow phase diagrams,
964: flow curves and shear banded profiles. Because one of our aims is to
965: show that the shear-banded state depends on the nature of the
966: interfacial terms, we consider three separate cases: (A) interfacial
967: terms only in the viscoelastic constitutive equation~\ref{eqn:JSd}
968: ($l\neq 0,\,\xi=0$, $r\equiv l/\xi=\infty$); (B) interfacial terms in
969: both the constitutive and concentration equations,~\ref{eqn:JSd}
970: and~\ref{eqn:relative} ($l\neq 0,\xi \neq 0$, $r=O(1)$); and (C)
971: interfacial terms only in the concentration
972: equation~\ref{eqn:relative} ($l=0,\xi\neq 0,r=0$).
973:
974: \subsection{Interfacial terms only in the viscoelastic constitutive
975: equation: $l\neq 0$, $\xi=0$.}
976: \label{sec:onlyvisco}
977:
978: In this section, we set the correlation length for concentration
979: fluctuations, $\xi$, to zero and consider small but non-zero values
980: the interfacial lengthscale $l$ in the constitutive
981: equation~\ref{eqn:JSd}.
982:
983: \subsubsection{Flow phase diagrams}
984:
985: \begin{figure}[htb]
986: \includegraphics[scale=0.3]{sigmaVsPhi_alpha1p0e-345_gamma0p0.eps}
987: \caption{Phase diagrams for three different degrees of coupling to
988: concentration for $\xi=0$ and small $l/L$.
989: \label{fig:sigmaVsPhi_alpha1.0e-345_gamma0.0} }
990: \end{figure}
991:
992: For any given shear banded profile, the values of the order parameters
993: in each of the two homogeneous phases specify the two ends of one tie line
994: in the phase diagram. Analogously to equilibrium tie lines, the
995: concentrations and strain rates of the coexisting states are related
996: to the mean strain rate $\bar{\dot{\gamma}}$ and mean concentration
997: $\bar{\phi}$ by the lever rule,
998: \begin{align}
999: \bar{\phi} &= \beta\phi_1 + (1-\beta)\phi_2 \\
1000: \bar{\dot{\gamma}} &= \beta\dot{\gamma}_1 + (1-\beta)\dot{\gamma}_2,
1001: \end{align}
1002: where $\beta$ is the volume fraction of material in state $(\phi_1,
1003: \dot{\gamma}_1)$. For each of several values of the concentration
1004: coupling, $\alpha$, we calculated the full phase diagram via a
1005: succession of shear startup runs, all at the critical shear rate
1006: $\gdotc(\alpha)$ (determined from Fig.~\ref{fig:spinodals}), for
1007: average concentrations ranging from $\phib=0.15$ down to the critical
1008: value $\phic(\alpha)$. For concentrations below the critical point
1009: the response of the system is smooth as a function of stress. In our
1010: model, this arises because decreasing concentration reduces the
1011: viscosity of the low shear rate branch faster than it reduces the
1012: viscosity of high shear rate branch. Hence the stress maximum
1013: decreases with decreasing concentration, disappearing when the stress
1014: maximum vanishes. Alternatively, in a more dilute system the plateau
1015: modulus and Maxwell time are both smaller, and one expects a smaller
1016: stress and higher strain rate at the onset of instability.
1017:
1018: The results for $\alpha=10^{-2}$, which gives rather strong
1019: concentration coupling, are shown in
1020: Fig.~\ref{fig:alpha1.0e-3_gamma0.0}a,b,c. Because the width,
1021: $\delta$, of the interface in the banded state is set by $l$, but with
1022: a prefactor that diverges at the critical point, in each successive
1023: run we rescaled $l$ so that $\delta$ remained (approximately) equal to
1024: its value ($\ll L$) in the first run at $\phib=0.15$. We return below
1025: to study the divergence of $\delta/l$ at the critical point.
1026:
1027: To illustrate the finite size considerations of
1028: Sec.~\ref{sec:numerics} (above), in
1029: Fig.~\ref{fig:alpha1.0e-3_gamma0.0:a} we show the tie lines obtained
1030: for two different (starting) values of $l$. All the results are
1031: converged with respect to mesh fineness and timestep (not explicitly
1032: shown), but the tie lines differ slightly between the two values of
1033: $l$. However all seem to be consistent with one given binodal line: we
1034: do not have any explanation for this apparent consistency.
1035:
1036: To investigate the effect of reducing the coupling to concentration,
1037: we repeat the phase diagram for $\alpha=10^{-2}$ alongside that for
1038: $\alpha=10^{-3}$ and $\alpha=10^{-4}$ in
1039: Fig.~\ref{fig:sigmaVsPhi_alpha1.0e-345_gamma0.0}. As expected, the
1040: concentration difference between the bands tends to zero as
1041: $\alpha\rightarrow0$.
1042: %$\alpha\equiv G''(\phi=0.11)/2f''(\phi=0.11)$.
1043:
1044: \subsubsection{Flow curves}
1045:
1046: So far, we have discussed the flow phase diagrams. Measurement of
1047: these diagrams still presents an open challenge to experimentalists,
1048: due to the difficulty in measuring the concentration of micelles in
1049: each band (although SANS data has been used to estimate the bands'
1050: concentrations in systems near the I-N transition~\cite{BRL98}). In
1051: this section we discuss the macroscopic flow curves, which are
1052: relativly easily measured using conventional bulk rheology. However
1053: it is important to realise that a set of flow curves
1054: $\Sigma(\gdotb,\phib)$ measured for {\em several} values of $\phib$
1055: actually contains the same information as the phase diagram:
1056: reconstruction of the latter from the former is described in
1057: Fig.~\ref{fig:reconstruct} A full set of flow curves could therefore
1058: be used to check measurements of concentration differences.
1059:
1060: \begin{figure}[htb]
1061: \includegraphics[scale=0.3]{calced_fc_alpha_1p0e-3_gamma0p0.eps}
1062: \caption{Macroscopic flow curves for $\alpha=10^{-2}$ at
1063: $\phib=0.11$. Thick solid line: reconstructed from the tie lines of
1064: the phase diagram. Dot dashed and dashed lines: calculated by
1065: directly measuring the average stress and strain rate during a
1066: strain rate sweep for $l=0.016,\ny=100,\dt=0.05$ (dot-dashed) and
1067: $l=0.008,\ny=200,\dt=0.05$ (dashed). (The slight discrepancy
1068: between these three curves is discussed in the text.) The thin
1069: dotted line is the intrinsic (homogeneous) constitutive curve.
1070: \label{fig:calced_fc_alpha_1.0e-3_gamma0.0} }
1071: \end{figure}
1072:
1073:
1074: \begin{figure}[htb]
1075: \includegraphics[scale=0.3]{fc_alpha1p0e-345_gamma0p0.eps}
1076: \caption{Macroscopic flow curves (from direct measurements of the
1077: stress and strain rate) for three different degrees of coupling to
1078: concentration.
1079: \label{fig:fc_alpha1.0e-345_gamma0.0} }
1080: \end{figure}
1081:
1082: %\bw
1083: \begin{figure*}
1084: \subfigure[]{
1085: \includegraphics[scale=0.19]{reconstruct1.eps}
1086: \label{fig:reconstruct:a}
1087: }
1088: \subfigure[]{
1089: \includegraphics[scale=0.8]{reconstruct2fig.eps}
1090: \label{fig:reconstruct:b}
1091: }
1092: \subfigure[]{
1093: \includegraphics[scale=0.8]{reconstruct3fig.eps}
1094: \label{fig:reconstruct:c}
1095: }
1096: \caption{Reconstruction of the flow phase diagram from a family of
1097: macroscopic flow curves $\Sigma(\gdotb,\phib)$, measured for several
1098: different average concentrations $\phib$. Consider the flow curves
1099: of Fig.~\ref{fig:reconstruct:a}. The curve that starts at A and ends
1100: at B is for an average concentration $\phib=0.08$. Points A and B
1101: are at the edge of the two-phase region. Reading off the stress
1102: from Fig.~\ref{fig:reconstruct:a}, A and B give use two points on
1103: the binodal in Fig.~\ref{fig:reconstruct:b}. Likewise reading off
1104: the strain rate, we get points A and B in
1105: Fig.~\ref{fig:reconstruct:c}. Repeating this for all the circles in
1106: Fig.~\ref{fig:reconstruct:a}, we can construct many points on the
1107: binodal in Figs.~\ref{fig:reconstruct:b} and
1108: \ref{fig:reconstruct:c}, which can then be interpolated over to give
1109: the full binodal. We now just need to specify the tie lines. In
1110: Fig.~\ref{fig:reconstruct:b} this is trivial: all tie lines are
1111: horizontal since the coexistence occurs at common stress (for
1112: gradient banding). In Fig.~\ref{fig:reconstruct:c}, to get the slope
1113: of the tie line that starts at $B$ we proceed by recalling that the
1114: tie line represents constant shear stress. Therefore we find another
1115: point, D, in Fig.~\ref{fig:reconstruct:a} that is at the same stress
1116: as point $A$, and read off its average strain-rate. Its average
1117: concentration is already known. This gives point $C$ in
1118: Fig.~\ref{fig:reconstruct:c}. Similarly, $D$ is the image of point
1119: $B$ at constant stress. Repeating this process we can fill in all
1120: the tie lines of the phase diagram. }
1121: \label{fig:reconstruct}
1122: \end{figure*}
1123: %\ew
1124:
1125: In this work, we take the opposite approach for convenience, and
1126: reconstruct the steady-state flow curves from the tie lines of the
1127: phase diagram. The results are shown in
1128: Fig.~\ref{fig:alpha1.0e-3_gamma0.0:d}. The inset shows the same data
1129: on a log-log plot, to enable comparison with Ref.\cite{berret94a} in
1130: which the coexistence plateau in a log-log representation was a
1131: reasonably straight line (over the shear-rate range investigated) with
1132: slope $0.3$. Note that the results shown in
1133: Fig.~\ref{fig:alpha1.0e-3_gamma0.0:d} are in units of $G(\phi=0.11)$
1134: and $\tau(\phi=0.11)$. In Ref.~\cite{BRP94}, Berret replotted the flow
1135: curves in units of $G(\phib)$ and $\tau(\phib)$, finding scaling
1136: collapse of the family $\Sigma(\gdotb,\phib)/G(\phib)$ \versus\
1137: $\gdotb\tau(\phib)$ in the low shear regime $\gdotb\to 0$. We do not
1138: find this scaling collapse
1139: (Fig.~\ref{fig:fc_alpha1.0e-3_gamma0.0_scaled}) because we have used
1140: an artificially large high-shear Newtonian contribution $\eta\gdot$
1141: for numerical convenience (recall Sec.~\ref{sec:parameters}): the
1142: overall zero shear viscosity, $G(\phi)\tau(\phi)+\eta(\phi)$ therefore
1143: does not scale as $G(\phi)\tau(\phi)$, even approximately.
1144:
1145: \begin{figure}[htb]
1146: \includegraphics[scale=0.3]{fc_alpha1p0e-3_gamma0p0_scaled.eps}
1147: \caption{Macroscopic flow curves as shown in
1148: Fig.~\ref{fig:alpha1.0e-3_gamma0.0:d} above, but now with the stress
1149: in units of $G(\phi)$ and the strain rate in units of $\tau(\phi)$.
1150: \label{fig:fc_alpha1.0e-3_gamma0.0_scaled} }
1151: \end{figure}
1152:
1153: To check the reconstruction of flow curves from the phase diagram, we
1154: also explicitly calculated the flow curve at a single $\phib=0.11$. To
1155: do this, we first performed a shear startup at a given $\gdotb$ in the
1156: unstable region. We then (without reinitialising the system)
1157: decreased $\gdotb$ in steps to the edge of the coexistence regime,
1158: ensuring that a steady state was reached before measuring the total
1159: stress. We then reinitialised the system and repeated the entire
1160: procedure, but now with increasing $\gdotb-$jumps. The results are
1161: shown in \ref{fig:calced_fc_alpha_1.0e-3_gamma0.0} for two different
1162: values of $l$. The slight discrepancy between the directly measured
1163: flow curve ``plateaus'' (i.e. the inhomogeneouse part of the flow
1164: curve) and those reconstructed from the tie lines is due to the finite
1165: size of the interface $\delta$ relative to the cell $L$, and so is
1166: smaller for the smaller value of $\delta/L$. The construction
1167: described in Fig.~\ref{fig:reconstruct} implicitly assumes that
1168: $\delta/L=0$.
1169:
1170:
1171: As expected for this value of $\alpha$ (which gives a large
1172: concentration difference between the bands;
1173: Fig.~\ref{fig:alpha1.0e-3_gamma0.0:a}), the steady state flow curve
1174: ``plateau'' slopes strongly upwards in $\gdotb$. In
1175: Fig.~\ref{fig:fc_alpha1.0e-345_gamma0.0} we compare the (directly
1176: measured) macroscopic flow curve for the three levels of concentration
1177: coupling shown in Fig.~\ref{fig:sigmaVsPhi_alpha1.0e-345_gamma0.0}: as
1178: expected, the slope of the flow curve tends to zero with the degree of
1179: concentration coupling $\alpha$.
1180:
1181: The upturn in the measured flow curve at the edge of the coexistence
1182: plateau (apparent at the lower binodal for $\alpha=10^{-2}$ in
1183: Fig.~\ref{fig:fc_alpha1.0e-345_gamma0.0}) results again from the
1184: finite value of $\delta/L$: the interface bumps into the edge of the
1185: rheometer when one of the bands gets very narrow. We expect this
1186: (steady-state) effect to be much less pronounced in experimental
1187: systems, since realistic interfaces are much smaller than those used
1188: in our numerical study. Only near a critical point, where the
1189: interface becomes very broad (for fixed $l$), would we expect to see a
1190: true steady-state bump at the edge of the plateau. Nonetheless,
1191: pronounced bumps are often apparent in data obtained via upward
1192: strain-rate sweeps. However in most cases this is likely to be a
1193: metastable effect, so that the bump could be eliminated (or at least
1194: reduced) by reducing the rate of the sweep~\cite{grand97}.
1195:
1196: As noted in Sec.~\ref{sec:intro}, in a curved Couette geometry the
1197: ``plateau'' (B'F' of Fig.~\ref{fig:schem}) in the flow curve will
1198: slope upwards due to the inhomogeneity of the stress field, even
1199: without concentration coupling. It should be noted that all
1200: calculations in this paper are for a planar shear geometry, and the
1201: slope of our flow curves in the coexistence regime results solely from
1202: concentration coupling. In fact, the slope in
1203: Fig.~\ref{fig:alpha1.0e-3_gamma0.0:d} is far greater than one would
1204: typically expect from curvature effects: for a Couette cell with
1205: radius $R$ and gap $\delta R$, the stress measured at the inner
1206: Couette wall would change by $\delta\Sigma/\Sigma=2\delta R/R$ over
1207: the coexistence regime, and so too would the relative change in torque
1208: through the coexistence ``plateau''. The slope of
1209: Fig.~\ref{fig:alpha1.0e-3_gamma0.0:d} would therefore require an
1210: atypically large curvature of $\delta R /R\sim0.5$.
1211:
1212: \subsubsection{Interfacial profiles; divergence of interface width at
1213: the critical point}
1214:
1215: \begin{figure}[htb]
1216: \includegraphics[scale=0.4]{profiles_compared.eps}
1217: \caption{Steady state shear banded profile at $\alpha=10^{-2}$,
1218: $\gdotb=4.64$, $\phib=0.15$ for two different ratios $r=l/\xi$. The
1219: thick lines are for $r=\infty$ ($l=0.008$, $\xi=0.0$), $\ny=200$,
1220: $\dt=0.05$, as considered in this section. The thin lines show the
1221: corresponding results for $r=0.4$ ($l=0.008$, $\xi=0.002$),
1222: $\ny=200$, $\dt=0.00625$ (to be discussed in Sec.~\ref{sec:both}
1223: below), for comparison.
1224: \label{fig:profiles_compared} }
1225: \end{figure}
1226:
1227: \begin{figure}[htb]
1228: \includegraphics[scale=0.3]{widths_alpha1p0e-3_gamma0p0.eps}
1229: \caption{Scaled interface width $\delta/l$ versus the distance from the critical stress $\Sigma-\Sigma_{\rm c}$. The dotted line is a power $-0.5$.
1230: \label{fig:widths_alpha1.0e-3_gamma0.0} }
1231: \end{figure}
1232:
1233: We now turn to the interfacial profiles and widths. A full steady
1234: state banded profile for $\alpha=10^{-2}$ (corresponding to the
1235: rightmost/uppermost tie line in
1236: fig.~\ref{fig:alpha1.0e-3_gamma0.0}a,b,c) is shown by the thick lines
1237: in Fig.~\ref{fig:profiles_compared}. As required, the interface is
1238: smooth on the scale of the mesh, but sharp on the scale of the gap
1239: size, \ie\ $L/\ny\ll \delta \ll L\equiv 1$ where $\delta$ is the width
1240: of the interface. Note that the shear rate is negative across the gap
1241: since we have chosen to move the wall at $y=0$; accordingly we have
1242: plotted $-W_{xy}$, since $W_{xy}$ is antisymmetric in shear rate.
1243: $-W_{wy}$ is rather small in the high shear band, as expected from the
1244: underlying constitutive non-monotonicity. Meanwhile $W_{xx}$ is very
1245: large, while $W_{yy}\approx -0.5$ (recall that $\tens{W}$ measures
1246: deformation relative to the unit tensor $\tens{\delta}$): this
1247: corresponds to the micelles being highly stretched along the flow
1248: direction and is consistent with the experimental observation that the
1249: first normal stress difference progressively increases throughout the
1250: banding regime~\cite{rehage91}. The concentration is lower in the high
1251: shear band, where $W_{yy}$ is smaller (more negative): this is a
1252: direct result of the tendency of mielles to move up gradients in
1253: $W_{yy}$, as determined by Eqn.~\ref{eqn:relative} above.
1254:
1255: In fact the interface width, $\delta$, is slightly different for each
1256: order parameter: we define it to be the distance between the two
1257: points where the change in that order parameter between the two
1258: homogeneous phases is $25\%$ and $75\%$ complete. For a fixed value
1259: of $l$ (which sets the overall scale of the interface width), $\delta$
1260: diverges at the critical point (for each order parameter). In
1261: tracking $\phib$ down towards the critical point, therefore, we
1262: continually rescaled $l$ to ensure that the interface width remained
1263: approximately constant. In each case, we measured $\delta/l$, for each
1264: of $W_{xy}$, $W_{xx}$, $W_{yy}$ and $\phi$: see
1265: Fig.~\ref{fig:widths_alpha1.0e-3_gamma0.0}. According to mean field
1266: theory, the divergence should be of the form $\delta/l \sim
1267: (\Sigma-\Sigma_{\rm c})^{-1/2}$. The power $-1/2$ is accordingly shown
1268: in Fig.~\ref{fig:widths_alpha1.0e-3_gamma0.0} as a guide for the eye.
1269:
1270: \subsection{Interfacial terms in both the viscoelastic constitutive
1271: equation, and in the concentration equation: $l\neq 0$, $\xi\neq
1272: 0$.}
1273: \label{sec:both}
1274:
1275: %\afterpage{\clearpage}
1276:
1277: %\bw
1278: \begin{figure}[htb]
1279: \centering \subfigure[Tie lines]{
1280: \includegraphics[scale=0.32]{gdotVsPhi_alpha1p0e-3_betaonly_and_both.eps}
1281: \label{fig:alpha1.0e-3_betaonly_and_both:a}
1282: } \subfigure[Partially reconstructed macroscopic flow curves
1283: (solid); homogeneous constitutive curve (dotted).]
1284: {\includegraphics[scale=0.32]{fc_alpha1p0e-3_betaonly_and_both.eps}
1285: \label{fig:alpha1.0e-3_betaonly_and_both:b}
1286: }
1287: \caption{Phase diagrams and flow curves for $\alpha=10^{-2}$ and
1288: $r=0.4$ ($l=0.008$, $\xi=0.02$) with the corresponding data for
1289: $\alpha=10^{-2}$ and $r=\infty$ ($l=0.008$, $\xi=0$) for
1290: comparison. }
1291: \label{fig:alpha1.0e-3_betaonly_and_both}
1292: \end{figure}
1293: %\ew
1294:
1295: We now study the effect of including interfacial gradient terms in the
1296: concentration equation~\ref{eqn:relative} (so that now $\xi \neq 0$)
1297: as well as in the viscoelastic equation~\ref{eqn:JSd}, $l\neq 0$.
1298: Hence, while in the previous section we considered $r\equiv
1299: l/\xi=\infty$, then, we now consider $r=O(1)$. In
1300: Fig.~\ref{fig:alpha1.0e-3_betaonly_and_both:a}, we give the phase
1301: diagram for $r=0.4$. Comparing it with our results for $r=\infty$
1302: (also shown in Fig.~\ref{fig:alpha1.0e-3_betaonly_and_both:a}), we see
1303: that the slopes of the tie lines and the overall binodal both depend
1304: quantitatively on $r$. [The difference between the results for
1305: $r=\infty$ and $r=0.4$ is far greater than any ``error'' associated
1306: with the fact that we are not quite in the limit $\dt\to 0$,
1307: $l\ny\to\infty $, $\xi\ny\to\infty $, $l/L\to 0$ and $\xi/L \to 0$.]
1308: This provides a concrete example of the fact that shear-banding
1309: coexistence is determined by, and non-universal with respect to, the
1310: interfacial terms~\cite{lu99}. As noted above, this contrasts sharply
1311: with the equilibrium case, in which the equations of motion are
1312: integrable and so the phase diagram is independent of the interfacial
1313: terms. Although conceptually important, this dependence is in
1314: practice rather weak: the overall features of the phase diagram are
1315: unchanged. The critical point is unaffected.
1316:
1317:
1318: In Fig.~\ref{fig:alpha1.0e-3_betaonly_and_both:b} we show the
1319: corresponding macroscopic flow curves, reconstructed using the tie
1320: lines of Fig.~\ref{fig:alpha1.0e-3_betaonly_and_both:a}. Because we
1321: only calculated a few tie lines in this case, the recontruction is
1322: rather sparse. Nonetheless, the slight difference between $r=\infty$
1323: and $r=0.4$ is apparent.
1324:
1325: In Fig.~\ref{fig:profiles_compared}, we compare a full banded profiles
1326: for $r=\infty$ and $r=0.4$. The slight dependence on $r$ is again
1327: apparent.
1328:
1329: %\newpage
1330:
1331: \subsection{Interfacial gradient terms only in the concentration equation: $l=0$, $\xi\neq 0$.}
1332: \label{sec:onlyconc}
1333:
1334: %\bw
1335: \begin{figure*}
1336: \centering \subfigure[Tie lines, only shown for the case $l=r=0$
1337: when close to the critical point]{
1338: \includegraphics[scale=0.28]{gdotVsPhi_gammonly_and_both.eps}
1339: \label{fig:gdotVsPhi_gammonly_and_both}
1340: }
1341: \subfigure[Steady-state profiles for initial condition
1342: $\phi(y)=\phib+0.4\cos(\pi y)$: there is no selected smoothly banded
1343: state ]{ \includegraphics[scale=0.33]{profiles_gammonly.eps}
1344: \label{fig:profiles_gammonly}
1345: }\\
1346: \subfigure[Steady-state profiles for initial conditions
1347: $\phi(y)=\phib+0.4\cos(\pi y)$ and $\phi(y)=\phib+0.7\cos(\pi y)$: the
1348: steady state depends on the initial condition.]
1349: {
1350: \includegraphics[scale=0.25]{no_selection.eps}
1351: \label{fig:no_selection}
1352: }
1353: \caption{(a) Phase diagram at $\alpha=10^{-2}$, for $l=0.0$,
1354: $\xi=0.02$, $r\equiv l/\xi=0.0$, shown with the corresponding data
1355: for $l=0.008$, $\xi=0.02$, $r=0.4$ for comparison. Tie lines are
1356: only shown near the critical point because for larger values of
1357: $\phib$, there is no uniquely selected, smoothly banded state. This
1358: is shown in Figs. b and c. In Fig.b the steady state profiles from
1359: left to right at fixed ordinate correspond to tie lines left to
1360: right in the upper Fig.(a). Fig (c) shows the steady state profile
1361: in $\gdot$ (upper two curves) and in $\phi$ (lower two curves) for
1362: $\phib=0.16$ and $\gdotb=4.66$ with initial condition
1363: $\phi(y)=\phib+0.4\cos(\pi y)$ (solid lines) and with
1364: $\phi(y)=\phib+0.7\cos(\pi y)$ (dashed lines): the ``selected''
1365: state depends upon the initial condition -- \ie\ there is no state
1366: selection for $l=0$ for stresses far enough above the critical
1367: point.}
1368: \label{fig:alpha1.0e-3_gammonly}
1369: \end{figure*}
1370: %\ew
1371:
1372: %\afterpage{\clearpage}
1373:
1374: Finally we set the interfacial length $l$ in the constitutive equation
1375: equal to zero. The constitutive equation is now local, and the only
1376: source of spatial gradients is the equilibrium correlation length for
1377: concentration fluctuations (Eqns.~\ref{eqn:relative}
1378: and~\ref{eqn:free_energy}): $r\equiv l/\xi=0$. In the absence of
1379: concentration coupling, $\alpha=0$, it is known that there is no
1380: uniquely selected, smoothly shear banded state when $l=0$~\cite{lu99}.
1381: Here we investigate whether a smoothly banded state is selected for
1382: $\alpha\neq 0$, by virtue of the interfacial terms in the
1383: concentration equation.
1384:
1385: Our numerics only gave a smoothly banded profile for stresses near the
1386: critical point, even for the largest accessible values of $\xi$ and
1387: $\ny$. The profiles shown from left to right in
1388: Fig.~\ref{fig:profiles_gammonly} are progressively further above the
1389: critical point. The tie lines correponding to the smooth profiles near
1390: the critical point are shown in
1391: Fig.~\ref{fig:gdotVsPhi_gammonly_and_both}, alongside the correponding
1392: results at $r=0.4$ for comparison. Consistent with the discussion of
1393: non-universality in the previous section, the phase diagram for
1394: $r=0.0$ is slightly different from that for $r=0.4$ (and is different
1395: again from the case $r=\infty$; not shown).
1396:
1397: For the spiky profiles, further from the critical point, the binodal
1398: of the associated tie lines is irregular (not shown in
1399: Fig.~\ref{fig:gdotVsPhi_gammonly_and_both}), suggesting that the
1400: steady state is not uniquely selected. In view of this, a natural
1401: question is whether selection could occur in principle (but is
1402: inaccessible with any realistic mesh due to the pronounced
1403: non-monotonicity in $W_{xy}(y)$), or whether selection cannot occur,
1404: even in principle. In Fig.~\ref{fig:no_selection} we show that the
1405: steady state depends on the initial condition; so state selection
1406: appears to be lost when $l=0$. This numerical observation is backed up
1407: by the following analytical argument.
1408:
1409: In steady state, the system must obey:
1410: %
1411: \begin{itemize}
1412: \item The force-balance equation,
1413: \begin{equation}
1414: \label{eqn:simplenavier}
1415: S(\gdot,\phi)\equiv G(\phi)W_{xy}[\gdot\tau(\phi)]+\etabar(\phi)\gdot=\Sigma={\rm const.}
1416: \end{equation}
1417: %
1418: \item The (now local) constitutive equation, equation~(\ref{eqn:JSd}),
1419: \begin{equation}
1420: \label{eqn:solveW}
1421: W_{\alpha\beta}=W_{\alpha\beta}[\gdot \tau(\phi)] \;\;\mbox{for }\;\;\alpha\beta=xx,xy,yy.
1422: \end{equation}
1423: %
1424: \item The steady-state of equation~\ref{eqn:relative}. For the
1425: purposes of this analytical argument we use a simplified version of
1426: this equation, which we believe still captures the essential physics:
1427: %
1428: \begin{equation}
1429: \label{eqn:simpleconc}
1430: 0=\partial^2_y\left\{\Lambda-\partial_y^2\phi \right\}
1431: \end{equation}
1432: %
1433: with
1434: %
1435: \begin{equation}
1436: \label{eqn:defineF}
1437: \Lambda=f'(\phi)-\frac{G(\phi)W_{yy}[\gdot\tau(\phi)]}{\phi}.
1438: \end{equation}
1439: %
1440: %$F$ can be written as either a single-valued function of $\gdot$
1441: %(Fig.~\ref{fig:construction}b), or a multi-valued function of $\phi$
1442: %(Fig.~\ref{fig:construction}c).
1443: Integrating Eqn.~\ref{eqn:simpleconc}
1444: twice, and using the boundary conditions
1445: $\partial_y\phi=0,\partial^3_y\phi=0$ for $y=0,L$, we obtain
1446: %
1447: \begin{equation}
1448: \label{eqn:third}
1449: \mu={\rm const.}=\Lambda-\partial_y^2\phi,
1450: \end{equation}
1451: where $\mu$ is an integration constant.
1452: %
1453: \end{itemize}
1454: %
1455: We now show that a solution satisfying Eqns.~\ref{eqn:simplenavier}
1456: and~\ref{eqn:solveW} cannot in general simultaneously satisfy
1457: Eqn.~\ref{eqn:third}.
1458:
1459: %
1460: \begin{figure*}[htb]
1461: \includegraphics[scale=0.6]{construction1.eps}
1462: \caption{a) Dotted lines: relation between $\phi$ and $\gdot$ for the
1463: case of a local constitutive equation, for several values of the
1464: shear stress, $\Sigma$, calculated using
1465: Eqns.~(\ref{eqn:simplenavier}) and~(\ref{eqn:solveW}). Solid lines:
1466: the results of our numerics, showing that the $\phi-$ profile cannot
1467: properly negotiate the interface, as described in the main text.
1468: (Note that the stresses used to generate the dotted and the solid
1469: lines differ slightly, but the overall trend is still clear.) b) The
1470: function $\Lambda$ of Eqn.~\ref{eqn:defineF} plotted \versus\
1471: $\gdot$ using the relation of Fig.a). c) $\Lambda$ replotted
1472: \versus\ $\phi$.
1473: \label{fig:construction} }
1474: \end{figure*}
1475: %
1476: Consider firstly Eqns.~\ref{eqn:simplenavier} and~\ref{eqn:solveW}.
1477: Sustituting $W_{xy}$ from Eqn.~\ref{eqn:solveW} into
1478: Eqn.~\ref{eqn:simplenavier}, we obtain an expression for
1479: $S(\gdot,\phi)$: this is just the family of homogeneous constitutive
1480: curves, as plotted in Fig.~\ref{fig:spinodals} above. Because the
1481: constitutive equation is local, the solution at all points across the
1482: rheometer cell must lie on one of these intrinsic constitutive curves.
1483: Indeed, as the shear rate changes across the interface, the system
1484: must pass through constitutive curves of differing concentrations to
1485: maintain a uniform stress $\Sigma$.
1486: %($\partial_y\Sigma=0$ by force balance.)
1487: In other words, a relation $\phi=\phi(\gdot,\Sigma)$ must be
1488: obeyed. The family of these curves is shown as dotted lines in
1489: Fig.~\ref{fig:construction}a.
1490: For the range of stresses at which $\phi(\gdot,\Sigma)$ is
1491: non-monotonic, $\phi$ must have the form shown in
1492: Fig.~\ref{fig:proof}b in which the derivative $\partial^2_y\phi$
1493: changes sign three times across the interface, as in
1494: Fig.~\ref{fig:proof}c. [Actually, the forms of
1495: Fig.~\ref{fig:proof}b,c assume that the profile in $\gdot$ increases
1496: monotonically through the interface (Fig.~\ref{fig:proof}a). However
1497: this monotonicity will emerge self consistently from our argument
1498: below.]
1499:
1500: However we know from Eqn.~\ref{eqn:third} that
1501: $\partial_y^2\phi=\Lambda-\mu$. $\Lambda$ is plotted in
1502: Fig.~\ref{fig:construction}b,c using Eqn.~\ref{eqn:defineF} together
1503: with the constraint $\phi=\phi(\gdot,\Sigma)$ (imposed from
1504: Eqns.~\ref{eqn:simplenavier} and~\ref{eqn:solveW}, as discussed
1505: above). From this plot we see that, for any $\mu$, a solution that
1506: starts and ends in homogeneous phases (for which
1507: $\partial_y^2\phi=\Lambda-\mu=0$)~\footnote{While this boundary
1508: condition is intuitively clear (the homogeneous phases are in
1509: general large compared with the interface), in our numerics we only
1510: actually imposed the boundary condtion $\partial_y\phi=0$. We
1511: therefore did not, a priori, ``overspecify'' the differential
1512: equation~\ref{eqn:third}. However we shall show below that
1513: $\partial^2_y\phi$ must {\em automatically} equal zero at each
1514: interface.} can only involve at most one sign change of
1515: $\partial_y^2\phi$ between the boundaries. This inconsistency with
1516: Fig.~\ref{fig:proof}c means that a steady banded solution cannot exist
1517: for these stress values for which $\phi(\gdot,\Sigma)$ is
1518: non-monotonic. To summarize: for stresses far enough above the
1519: critical point that $\phi(\gdot,\Sigma)$ is non-monotonic, a steady
1520: state solution cannot simultaneously satisfy
1521: Eqns.~\ref{eqn:simplenavier} and~\ref{eqn:solveW} (which imply three
1522: sign changes of $\partial_y^2\phi$) at the same time as
1523: Eqn.~\ref{eqn:third} (which only allows one sign change). Therefore
1524: there a steady, smoothly banded profile cannot exist for such
1525: stresses.
1526:
1527: This argument is consistent with the sharp numerical profiles of
1528: Fig.~\ref{fig:profiles_gammonly}, which are replotted in
1529: Fig.~\ref{fig:construction}a (solid lines): each solution should have
1530: followed a local (dotted) curve $\phi(\gdot)$, but instead has jumped
1531: across the region in which this curve is non-monotonic. (The stresses
1532: used to generate the homogeneous solutions $\phi(\gdot,\Sigma)$ in
1533: Fig.~\ref{fig:construction}a were slightly different from those of the
1534: numerical profiles: however the trend is still clear.) Although
1535: Eqn.~(\ref{eqn:simpleconc}) is highly oversimplified, we believe that
1536: the failure to negotiate the interface due to the conflict described
1537: above is the reason for non-selection in the full, numerically solved
1538: model.
1539:
1540: %
1541: \begin{figure}[htb]
1542: \includegraphics[scale=0.6]{proof.eps}
1543: \caption{Assuming that the shear rate varies monotonically across the
1544: interface (a), then for a relation $\phi(\gdot,\Sigma)$ of
1545: Fig.~\ref{fig:construction}a that is non-monotonic, the
1546: concentration $\phi$ must vary as in b), with three sign changes in
1547: $\phi''\equiv\partial_y^2\phi$ as in c).
1548: \label{fig:proof} }
1549: \end{figure}
1550: %
1551:
1552: We return finally to justify our assumption that the shear rate must
1553: increase monotonically through the interface, and to discuss in more
1554: detail the nature of the banded solution when it {\em can} exist (\ie\
1555: for stresses near crticial point where $\phi(\gdot)$ is monotonic).
1556: Multiplying Eqn.~\ref{eqn:third} across by $d\phi/dy$, integrating on
1557: $\phi$, and imposing $\partial_y\phi=0$ at each boundary, we find,
1558: {\em for the simplified model of Eqn.~\ref{eqn:simpleconc}},
1559: %
1560: \begin{equation}
1561: \label{eqn:equalareas}
1562: \int_{\phi_l}^{\phi_r}d\phi\,\left[\mu-\Lambda\right]=0,
1563: \end{equation}
1564: %
1565: which is an ``equal areas'' construction. ($\phi_l$ and $\phi_r$
1566: denote the boundary values at $y=0,L$.)
1567: %
1568: \begin{figure}[htb]
1569: \includegraphics[scale=0.3]{equalareas.eps}
1570: \caption{Of these three proposed constructions specifying the banded
1571: state (when it is selected, near the critical point), only a) and b)
1572: are consistent with the boundary conditions $\partial_y\phi=0$. Of
1573: these, b) is for a finite system for which $\partial_y^2\phi\neq 0$
1574: at the boundary while a) is for the realistic physical limit in
1575: which the interface is narrow compared with the gap size,
1576: connecting two homogeneous phases in which
1577: $\partial_y^n\phi=0\,\forall\,n$.
1578: \label{fig:equalareas} }
1579: \end{figure}
1580: %
1581: If, in addition, we were to impose that $\partial_y^2\phi=0$ at each
1582: boundary, then the construction must automatically be as shown in
1583: Fig.~\ref{fig:equalareas}a. However we did not actually impose this
1584: condition in our numerics, so the construction of
1585: Fig.~\ref{fig:equalareas}b is also possible. This in fact corresponds
1586: to a finite system, where the true homogeneous state
1587: $\partial_y^n\phi=0\,\forall n$ is not quite reached at the
1588: boundaries. Any other equal areas construction
1589: (Fig.~\ref{fig:equalareas}c) is not possible, for the following
1590: reason. Consider starting at point C with $\partial_y\phi=0$ (which
1591: we {\em do} impose at the boundary in our numerics).
1592: Eqn.~\ref{eqn:third} then tells us that $\partial_y^2\phi<0$ at this
1593: point, so the function $\phi(y)$ must curve downwards from its
1594: starting point of zero slope. Therefore $\phi$ locally decreases, and
1595: the system moves to point $C'$. Repeating this argument, we find that
1596: the system can never cross to the point $D$. By similar reasoning, the
1597: shear rate must rise monotonically through the interface since any
1598: initial fall (from the side of the low shear band) would be similarly
1599: unstable to point $C$ in Fig.~\ref{fig:equalareas}c above.
1600:
1601: Of course the concentration equation~(\ref{eqn:simpleconc}) is highly
1602: oversimplified. For instance, a more realistic model (such as the one
1603: of Eqn.~\ref{eqn:relative}) would have $\phi$ dependent prefactors to
1604: the $\partial_y^2\phi$ term. {\em The equal areas result of
1605: Eqn.~(\ref{eqn:equalareas}) is therefore specific to our
1606: oversimplified Eqn.~(\ref{eqn:simpleconc}), and does not hold in
1607: general.} Nonetheless we believe that Eqn.~\ref{eqn:simpleconc}
1608: correctly predicts the absence of a uniquely banded solution for
1609: stresses far above the critical point, via the basic conflict between
1610: the number of sign changes of $\partial_y^2\phi$ across the interface,
1611: described above.
1612:
1613: \section{Conclusions}
1614: \label{sec:conclusion}
1615:
1616: In this paper, we have studied the role of concentration coupling in
1617: the shear banding of complex fluids using the two-fluid, non-local
1618: Johnson-Segalman model. We have calculated phase diagrams for
1619: different degrees of coupling between concentration and mechanical
1620: degrees of freedom (molecular strain), and found a phase diagram
1621: qualitatively consistent with experiments on micellar solutions at
1622: dilutions well below the equilibrium isotropic-to-nematic transition
1623: \cite{berret94b}. Specific points to note are as follows.
1624:
1625: \begin{enumerate}
1626:
1627: \item The coexistence plateau in the steady-state flow curve slopes
1628: upward with shear rate, because of the concentration difference
1629: between the coexisting bands. The overall plateau height and width
1630: decrease with average concentration, terminating in a
1631: non-equilibrium critical point. CPCl/NaSal in
1632: brine~\cite{berret94b} shows the same trend.
1633:
1634: \item
1635:
1636: Of the two coexisting bands, the high shear band has a smaller
1637: concentration due to the fact that concentration tends to move up
1638: gradients in the normal micellar strain component $W_{yy}$ (where
1639: $y$ is the flow-gradient direction). ($\tens{W}$ describes
1640: deformation relative to the unit tensor $\tens{\delta}$, and
1641: $W_{yy}$ is more negative in the high-shear phase than in the lower
1642: shear phase.) Tie lines of the phase diagram in the $\gdot,\phi$
1643: plane therefore have negative slope.
1644:
1645: \item
1646: The concentration gap is smaller for smaller values of
1647: concentration-coupling $\alpha\propto G'(\phi)/f''(\phi)$, and
1648: tends to zero in the limit $\alpha\to 0$. Accordingly, the
1649: coexistence region of the steady-state flow curve becomes flat in
1650: this limit.
1651:
1652: \item We have described the way in which the flow phase diagram can be
1653: reconstructed from the family of flow curves $\Sigma(\gdotb,\phib)$,
1654: measured for several average concentrations $\phib$
1655: (Fig.~\ref{fig:reconstruct}).
1656:
1657: \item The phase diagram and flow curves depend slightly on the
1658: relative size of the interfacial term in the viscoelastic
1659: constitutive equation to that in the equation that specifies the
1660: concentration dynamics. This is a concrete demonstration of how
1661: stress selection and the coexistence conditions of driven systems
1662: depend on the nature of the interface, in contrast to equilibrium
1663: coexistence.
1664:
1665: \item We find \textit{no} unique state selection when there are no
1666: gradient terms in the viscoelastic constitutive equation, except for
1667: stresses that are close to the critical point. This implies that,
1668: for a model to reproduce a uniquely selected stress, it is not
1669: enough to simply have gradient terms only in, for example, the
1670: concentration dynamics. The dynamical equations of motion for each
1671: degree of freedom must possess inhomogeneous terms to attain
1672: selection in all situations. Conversely, in situations where such
1673: terms are physically absent, one can expect, under certain
1674: conditions, no selection and hence a range of control parameters
1675: (shear stress or strain rate) for which the steady states are
1676: \textit{intrinsically} history-dependent.
1677:
1678: \item The interface width diverges at the critical point as a power
1679: law $(\Sigma-\Sigma_{\rm c})^{-n}$ with $n\approx 0.5$, although
1680: $n$ differs slightly across the different order parameters.
1681:
1682: \end{enumerate}
1683:
1684: Although our \model\ model is highly oversimplified, we believe that
1685: it contains the basic ingredients required for a first description of
1686: wormlike micellar surfactant solutions at concentrations well below
1687: the isotropic-nematic (I-N) transition. In particular, it incorporates
1688: the minimal set of realistic degrees of freedom (tensorial order
1689: parameter for the micellar strain together with concentration), and
1690: unifies a non-monotonic flow curve with the Helfand-Fredrickson
1691: coupling between concentration and flow. Similar techniques could be
1692: applied to more involved Cates non-linear theory for wormlike micelles
1693: \cite{cates90,SpenCate94}.
1694:
1695: We recall a previous calculations by Olmsted \etal\ was aimed at
1696: systems of rigid rods near the I-N transition~\cite{olmsted99c}. In
1697: future work we hope to unify these two approaches into a description
1698: of wormlike micelles that is valid over the entire concentration
1699: range. This should provide a first step towards understanding the
1700: crossover regime in the data of Fig.~\ref{fig:coexistence}, in which the
1701: coexistence plateau stress is a non monotonic function of the micellar
1702: concentration.
1703:
1704: \begin{acknowledgements}
1705: We thank J-F Berret and Paul Callaghan for useful discussions, and
1706: EPSRC GR/N11735 for financial support.
1707: \end{acknowledgements}
1708:
1709: \appendix
1710: \bw
1711:
1712: \section{ \model\ Equations in Cartesian Coordinates}
1713: \label{app:fulleqns}
1714:
1715: In this appendix, we give the components of the \model\ model's
1716: equations for planar shear flow along the $x$ direction, allowing
1717: gradients only in the flow-gradient direction, $y$, as described in
1718: Sec.~\ref{sec:geometry}, above. The $x$ component of force-balance is
1719: (in the zero-Reynolds limit considered in this paper)
1720: %
1721: \be
1722: 0=\partial_y
1723: \left[ G \left( \phi \right) { W_{xy}} \right] +\,{ \etam}\,\partial_y
1724: \left[\,\phi\,\partial_y { v_{mx}} \right] +\,{ \etas}\,\partial_y
1725: \left[ \, \left( 1-\phi \right) \partial_y { v_{sx}} \right]. \ee
1726: %
1727: The $y$ component of force-balance is fixed by incompressibility,
1728: $\nablu.\vect{v}=0$, along with the boundary condition $v_y=0$:
1729: %
1730: \be
1731: 0=\phi v_{my}+(1-\phi)v_{sy}.
1732: \ee
1733: %
1734: The relative velocity between the micelles and solvent (again ignoring
1735: inertial terms) is
1736: %
1737: \begin{align}
1738: v_{my}-{ v_{sy}}&=\frac{\phi(1-\phi)}{\zeta} \left\{ \frac{1}{\phi}
1739: \partial_y \left[ G \left( \phi
1740: \right) { W_{yy}} \right] +2\,\frac{1}{\phi} { \etam}
1741: \,\partial_y \left[ \,\phi\,\partial_y { v_{my}} \right] -2\, \frac{1}{1-\phi} { \etas}\,\partial_y \left[ \, \left( 1-\phi
1742: \right) \partial_y { v_{sy}} \right] - \partial_y \mathfrak{F} \right\}\\
1743: { v_{mx}}-{ v_{sx}}&=\frac{\phi(1-\phi)}{\zeta}\left\{ \frac{1}{\phi}
1744: \partial_y \left[ G \left( \phi
1745: \right) { W_{xy}} \right] +\,\frac{1}{\phi} { \etam}
1746: \,\partial_y \left[ \,\phi\,\partial_y { v_{mx}} \right] - \,\frac{1}{1-\phi} { \etas}\,\partial_y \left[ \, \left( 1-\phi
1747: \right) \partial_y { v_{sx}} \right]\right\}\\
1748: \mathfrak{F}&= { f'} \left( \phi \right) - \,g \partial^2_y \phi + \tfrac{1}{2}\,{ G'}
1749: \left( \phi \right) \left[ { W_{yy}}+{ W_{xx}}-\ln \left( { W_{yy}}
1750: \,{ W_{xx}}+{ W_{yy}}+{ W_{xx}}+1-{{ W_{xy}}}^{2} \right) \right].
1751: \end{align}
1752: The evolution of the micellar strain tensor is given by
1753: \begin{align}
1754: \partial_t W_{xy}+{ v_{my}}\,\partial_y { W_{xy}}&= \tfrac{1}{2}(a-1) W_{xx} \,\partial_y {
1755: v_{mx}} + \tfrac{1}{2}(1+a)\,W_{yy} \partial_y { v_{mx}} +a
1756: \,W_{xy} \partial_y { v_{my}} + \,\partial_y {
1757: v_{mx}}
1758: -{\frac {{ W_{xy}}}{\tau \left( \phi \right) }}+{\frac {
1759: l \left( \phi \right) ^{2} \partial^2_y { W_{xy}}
1760: }{\tau \left( \phi \right) }},\\
1761: \partial_t W_{yy}+{ v_{my}}\,\partial_y { W_{yy}} &= (a-1)
1762: \,W_{xy}\partial_y {
1763: v_{mx}} +2\,a \,W_{yy} \partial_y { v_{my}} + 2\,\partial_y { v_{my}} -{\frac {{
1764: W_{yy}}}{\tau \left( \phi \right) }}+{\frac { l \left( \phi
1765: \right) ^{2} \partial^2_y { W_{yy}} }{
1766: \tau \left( \phi \right) }},\\
1767: \partial_t W_{xx}+{ v_{my}}\,\partial_y { W_{xx}} &= (1+a)\,W_{xy}\partial_y {
1768: v_{mx}} -{
1769: \frac {{ W_{xx}}}{\tau \left( \phi \right) }}+{\frac { l
1770: \left( \phi \right) ^{2}\partial^2_y { W_{xx}}
1771: }{\tau \left( \phi \right) }}.
1772: \end{align}
1773: Finally, the concentration dynamics are
1774: %
1775: \bea
1776: \partial_t\phi&=&-\partial_y \left\{ \frac{\phi^2\, \left( 1-\phi
1777: \right)^2}{\phi\zeta}
1778: \left\{ \partial_y \left[ G \left( \phi
1779: \right) { W_{yy}} \right] +2\,\etam \,\partial_y \left[
1780: \,\phi\,\partial_y { v_{my}}
1781: \right] -\frac{2\phi}{1-\phi} {
1782: \etas}\,\partial_y \left[ \, \left( 1-\phi \right) \partial_y
1783: { v_{sy}}\right] - \partial_y \mathfrak{F} \right\} \right\}. \eea
1784:
1785: \ew
1786:
1787: \section{Stationary homogeneous solutions of the \model\ model}
1788: \label{app:homogen}
1789:
1790: In planar shear, the stationary homogeneous solutions to
1791: Eqns.~(\ref{eqn:navier}-\ref{eqn:JSd}) for given $\gdot$ and $\phi$
1792: are $\vrel\equiv\vm-\vs=0$ and
1793: %
1794: \begin{subequations}
1795: \label{eqn:intrinsic_flow_curve}
1796: \begin{align}
1797: W_{xy}&=\frac{\gdot\tau(\phi)}{1+b\gdot^2\tau^2(\phi)},\\
1798: W_{yy}&= \frac{a-1}{1+a}W_{xx}=
1799: -\frac{1}{(1+a)}\frac{b\gdot^2}{1+b\gdot^2}\\
1800: W_{zz}&=W_{xz}=W_{yz}=0,
1801: \end{align}
1802:
1803: \end{subequations}
1804: %
1805: where $b=1-a^2$. The steady state shear stress is given by
1806: \begin{equation}
1807: \label{eq:3}
1808: \Sigma_{xy}=G(\phib)W_{xy}+\phi\etam+(1-\phi)\etas\gdot=\textrm{constant}.
1809: \end{equation}
1810:
1811: %\bibliographystyle{epjsty}
1812:
1813: %\bibliography{ackerson,actin,articles,banding,barham,berrportdecruppe,books,callaghan,chandcits,crystal,crystal_theory,dnatheory,flowcryst,fredrickson,gelbart,helfrich,hsiao,LCtheory,lifshitz,line_tension,malkus,master,mccoy,membs,mine,noirez,notes,onions,pineweitz,pomeau,poon,pratt,psolutions,rheofolks,ryan,schoot,semenov,sriram,stein,worms2,worms3,worms,zubarev}
1814:
1815: \begin{thebibliography}{10}
1816:
1817: \bibitem{cates90}
1818: M.~E. Cates, J.~Phys. Chem. {\bf 94}, 371 (1990).
1819:
1820: \bibitem{SpenCate94}
1821: N.~A. Spenley and M.~E. Cates, Macromolecules {\bf 27}, 3850 (1994).
1822:
1823: \bibitem{SCM93}
1824: N.~A. Spenley, M.~E. Cates, and T.~C.~B. McLeish, Phys. Rev. Lett. {\bf 71},
1825: 939 (1993).
1826:
1827: \bibitem{Yerushalmi70}
1828: J. Yerushalmi, S. Katz, and R. Shinnar, Chemical Engineering Science {\bf 25},
1829: 1891 (1970).
1830:
1831: \bibitem{olmsted99a}
1832: P.~D. Olmsted, O. Radulescu, and C.-Y.~D. Lu, J. Rheology {\bf 44}, 257
1833: (2000).
1834:
1835: \bibitem{lu99}
1836: C.-Y.~D. Lu, P.~D. Olmsted, and R.~C. Ball, Phys. Rev. Lett. {\bf 84}, 642
1837: (2000).
1838:
1839: \bibitem{olmstedlu97}
1840: P.~D. Olmsted and C.-Y.~D. Lu, Phys.~Rev. {\bf E56}, 55 (1997).
1841:
1842: \bibitem{spenley96}
1843: N.~A. Spenley, X.~F. Yuan, and M.~E. Cates, J.~Phys.~II (France) {\bf 6}, 551
1844: (1996).
1845:
1846: \bibitem{berret94b}
1847: J.~F. Berret, D.~C. Roux, and G. Porte, J.~Phys.~II (France) {\bf 4}, 1261
1848: (1994).
1849:
1850: \bibitem{Call+96}
1851: P.~T. Callaghan, M.~E. Cates, C.~J. Rofe, and J.~B. A.~F. Smeulders, J. Phys.
1852: II (France) {\bf 6}, 375 (1996).
1853:
1854: \bibitem{grand97}
1855: C. Grand, J. Arrault, and M.~E. Cates, J.~Phys.~II (France) {\bf 7}, 1071
1856: (1997).
1857:
1858: \bibitem{MairCall96}
1859: R.~W. Mair and P.~T. Callaghan, Europhys. Lett. {\bf 36}, 719 (1996).
1860:
1861: \bibitem{MairCall96c}
1862: R.~W. Mair and P.~T. Callaghan, Europhysics Letters {\bf 65}, 241 (1996).
1863:
1864: \bibitem{BritCall97}
1865: M.~M. Britton and P.~T. Callaghan, Phys. Rev. Lett. {\bf 78}, 4930 (1997).
1866:
1867: \bibitem{berret94a}
1868: J.~F. Berret, D.~C. Roux, G. Porte, and P. Lindner, Europhys. Lett. {\bf 25},
1869: 521 (1994).
1870:
1871: \bibitem{schmitt94}
1872: V. Schmitt, F. Lequeux, A. Pousse, and D. Roux, Langmuir {\bf 10}, 955
1873: (1994).
1874:
1875: \bibitem{Capp+97}
1876: E. Cappelaere, J.~F. Berret, J.~P. Decruppe, R. Cressely, and P. Lindner, Phys.
1877: Rev. {\bf E} {\bf 56}, 1869 (1997).
1878:
1879: \bibitem{rehage91}
1880: H. Rehage and H. Hoffmann, Mol. Phys. {\bf 74}, 933 (1991).
1881:
1882: \bibitem{Decr+95}
1883: J.~P. Decruppe, R. Cressely, R. Makhloufi, and E. Cappelaere, Coll. Polym. Sci.
1884: {\bf 273}, 346 (1995).
1885:
1886: \bibitem{Makh+95}
1887: R. Makhloufi, J.~P. Decruppe, A. Aitali, and R. Cressely, Europhys. Lett. {\bf
1888: 32}, 253 (1995).
1889:
1890: \bibitem{DCC97}
1891: J.~P. Decruppe, E. Cappelaere, and R. Cressely, J. Phys. II (France) {\bf 7},
1892: 257 (1997).
1893:
1894: \bibitem{BPD97}
1895: J.~F. Berret, G. Porte, and J.~P. Decruppe, Phys. Rev. {\bf E} {\bf 55}, 1668
1896: (1997).
1897:
1898: \bibitem{FisCal01}
1899: E. Fischer and P.~T. Callaghan, Phys.\ Rev.\ E {\bf 6401}, 1501 (2001).
1900:
1901: \bibitem{FisCal00}
1902: E. Fischer and P.~T. Callaghan, Europhys.\ Lett. {\bf 50}, 803 (2000).
1903:
1904: \bibitem{LerDecBer00}
1905: S. Lerouge, J.~P. Decruppe, and J.~F. Berret, Langmuir {\bf 16}, 6464 (2000).
1906:
1907: \bibitem{BRL98}
1908: J.~F. Berret, D.~C. Roux, and P. Lindner, European Physical Journal B {\bf 5},
1909: 67 (1998).
1910:
1911: \bibitem{schmitt95}
1912: V. Schmitt, C.~M. Marques, and F. Lequeux, Phys.~Rev. {\bf E52}, 4009 (1995).
1913:
1914: \bibitem{brochdgen77}
1915: F. Brochard and P.-G. de~Gennes, Macromolecules {\bf 10}, 1157 (1977).
1916:
1917: \bibitem{HelfFred89}
1918: E. Helfand and G.~H. Fredrickson, Phys. Rev. Lett. {\bf 62}, 2468 (1989).
1919:
1920: \bibitem{DoiOnuk92}
1921: M. Doi and A. Onuki, J. Phys. II (France) {\bf 2}, 1631 (1992).
1922:
1923: \bibitem{milner93}
1924: S.~T. Milner, Phys.~Rev. {\bf E48}, 3674 (1993).
1925:
1926: \bibitem{WPD91}
1927: X.~L. Wu, D.~J. Pine, and P.~K. Dixon, Phys. Rev. Lett. {\bf 66}, 2408
1928: (1991).
1929:
1930: \bibitem{beris94}
1931: A.~N. Beris and V.~G. Mavrantzas, J. Rheol. {\bf 38}, 1235 (1994).
1932:
1933: \bibitem{SBJ97}
1934: T. Sun, A.~C. Balazs, and D. Jasnow, Phys. Rev. {\bf E} {\bf 55}, R6344
1935: (1997).
1936:
1937: \bibitem{tanak96}
1938: H. Tanaka, Phys. Rev. Lett. {\bf 76}, 787 (1996).
1939:
1940: \bibitem{DecLerBer01}
1941: J.~P. Decruppe, S. Lerouge, and J.~F. Berret, Phys.\ Rev.\ E {\bf 6302}, 2501
1942: (2001).
1943:
1944: \bibitem{FieOlm02}
1945: S.~M. Fielding and P.~D. Olmsted, (2002), preprint cond-mat/0207344.
1946:
1947: \bibitem{FieOlm02b}
1948: S.~M. Fielding and P.~D. Olmsted, (2002), preprint, cond-mat/0208599.
1949:
1950: \bibitem{johnson77}
1951: M. Johnson and D. Segalman, J. Non-Newt. Fl. Mech {\bf 2}, 255 (1977).
1952:
1953: \bibitem{milner91}
1954: S.~T. Milner, Phys. Rev. Lett. {\bf 66}, 1477 (1991).
1955:
1956: \bibitem{deGen76}
1957: P.-G. de~Gennes, Macromolecules {\bf 9}, 587 (1976).
1958:
1959: \bibitem{Brochard83}
1960: F. Brochard, J.Phys. (Paris) {\bf 44}, 39 (1983).
1961:
1962: \bibitem{olmsted99c}
1963: P.~D. Olmsted and C.-Y.~D. Lu, Phys. Rev. {\bf E60}, 4397 (1999).
1964:
1965: \bibitem{FieOlm02f}
1966: S.~M. Fielding and P.~D. Olmsted, (2002), in preparation.
1967:
1968: \bibitem{elkareh89}
1969: A.~W. El-Kareh and L.~G. Leal, J. Non-Newt. Fl.~Mech. {\bf 33}, 257 (1989).
1970:
1971: \bibitem{BRP94}
1972: J.~F. Berret, D.~C. Roux, and G. Porte, J. Phys. II (France) {\bf 4}, 1261
1973: (1994).
1974:
1975: \end{thebibliography}
1976:
1977:
1978: \end{document}
1979:
1980:
1981: