nlin0205062/rama6.tex
1: \documentclass[aps,twocolumn]{revtex4}
2: \usepackage{epsfig}
3: \usepackage{bm}
4: %\input REVTeXDraft1.tex
5: \def\B.#1{{\bm{#1}}}
6: \def\C.#1{{\cal{#1}}}
7: \def\pop#1{{{\partial}\over{\partial#1}}}
8: \def\popa#1#2{{{\partial#1}\over{\partial#2}}}
9: \def\p#1#2{\partial#1/\partial#2}
10: \def\i{{\rm i}}
11: \def\d{{\rm d}} 
12: \def\e{{\rm e}} 
13: \def\D{{\rm D}} 
14: \def\invr{O\left(R^{-1}\right)}
15: \def\siminvr{\sim O\left(R^{-1}\right)}
16: \def\abs#1{\left|#1\right|}
17: \def\ep{{\epsilon}}
18: \def\be{\begin{equation}}
19: \def\ee{\end{equation}}
20: \def\etal{{\em et al.} }
21: \def\R{{\rm Re}} 
22: \def\k{k} 
23: \def\Rec {${\rm R}_{\rm cr}$}
24: \def\s{\left(k_+^2 + \beta^2\right)}
25: \def\A{\left[\i (\omega_+ - k_+ U) + \nu(D^2 - k_+^2 - \beta^2) +(D\nu)
26: D\right]}
27: \def\DA{\left[-\i k_+ (DU) + (D\nu)(D^2 - k_+^2 - \beta^2) + (D^2\nu)
28: D\right]}
29: \def\beta{k_z}
30: 
31: \begin{document} 
32: 
33: \title{Stabilization of Hydrodynamic Flows by Small Viscosity Variations}
34: \author{Rama Govindarajan$^{\dag}$, Victor S. L'vov$^*$ and Itamar
35: Procaccia$^*$}
36: \affiliation{$^\dag $ Fluid Dynamics Unit, Jawaharlal Nehru Centre for
37: Advanced 
38: Scientific Research, Jakkur, Bangalore 560064, India. \\
39: $^*$ Dept. of Chemical Physics, The Weizmann Institute of Science, Rehovot
40: 76100, Israel. }
41: %%%%%%%%%%%%%%%%%%%%%%
42:  
43: \begin{abstract}  
44: Motivated by the large effect of turbulent drag reduction by minute
45: concentrations of polymers we study the effects of a weakly
46: space-dependent viscosity on the stability of hydrodynamic flows. In a
47: recent Letter [Phys. Rev. Lett. {\bf 87}, 174501, (2001)] we exposed
48: the crucial role played by a localized region where the energy of
49: fluctuations is produced by interactions with the mean flow (the
50: ``critical layer"). We showed that a layer of weakly space-dependent
51: viscosity placed near the critical layer can have a very large
52: stabilizing effect on hydrodynamic fluctuations, retarding
53: significantly the onset of turbulence.  In this paper we extend these
54: observation in two directions: first we show that the strong
55: stabilization of the primary instability is also obtained when the
56: viscosity profile is realistic (inferred from simulations of turbulent
57: flows with a small concentration of polymers).  Second, we analyze the
58: secondary instability (around the time-dependent primary instability)
59: and find similar strong stabilization. Since the secondary instability
60: develops around a time-dependent solution and is three-dimensional,
61: this brings us closer to the turbulent case. We reiterate that the
62: large effect is {\em not} due to a modified dissipation (as is assumed
63: in some theories of drag reduction), but due to reduced energy intake
64: from the mean flow to the fluctuations. We propose that similar
65: physics act in turbulent drag reduction.
66: \end{abstract}
67: 
68: \maketitle
69: 
70: \section{Introduction}
71:  
72: This paper is motivated by the dramatic effects that are observed with
73: the addition of small amounts of polymers to hydrodynamic flows.
74: While interesting effects were discussed in the context of the
75: transition to turbulence, vortex formation and turbulent transport
76: \cite{95NH}, the phenomenon that attracted the most attention was, for
77: obvious reasons, the reduction of friction drag by up to 80\% when
78: very small concentrations of long-chain polymers were added to
79: turbulent flows \cite{69Lum,00SW}. In spite of the fact that the
80: phenomenon is robust and the effect huge, there exists no accepted
81: theory that can claim quantitative agreement with the experimental
82: facts. Moreover, it appears that there is no mechanistic
83: explanation. In the current theory that is due to de Gennes
84: \cite{86TG,90Gennes} one expects the Kolmogorov cascade to be
85: terminated at scales larger than Kolmogorov scale, leading somehow to
86: an increased buffer layer thickness and reduced drag, but how this
87: happens and what is the fate of the turbulent energy is not being made
88: clear.
89: 
90: In a recent Letter \cite{01GLP} we proposed that the crucial issue is
91: in the {\em production} of energy of hydrodynamic fluctuations by
92: their interaction with the mean flow. For the sake of concreteness we
93: examined a Poiseuille laminar flow and its loss of linear stability,
94: and showed how small viscosity contrasts lead to an order of magnitude
95: retardation in the onset of instability of ``dangerous"
96: disturbances. Specifically, we considered a flow in a channel of
97: dimensionless width 2, in which there are two fluids: one fluid of
98: viscosity $\mu_1$ flows near the walls, and the other fluid of
99: viscosity $\mu_2$ flows at the center, see Fig. \ref{f:scheme}. The
100: viscosities differ slightly, for example we considered (in
101: dimensionless units) $\mu_2=1$ and $m=\mu_1/\mu_2 =0.9$. The main
102: ingredient of the calculation was that all the viscosity difference of
103: 0.1 concentrated in a ``mixed" layer of width 0.10. The motivation
104: behind these numbers was the observation that the inferred effective
105: viscosity in polymer drag reduction increases towards the center by
106: about 30\% over about a 1/3 of the half-channel \cite{97SBH}. With our
107: choice we have comparable viscosity gradients in the mixed layer.
108: 
109: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
110: \begin{figure} 
111: \epsfxsize=7.5cm \epsfbox{Ramaj1.eps}
112: \caption{Schematic of the flow: the fluid near the walls has a viscosity
113: $\mu_1$, and that flowing at the center is of viscosity $\mu_2$. In
114: the mixed layer (of width $q$) the viscosity varies gradually between
115: $\mu_1$ and $\mu_2$. The parameter $p$ controls the position of the
116: mixed layer. For simplicity we neglect the down-stream growth in $q$.}
117: \label{f:scheme} 
118: \end{figure} 
119: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
120: 
121: In this model everything was explicitly calculable. The main point of
122: our analysis (see Sect. \ref{primary} for further details) was that
123: there exists a position in the channel where the velocity of the mean
124: flow is the same as the velocity of the most dangerous primary
125: instability. Below we refer to the layer around this position as the
126: ``critical layer". If we placed the mixed layer in the vicinity of the
127: critical layer, we got a giant effect of stabilization.  Analyzing
128: this phenomenon, we demonstrated that nothing special happened to the
129: dissipation.  Rather, it was the energy intake from the mean flow to
130: the unstable mode that was dramatically reduced, giving rise to a
131: large effect for a small cause. In this paper we extend these
132: observations in two directions. In Sect. \ref{primary}, after
133: reviewing the results of the simple model, we extend the analysis of
134: the primary instability to a case in which the viscosity profile is
135: that inferred from direct numerical simulations of turbulent channel
136: flow of a dilute polymeric solution \cite{97SBH}. We will see that
137: very similar effects are found. In other words, one does not need to
138: put by hand the region of viscosity variation in the vicinity of the
139: critical layer. When we have a continuous variation of the viscosity
140: in the region near the wall, the effect is the same, since it is only
141: crucial that there will be {\em some} space dependence of the
142: viscosity in the critical layer, which is usually not too far from the
143: wall.
144: 
145: A possible criticism of our results can be that the primary
146: instability is still too far from typical turbulent fluctuations. This
147: is in particular true since the most unstable primary modes are
148: 2-dimensional, whereas typical turbulent fluctuations are
149: 3-dimensional. For these reasons we present in Sect. \ref{secondary}
150: the analysis of the effect of small viscosity variations on the
151: secondary instability, for which the most ``dangerous" modes are
152: 3-dimensional. The tactics are similar to those taken for the primary
153: instability. First we discuss the effects of a mixed layer put at the
154: ``right" place in the channel, and second we show that continuous
155: viscosity profiles do exactly the same. We find again the giant effect
156: of stabilization for relatively small viscosity variations, lending
157: further support to our proposition that similar effects may very well
158: play a crucial role in turbulent drag reduction.  In Sect.
159: \ref{conclude} we present concluding remarks and suggestions for the
160: road ahead.
161: 
162: \section{Primary instability of Poiseuille flow}
163: \label{primary}
164: 
165: It is well known that parallel Poiseuille flow loses linear stability
166: at some threshold Reynolds number Re=Re$_{\rm th}$ (close to 5772).
167: It is also well known that the instability is ``convective", with the
168: most unstable mode having a phase velocity $c_p$. Analytically it has
169: the form \be \hat \phi_p(x,y,t) = \frac{1}{2}\big\{\phi_p(y) \exp\left
170: [\i \k_p(x-c_p \, t) \right ] + \mbox{c.c.}\big\} \exp(\gamma_p t) \ ,
171: \label{eq:inst} 
172: \ee where the subscript $p$ stands for the primary instability, $\hat
173: \phi(x,y,t)$ is the disturbance streamfunction and $\phi(y)$ is the
174: complex envelope of $\hat \phi(x,y,t)$. We have chosen $x$ and $y$ as
175: the streamwise and wall-normal coordinates respectively, $\k$ as the
176: streamwise wavenumber of the disturbance and $t$ as time. $\gamma_p$
177: is the growth rate of the primary instability.  What is not usually
178: emphasized is that the main interactions leading to the loss of
179: stability occur in a sharply defined region in the channel, i.e. at
180: the critical layer whose distance from the wall is such that the phase
181: velocity $c$ is identical to the velocity of the mean flow somewhere
182: within this layer. It is thus worthwhile to examine the effect on the
183: stability of Poiseuille flow of a viscosity gradient placed in the
184: vicinity of the critical layer. This will provide us with a very sharp
185: understanding of the mechanism of the stabilization of the flow by
186: viscosity variations. In the following subsection we will examine the
187: case of continuous viscosity profiles.
188: 
189: \subsection{Mixed Layer}
190: \label{primix}
191: 
192: A report of the results of this subsection was provided in
193: \cite{01GLP}. We examine a channel flow of two fluids with different
194: viscosities $\mu_1$ and $\mu_2$, see Fig. \ref{f:scheme}.
195: 
196: The observation that we want to focus on is shown in Fig. \ref{f:rcr}:
197: the threshold Reynolds number for the loss of stability of the mode as
198: in Eq.  (\ref{eq:inst}) depends crucially on the position of the mixed
199: layer. When the latter hits the critical layer, the threshold Reynolds
200: number for the loss of stability reaches as much as 88000. In other
201: words, one can increase the threshold of instability {\em for a given
202: mode} 15 times, and by making the mixed layer thinner one can reach
203: even higher threshold Reynolds values.
204: %%%%%%%%%%%%%%%%%%%%%%%
205: \begin{figure}
206: \epsfxsize=8.3
207: cm
208: \epsfbox{Ramarcr.eps}
209: \caption{The dependence of the threshold Reynolds
210: number Re$_{\rm th}$ on the position of the viscosity stratified layer
211: for $m=0.9$. The dashed line pertains to the neat fluid. Note the huge
212: increase in $\Re_{\rm th}$ within a small range. This occurs when the
213: stratified layer overlaps the critical layer.}
214: \label{f:rcr}
215: \end{figure}
216: %%%%%%%%%%%%%%%%%%%%%%%
217: In \cite{01GLP} we analyzed the physical origin of this huge
218: sensitivity of the flow stability to the profile of the viscosity.
219: 
220: The stability of this flow is governed by the modified Orr-Sommerfeld
221: equation \cite{white}
222: \begin{eqnarray}
223: \nonumber
224: && \i\k_p
225: \left[\left(\phi_p''-\k_p^2\phi_p\right)(\bar U-c_p-\i
226: \gamma_p) - \bar
227: U''\phi_p \right] \nonumber\\&&= {1\over {\rm Re}}\bigg[\mu
228: \phi_p^{\rm (4)}
229: + 2
230: \mu' \phi_p''' +\left(\mu'' - 2 \k_p^2 \mu \right)\phi_p''
231: \nonumber\\&&-
232: 2
233: \k_p^2 \mu' \phi_p' +
234: \left(\k_p^2 \mu'' + \k_p^4 \mu\right)
235: \phi_p\bigg]\,,
236: \label{modOS} 
237: \end{eqnarray} 
238: in which $\bar U(y)$ is the basic laminar velocity, and $\mu$ is a
239: function of $y$. The boundary conditions are $\phi_p(\pm 1) = \phi_p'
240: (\pm 1) = 0$. All quantities have been non-dimensionalised using the
241: half-width $H$ of the channel and the centerline velocity $U_0$ as the
242: length and velocity scales respectively.  The Reynolds number is
243: defined as Re$\equiv \rho U_0 H /\mu_2$, where $\rho$ is the density
244: (equal for the two fluids). The primes stand for derivative with
245: respect to $y$.  At $y=0$, we use the even symmetry conditions
246: $\phi(0)=1$ and $\phi'(0)=0$, as the even mode is always more unstable
247: than the odd.
248: 
249: Since the flow is symmetric with respect to the channel centreline, we
250: restrict our attention to the upper half-channel. Fluid 2 occupies the
251: region $0 \le y \le p$. Fluid 1 lies between $p+q \le y \le 1$. The
252: region $p \le y \le p+q$ contains mixed fluid. The viscosity is
253: described by a steady function of $y$, scaled by the inner fluid
254: viscosity $\mu_2$:
255: \begin{eqnarray}\label{muin}
256: \hskip -0.4cm \mu(y) &=& 1
257: \,,\quad  \mbox{for} \quad 0 \le y \le p\,,\\
258: \label{vis5}
259: \hskip -0.4cm \mu(y) &=& 1 +
260: (m-1)\, \xi^3\left[10 - 15\, \xi + 6 \xi^2
261: \right], \ 0 \le \xi \le 1\,,
262: \\
263: \label{muout}
264: \hskip -0.4cm \mu(y) &=& m \,,\quad  \mbox{for} \quad p+q \le y \le
265: 1.
266: \end{eqnarray}
267: Here $\xi\equiv (y-p)/q$ is the mixed layer coordinate.  We have
268: assumed a 5th-order polynomial profile for the viscosity in the mixed
269: layer, whose coefficients maintain the viscosity and its first two
270: derivatives continuous across the mixed layer. The exact form of the
271: profile is unimportant. For a plot of the profile $m=0.9$, see
272: Fig.~\ref{f:thinprof}.
273: 
274: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
275: \begin{figure}
276: \epsfxsize=7.5cm  \epsfbox{RamaU.eps}
277: \caption{Profiles of the normalized
278: viscosity $\mu(y)$ and normalized velocity $\bar U(y)$ and the second
279: derivative $\bar U^{''}(y)$ for $m=0.9$ (solid lines) and $m=1.0$
280: (dashed lines). The mixed layer is between the vertical dashed lines.
281: }
282: \label{f:thinprof} 
283: \end{figure} 
284: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
285: 
286: The basic flow $\bar U(y)$ is obtained by requiring the velocity and
287: all relevant derivatives to be continuous at the edges of the mixed
288: layer:
289: \begin{eqnarray}\label{ui}
290: \hskip -0.5cm &&   \bar U(y) = 1-Gy^2/2\,, \quad
291: \mbox{for} \quad y \le p\,,\\
292: \hskip -0.5cm &&\bar U(y) = U(p) - G\int_p^ydy~ y/ \mu
293: \,,\quad  \mbox{for}
294: \quad p \le y \le p+ q,
295: \label{um}\\
296: \hskip -0.5cm && \bar U(y) =
297: G\left(1-y^2\right)/2m, \quad  \mbox{for}
298: \quad y \ge p+q\
299: .\label{uo}
300: \end{eqnarray}
301: Here $G$ is the streamwise pressure gradient.
302: 
303: It can be seen, comparing the mean profile $\bar U(y)$ to that of the
304: neat fluid (cf. Fig. \ref{f:thinprof}), that nothing dramatic happens
305: to this profile even when the mixed layer is chosen to overlap a
306: typical critical layer. Accordingly we need to look for the origin of
307: the large effect of Fig. 2 in the energetics of the disturbances. To
308: do so, recall that the streamwise and normal components of the
309: disturbance velocity $\hat u_p(x,y,t)$ and $\hat v_p(x,y,t)$ may be
310: expressed via streamfunction as usual: $\hat u_p(x,y,t) = \partial\hat
311: \phi_p/\partial y$, ${\rm and} \quad \hat v_p(x,y,t) = -\partial \hat
312: \phi_p/\partial x$.  These functions may be written in terms of
313: complex envelopes similar to Eq.~(\ref{eq:inst}):
314: \begin{eqnarray} 
315:   \label{eq:vel-env}
316:   \hat u_p(x,y,t) &\!\!\!=\!\!\!& \frac{1}{2}\big\{u_p(y)
317:   \exp\left[\i\k_p(x\!-\!c_p\, t)\right]\!+\!  \mbox{c.c.}\big\}
318:   \exp(\gamma_p t),\\ \nonumber \hat v_p(x,y,t) &\!\!=\!\!&
319:   \frac{1}{2}\big\{v_p(y) \exp\left[\i\k_p(x-c_p\, t)\right] +
320:   \mbox{c.c.}\big\} \exp(\gamma_p t) \ .
321: \end{eqnarray} 
322: The pressure
323: disturbance $\hat p_p$ is defined similarly.
324: 
325: Define now a disturbance of the density of the kinetic ene rgy of the
326: primary instability 
327: \be\label{kinen} \hat E_p(x,y,t) =
328: \frac{1}{2}\left[ \hat u_p(x,y,t)^2 + \hat v_p(x,y,t)^2\right] \ .
329: \ee We can express the mean (over $x$) density of the kinetic energy
330: as follows:
331: \begin{eqnarray}\label{averen}
332:   E_p(y,t)&\equiv& \left< \hat
333: E_p(x,y,t)\right>_x =\C.E_p (y)\exp\,(2
334: \gamma_p t)\,,
335: \\ \nonumber
336: \C.E_p
337: (y)&=&\frac{1}{4}\left( |u_p(y)|^2 + |v_p(y)|^2 \right)\ .
338: \end{eqnarray}
339: 
340: The physics of our phenomenon will be discussed in terms of the
341: balance equation for the averaged disturbance kinetic energy. Starting
342: from the linearized Navier-Stokes equations for $\hat u_p$ and $\hat
343: v_p$, dotting it with the disturbance velocity vector, averaging over
344: one cycle in $x$ and using Eqs. (\ref{eq:vel-env})-(\ref{averen})
345: leads to \be 2 \gamma_p \,\C.E_p (y) = \nabla \cdot J_p(y) + W_{p+}(y)
346: - W_{p-}(y) \,,
347: \label{enbal} 
348: \ee where the energy flux $J_p(y)$ in the $y$ direction, rates of
349: energy production (energy taken up by the primary instability from the
350: mean flow) $W_{p+}(y)$ and energy dissipation (by the viscosity)
351: $W_{p-}(y)$ are given by
352: \begin{eqnarray}\label{current}
353:   J_p(y) &\equiv&
354: {\left[u_p(y) p_p^*(y)+\mbox{c.c.}\right]
355: \over 4 \rho} + {1 \over \R} \mu
356: (y) \nabla \C.E_p(y)\,,\\
357: W_{p+}(y) &\equiv&  - {1 \over 4} \bar
358: U'(y)\left[u_p(y)
359:  v_p^*(y) +\mbox{c.c.}  \right]\,,
360: \label{prod}
361: \\
362: W_{p-}(y) &\equiv& {\mu(y) \over \R} \left\{2 \k_p^2 \C.E_p(y)
363: + {1 \over
364: 2}\left[|u_p'(y)|^2+|v_p'(y)|^2\right]\right\}\
365: .
366: \nonumber
367: \end{eqnarray}
368: The superscript $*$ denotes complex conjugate. To plot these functions
369: we need to solve Eq. (\ref{modOS}) as an eigenvalue problem, to obtain
370: $c_p$, $\gamma_p$, and $\phi_p(y)$ at given $\R$ and $\k_p$.  The
371: value of $c_p$ determines the position of the critical layer. It is
372: convenient to compute and compare the space averaged production and
373: dissipation terms $\Gamma_{p+}$ and $\Gamma_{p-}$ defined by:
374: %%%%%%%%%%%%%%%%%%
375: \be\label{totpd} \Gamma_{p\pm} \equiv\int_0^1 W_{p\pm}(y) \d y \ \Big/
376: \int_0^1 \C.E_p(y) \d y \ .  \ee The local production of energy can be
377: positive or negative, indicative of energy transfer from the mean flow
378: to the primary disturbance and vice-versa respectively. The production
379: in one region (where $ W_{p+}(y)>0$) can be partly canceled out by a
380: ``counter-production'' in other region (where $ W_{p+}(y)<0$).
381: 
382: The use of these measures can be exemplified with the neat fluid
383: ($m=1.0$ here). The laminar flow displays its first linear instability
384: at a threshold Reynolds number of $\R_{\rm th}=5772$, which means that
385: the total production $\Gamma_{p+}$ across the layer becomes equal to
386: the total dissipation $\Gamma_{p-}$ at this value of $\R$. Examining
387: Fig.  \ref{f:far} we can see that the disturbance kinetic energy is
388: produced predominantly within the critical layer, where the basic flow
389: velocity is close to the phase speed of the disturbance, while most of
390: the dissipation is in the wall layer.
391: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
392: \begin{figure} 
393: \epsfxsize=7.5cm
394: \epsfbox{Ramas0.eps}
395: \epsfxsize=7.5cm \epsfbox{Ramafar.eps}
396: \caption{Energy
397: balance: production $W_{p+}(y)$, solid line; dissipation
398: $W_{p-}(y)$,
399: dot-dashed line, $\R=5772$. Top: $m=1$, $\Gamma_{p+}=
400: \Gamma_{p-}=0.0148$.
401: Bottom: $m=0.9$, $p=0.3$, $\Gamma_{p+}=0.0158$,
402: $\Gamma_{p-}=0.0148$. In
403: this and the two subsequent figures the solid
404: vertical lines show the
405: location $y_c$ of the critical lines, whereas the
406: region
407: between the dotted
408: lines is the mixed layer.}
409: \label{f:far} 
410: \end{figure}
411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
412: The balance is not changed significantly when the viscosity ratio is
413: changed to $0.9$ so long as the mixed layer is not close to the
414: critical layer.  There is a small region of production and one of
415: counter-production within the mixed layer, whose effects cancel out,
416: leaving the system close to marginal stability.
417: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
418: \begin{figure}
419: \epsfxsize=7.5cm \epsfbox{Ramanear.eps}
420: \caption{Energy balance: production
421: $W_{p+}(y)$, solid line; dissipation
422: $W_{p-}(y)$, dot-dashed line.
423: $\R=5772$, $m=0.9$, $p=0.85$, $\Gamma_{p+}=
424: -0.0114$, $\Gamma_{p-}=0.0122$.
425: }
426:  \label{f:near} 
427: \end{figure} 
428: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
429: 
430: We now turn our attention to Fig. \ref{f:near}, in which our main
431: point is demonstrated.  The Reynolds number is the same as before, but
432: the mixed layer has been moved close to the critical layer. It is
433: immediately obvious that the earlier balance is destroyed. The
434: counter-production peak in the mixed layer is much larger than before,
435: making the flow more stable.  The wavenumber used is that at which the
436: flow is least stable for the given Reynolds number at this $p$. For
437: $m=0.9$, the threshold Reynolds number is $46400$.  Fig.  \ref{f:marg}
438: shows the energy balances at marginal stability - the picture is
439: qualitatively the same here as at $\R\approx 5772$ for the neat fluid.
440: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
441: \begin{figure}
442: \epsfxsize=7.5cm
443: \epsfbox{Ramamarg.eps}
444: \caption{Energy balance: production $W_{p+}(y)$,
445: solid line; dissipation
446: $W_{p-}(y)$, dot-dashed line. $\R=46400$, $m=0.9$,
447: $p=0.85$, $\Gamma_{p+}=
448: \Gamma_{p-}=0.0053$. }
449:  \label{f:marg} 
450: \end{figure}
451: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
452: \subsection{The mechanism of
453: stabilization}
454:  
455: The main factor determining the instability is the energy intake from
456: the mean flow, which is driven by the phase change caused by the
457: viscosity stratification. The dissipation on the other hand depends
458: only on Reynolds number and does not respond disproportionately to
459: changes in viscosity. In neat fluids, the term containing $\bar
460: U''(y)$ in (\ref{modOS}) is always of higher order within the critical
461: layer. However, with the introduction of a viscosity gradient within
462: the critical layer, the gradients of the basic velocity profile will
463: scale according to the mixed layer coordinate $\xi$. We show in the
464: analysis that follows that for $q \le O(\R^{-1/3})$, the term
465: containing $\bar U''$ is now among the most dominant. Since most of
466: the production of disturbance kinetic energy takes place within the
467: critical layer, we return to equation (\ref{modOS}) and isolate all
468: lowest-order effects within the critical layer. The relevant normal
469: coordinate in the critical layer is \be \eta \equiv {y-y_c \over \ep}
470: \label{eta}
471: \ee
472: where $y_c$ is the critical point defined by $U(y_c)=c$, and $\ep$ is the
473: critical layer thickness, which is a small parameter at large Reynolds
474: number. The basic channel flow velocity may be expanded in the vicinity of
475: the critical point as
476: \be
477: U(y) = c + (y-y_c) U'(y_c) + {(y-y_c)^2 \over 2!} U''(y_c) + \cdots.
478: \label{ucrit}
479: \ee
480: We use (\ref{ucrit}), and redefine $\phi_p(y)\equiv\Phi(\eta)$ and
481: $\mu(y)\equiv \nu(\xi)$, to rewrite
482: (\ref{modOS}) within the critical layer. We obtain
483: \be
484: \ep \sim \R^{-1/3} \equiv \left(\k_p\R\right)^{-1/3},
485: \label{epsilon}
486: \ee and the lowest order equation in the critical layer: \be \i \eta
487: {\d U \over \d y}\bigg|_c\Phi'' - {\i G p \over \nu^2} \chi\nu'\Phi =
488: \nu \Phi^{\rm (4)} + 2 \chi\nu' \Phi''' + \chi^2\nu''\Phi''\,,
489: \label{critlow}
490: \ee where $\chi \equiv \ep /q$ is $O(1)$ for the mixed layer. In the
491: absence of a viscosity gradient in the critical layer (i.e. $\nu=1$),
492: equation (\ref{critlow}) would reduce to \be \i \eta {\d U \over \d
493: y}\bigg|_c\Phi'' = \Phi^{\rm (4)}\,,
494: \label{lin}
495: \ee which is the traditional lowest-order critical layer equation for
496: a parallel shear flow \cite{45Lin}. The mechanism for the
497: stabilization now begins to be apparent: there are several new terms
498: which can upset the traditional balance between inertial and viscous
499: forces. In order to narrow down the search further, we resort to
500: numerical experimentation, because although all terms in
501: (\ref{critlow}) are estimated to be of $O(1)$, their numerical
502: contributions are different.  It transpires that the second term on
503: the left hand side of (\ref{critlow}) is particularly responsible: it
504: is straightforward to verify that it originates from the term
505: containing $\bar U''(y)$ in the modified Orr-Sommerfeld equation. As
506: testimony, note the dramatic effect on $\bar U''$ in Fig. 3. Any
507: reasonable viscosity gradient of the right sign will pick up this
508: term, leading to vastly enhanced stability.
509: 
510: Indeed, in the light of this discussion we can expect that the large
511: effect of retardation of the instability would even increase if we
512: make the mixed layer thinner. This is indeed so. Nevertheless, one
513: cannot conclude that instability can be retarded at will, since other
514: disturbances, differing from the primary mode, become unstable first,
515: albeit at a much higher Reynolds number than the primary mode; when we
516: stabilize a given mode substantially, we should watch out for other
517: pre-existing/newly destabilized modes which may now be the least
518: stable.
519: 
520: Finally, we connect our findings to the phenomenon of drag reduction
521: in turbulent flows. Since the total dissipation can be computed just
522: from the knowledge of the velocity profile at the walls, any amount of
523: drag reduction must be reflected by a corresponding reduction of the
524: gradient at the walls.  Concurrently, the energy intake by the
525: fluctuations from the mean flow should reduce as well. Indeed, the
526: latter effect was measured in both experiments \cite{97THKN} and
527: simulations \cite{01DSBH,00Ang}. The question is which is the chicken
528: and which is the egg. In our calculation we identified that the
529: reduction in production comes first. From Figs. 4 and 5 which are at
530: the same value of $\R$ we see that the dissipation does not change at
531: all when the mixed layer moves, but the production is strongly
532: affected. Of course, at steady state the velocity gradient at the wall
533: must adjust as shown in Fig. 6.
534: 
535: \subsection{Continuous Viscosity
536: Profile}
537: \label{contprim}
538: 
539: One could think that the strong stabilization discussed in the
540: previous subsection is only due to the precise positioning of the
541: mixed layer at the critical layer. If so, the result would have very
542: little generic consequence.  In this subsection we show that any
543: reasonable viscosity profile achieves the same effects. To this aim we
544: consider the effective viscosity profile reported in \cite{97SBH} (in
545: their Fig. 5) which is obtained from simulations of a turbulent
546: channel flow with polymer additive. It may be prescribed as
547: \begin{eqnarray}\label{cmuin}
548: \mu(y) &=& 1
549: \,,\quad  \mbox{for} \quad 0 \le y \le p\,,\\
550: \label{vis3}
551: \mu(y) &=& 1 +
552: (m-1)\, \left({y-p \over q}\right)^3,
553: \end{eqnarray}
554: with $q \sim 0.4$, and
555: $m \sim 0.7$, as shown in Fig.
556: \ref{f:contprof}.
557: \begin{figure}
558: \epsfxsize=7.5cm
559: \epsfbox{Ramacontprof.eps}
560: \caption{Prescribed continuous viscosity profile
561: (in accordance with that
562: obtained in direct numerical simulations of
563: polymeric flow). The
564: corresponding laminar velocity profile $\bar U(y)$ and
565: its second derivative
566: are also shown.}
567: \label{f:contprof}
568: \end{figure}
569: The energy balance for the least stable primary mode at $\R=5772$ for
570: this case (Fig. \ref{f:polyprim}) shows a large counter-production of
571: disturbance kinetic energy, which is in fact more pronounced than what
572: we obtained with a mixed layer (Fig. \ref{f:near}). Thus the strong
573: stabilization effect does not require careful placing of the viscosity
574: variation at a particular layer.  It is sufficient that there exist a
575: viscosity variation in the region of the critical layer (indicated as
576: the vertical line in Fig. \ref{f:polyprim}) to achieve the
577: stabilization.
578: 
579: It comes as no surprise that this continuous viscosity profile behaves
580: very similarly to the thin mixed-layer. If we return to equation
581: (\ref{critlow}), we will see that all we have now done is to increase
582: both $\nu'$ (which is proportional to $m-1$) and $q$ threefold (the
583: effective $q$ here is closer to 0.3 than 0.4, as we can see from
584: Fig. \ref{f:contprof}), so the ratio remains the same.
585: 
586: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
587: \begin{figure} 
588: \epsfxsize=7.5cm \epsfbox{Ramapolyprim.eps}
589: \caption{Energy balance: production $W_{p+}(y)$, solid line; dissipation
590: $W_{p-}(y)$, dot-dashed line. $\R=5772$, $m=0.7$, $p=0.6$, $q=0.4$,
591: $\Gamma_{p+}= -0.0345$, $\Gamma_{p-}=0.0138$. }
592:  \label{f:polyprim}
593: \end{figure} 
594: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
595: \section{Secondary Instabilities}
596: \label{secondary}  
597: 
598: A laminar flow through a channel is linearly unstable at $\R=5772$. In
599: all except the cleanest experiments, however, the flow becomes
600: turbulent at much lower Reynolds numbers, as low as $1000$
601: \cite{80OK,93TTRD}. This is because the linear stability analysis is
602: carried out on a steady laminar velocity profile, whereas a real flow,
603: except under carefully designed clean conditions, consists in addition
604: of small but finite disturbances (most of whom will decay at long
605: times). The stability behaviour of the real flow is quite different
606: from that of the steady profile: the actual flow is unstable to new
607: modes, often referred to as secondary modes. The secondary modes are
608: often three dimensional, and their signature is prominent in
609: fully-developed turbulence. As described below, the secondary
610: instabilities are studied by a Floquet analysis of the periodic
611: primary flow we obtained earlier.
612: 
613: As is usual in the analysis of secondary instabilities
614: \cite{88Herb,83Herb}, we begin by splitting the flow into a periodic
615: component (consisting of the mean laminar profile in addition to the
616: primary wave) and a secondary disturbance, e.g., \be \B.U_{\rm
617: total}(x,y,z,t) = \B.U(x,y,t) + \B.u_s(x,y,z,t), \ee where
618: \begin{eqnarray}
619: &&\!\!\B.U(x,y,t) = \bar U(y) \hat \B.x \label{basic}\\&&\!\!+ A_p(t)
620: \left\{\left[u_p(y) \hat \B.x  + v_p(y) \hat \B.y
621: \right] \exp \left[\i k_p (x - c_p t)\right] + {\rm c.c.} \right\}.
622: \nonumber
623: \end{eqnarray}
624: Here $\hat \B.x$ and $\hat \B.y$ are units vectors in the $x$
625: (steamwise) and $y$ (wall normal) directions.  The amplitude $A_p$ of
626: the primary disturbance changes very slowly with time, and $\d A_p /\d
627: t$ may be neglected during one time period. The spatial and temporal
628: dependence of the secondary disturbance is written in the form
629: \begin{eqnarray}
630: &&\B.u_s(y,\B.r_\perp,t)\equiv {\cal R}\!e\Big\{\B.u_{s+}(y) \exp\left[\i
631: \left(\B.k_+ \cdot \B.r_\perp - \omega_+ t\right)\right] \nonumber\\&&+
632: \B.u_{s-}(y) \exp\left[\i \left(\B.k_-\cdot \B.r_\perp - \omega_- t\right)
633: \right]\Big\},
634: \label{form}
635: \end{eqnarray}
636: where $\B.r_\perp\equiv x\hat \B.x +z\hat \B.z $, and $\B.k_\pm= k_\pm
637: \hat\B.x \pm k_z \hat\B.z$.  We substitute the above ansatz into the
638: Navier-Stokes and continuity equations, and retain linear terms in the
639: secondary. On averaging over $x$, $z$ and $t$, only the resonant modes
640: survive, which are related by \be \B.k_+ + \B.k_- = k_p \hat \B.x \ ,
641: \quad \text{therefore}~~\B.k_\pm=\pm \B.q+\frac{k_p}{2} \hat \B.x \ ,
642: \ee for any vector $\B.q$, and \be \omega_+ = \omega + \i\gamma_s
643: \qquad {\rm and} \qquad \omega_- = (\omega_p - \omega) + \i\gamma_s.
644: \label{om}
645: \ee Eliminating the disturbance pressure and streamwise component of
646: the velocity, we get the equations for the secondary disturbances
647: $v_s$ and $w_s$.  Using the operator $D$ for differentiation with
648: respect to the normal coordinate $y$, and the notation $f_\pm \equiv -
649: \i w_{s\pm}/k_z$, the equations read \cite{02SG}
650: \begin{eqnarray}
651: &&\A \nonumber\\&&
652: \times\left[\s f_+ - Dv_+\right] - \i k_+ U' v_+ \label{first}\\&&-
653: {A_p k_+ \over 2 k_-}\Bigg\{\left[\i k_+ u_p D + v_p D^2 +\i k_- Du_p
654: \right]v_-^*
655: \nonumber\\&&+ \left[\left(\beta^2-k_-k_+\right)v_pD + \i k_+
656: \left(k_-^2+\beta^2\right)u_p\right]f_-^*\Bigg\} = 0,
657: \nonumber
658: \end{eqnarray}
659: and
660: \begin{eqnarray}
661: &&\A 
662: \left(Df_+ - v_+\right) \nonumber\\&&+ \bigg[-\i k_+ (DU) + (D^2\nu) D +
663: (D\nu)(D^2 - k_+^2 - \beta^2)\bigg] f_+ \nonumber\\&&+ {A_p (k_p + k_-)
664: \over 2} \bigg[ \i u_p \left(v_-^* + Df_-^*\right)
665: -{v_p \over k_-}Dv_-^* \bigg] \nonumber\\&&+
666: {A_p \over 2}\left[v_p\left({k_p\beta^2 \over k_-} + D^2\right) -
667: \i k_-(Du_p) \right]f_-^* = 0,
668: \label{second}
669: \end{eqnarray}
670: The boundary conditions are
671: \be
672: \B.u_s=0 \qquad {\rm at} \qquad y = \pm 1.
673: \label{bcs}
674: \ee Equations (\ref{first}) and (\ref{second}), along with two
675: corresponding equations in $v_-^*$ and $f_-^*$, describe an eigenvalue
676: problem for the secondary instability. The four equations are solved
677: by a Chebychev collocation spectral method, details of the solution
678: procedure are available in \cite{02SG}.
679: 
680: The most unstable secondary mode in our case is found to be the
681: subharmonic, for which $\B.q=\beta \hat \B.z$.  The production and
682: dissipation are computed as before.
683: 
684: We survey in turn the thin mixed-layer profile, and the continuous
685: viscosity profile to see what viscosity variation does to the
686: secondary instability.
687: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
688: \subsection{Mixed Layer}
689: \label{thin-sec}
690: 
691: The velocity and viscosity profiles here are as given in Fig.
692: \ref{f:thinprof}, and the primary instability is that presented in
693: Sect.  \ref{primix}. Since the subharmonic ($k_+=k_-= k_p/2$) is the
694: least stable mode, we present this case alone. In Fig. \ref{f:thinmax}
695: a typical dependence of the growth rate of the secondary mode on the
696: spanwise wavenumber is shown. We can see that the viscosity variation
697: damps the secondary mode significantly, but it is still
698: unstable. However, there is a crucial difference in the {\em primary}
699: instabilities of the two: the primary is unstable for a constant
700: viscosity flow, but very stable in the mixed layer case. Therefore at
701: long times, the secondary mode, which feeds on the primary for its
702: existence, dies down in the latter case.  To compute the time
703: dependence of the amplitude of the secondary mode we computed the
704: growth rate $\gamma_s$ by neglecting the time dependence of the
705: amplitude of the primary words.  As a result we obtain the growth rate
706: $\gamma_s[A_p(t)]$, in which $A_p(t)$ can be an exponentially growing
707: or a decaying function of time. Having this growth rate we can present
708: the time dependence of the amplitude of the secondary mode, see Fig.
709: \ref{f:thinamp}. Without the viscosity contrast, the amplitude of the
710: secondary mode increases (essentially exponentially).  With the
711: viscosity contrast the amplitude decays in time.
712: 
713: We now observe the balances of energy initially and at a later time in
714: Figs.  \ref{f:thinprod0} and \ref{f:thinprod40} respectively. The
715: initial balance of energy is not very different from the constant
716: viscosity case.  At the later time, however, the production of
717: secondary kinetic energy is significantly lower. The location $y_c$ of
718: the critical point is seen from the figures to be close to the layer
719: of stratified viscosity. If the two were well-separated, the
720: stratification would do nothing to the secondary mode.
721: 
722: \begin{figure}
723: \epsfxsize=7.5cm \epsfbox{Ramathinmax.eps}
724: \caption{Dependence of growth rate on spanwise wavenumber. Solid line:
725: varying viscosity ($p=0.8, q=0.1, m=0.9$); dashed line: constant
726: viscosity ($m=1$). $k_p=1, A_p=0.005, \R=6000$. }
727: \label{f:thinmax}
728: \end{figure}
729: 
730: \begin{figure}
731: \epsfxsize=7.5cm \epsfbox{Ramathinamp.eps}
732: \caption{Amplitude of the secondary mode in logarithmic 
733: scale as a function of time. Dashed line: constant viscosity,
734: $m=1$. Here $\gamma_p=0.0003$, and the primary mode is unstable. Solid
735: line: varying viscosity; here $\gamma_p=-0.0206$, the primary mode is
736: stable. All conditions like in Fig. \ref{f:thinmax}, in particular
737: $A_p(t=0)=0.005$.}
738: \label{f:thinamp}
739: \end{figure}
740: 
741: \begin{figure}
742: \epsfxsize=7.5cm \epsfbox{Ramathinpr0.eps}
743: \caption{Production $W_{s+}$ and dissipation $W_{s-}$ of the 
744: kinetic energy
745: of the secondary disturbance at time=0. Solid line: $W_{s+}, m=0.9$;
746: dot-dashed line: $W_{s-}, m=0.9$; long dashes: $W_{s+}, m=1$; dotted
747: line: $W_{s-}, m=1$.  The vertical lines show $y_c$ (the critical
748: point location) for $m=0.9$ (solid) and $m=1$ (dotted).}
749: \label{f:thinprod0}
750: \end{figure}
751: 
752: \begin{figure}
753: \epsfxsize=7.5cm \epsfbox{Ramathinpr40.eps}
754: \caption{Production $W_{s+}$ and dissipation $W_{s-}$ of 
755: the kinetic energy of the secondary disturbance at time=40. Solid
756: line: $W_{s+}, m=0.9$; dot-dashed line: $W_{s-}, m=0.9, A_p=0.00215$;
757: long dashes: $W_{s+}, m=1$; dotted line: $W_{s-}, m=1, A_p=0.00506$.}
758: \label{f:thinprod40}
759: \end{figure}
760: 
761: A lowest-order analysis of the secondary stability equations is not as
762: straightforward as for the primary mode, since the secondary is highly
763: dependent on the amplitude of the primary \cite{02SG}. We may however
764: make the following observations from a critical layer analysis of
765: equations (\ref{first}) and (\ref{second}) and their
766: counterparts. When $A_p\gg \ep$, (cf. Eq.(\ref{epsilon})) only the
767: nonlinear terms appear at the lowest order, and the secondary mode is
768: completely driven by the primary.  When $A_p \sim O(\ep)$, both the
769: basic terms and the nonlinear terms contribute at the lowest order. It
770: may be numerically determined, however, that the secondary is slaved
771: to the primary here as well. When $A_p=o(\ep)$, the lowest-order
772: theory for the secondary is (not surprisingly) exactly that given by
773: (\ref{critlow}) for the primary.
774: 
775: A direct estimate of the effect of the viscosity stratification on the
776: secondary mode is obtained from the threshold amplitude $A_{\rm th}$ of
777: the primary for the instability. At a Reynolds number of 6000 and
778: primary wavenumber of $k_p=1$, for a neat fluid, all secondary modes
779: are damped if $A_{\rm th}<0.002$, while for the continuous viscosity
780: profile, all secondary modes continue to be damped even for larger
781: primary disturbances, up to $A_{\rm th}=0.005$. When the Reynolds number
782: is reduced to 2000, the threshold amplitudes are 0.012 and 0.016 for
783: the neat and viscosity-stratified fluids respectively.
784: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
785: \subsection{Continuous Viscosity Profile}
786: \label{cont-sec}
787: 
788: The velocity and viscosity profiles here are as given in Fig.
789: \ref{f:contprof}, and the primary instability is that presented in Sect.
790: \ref{contprim}. The counterparts for the continuous viscosity
791: profile of Figs. \ref{f:thinmax} to \ref{f:thinprod40} are presented
792: in Figs. \ref{f:max} to \ref{f:prod40} respectively. It is clear that
793: nothing has changed qualitatively.
794: 
795: \begin{figure}
796: \epsfxsize=7.5cm \epsfbox{Ramamax.eps}
797: \caption{Dependence of growth rate on spanwise wavenumber. Solid line:
798: varying viscosity [according to equation (\ref{vis3})]; dashed line:
799: constant viscosity. Wavenumbers and $\R$ as in Fig. \ref{f:thinmax}.}
800: \label{f:max}
801: \end{figure}
802: 
803: \begin{figure}
804: \epsfxsize=7.5cm \epsfbox{Ramaamp.eps}
805: \caption{Variation of the amplitude of the secondary instability mode with
806: time. Solid line: varying viscosity, $\gamma_p=-0.0244$; dashed line: 
807: constant viscosity, $\gamma_p=0.0003$.
808: Wavenumbers and $\R$ as in Fig. \ref{f:thinamp}.}
809: \label{f:amp}
810: \end{figure}
811: 
812: \begin{figure}
813: \epsfxsize=7.5cm \epsfbox{Ramaprod0.eps}
814: \caption{Production and dissipation at time=0. Solid line: varying
815: viscosity;
816: dashed line: constant viscosity, $A_p=0.005$ for both.}
817: \label{f:prod0}
818: \end{figure}
819: 
820: \begin{figure}
821: \epsfxsize=7.5cm \epsfbox{Ramaprod40.eps}
822: \caption{Production and dissipation at time=40. Solid line: varying
823: viscosity, $A_p=0.00184$;
824: dashed line: constant viscosity, $A_p=0.00506$.}
825: \label{f:prod40}
826: \end{figure}
827: 
828: Fig. \ref{f:ampdep} shows the dependence of the growth rate of the 
829: secondary mode on the amplitude of the primary disturbance. It is
830: clear that the instability is reduced by the stratification of
831: viscosity, but there is no dramatic effect in the secondary alone. We
832: may conclude that the large effect comes from the complete reliance 
833: of the secondary on the primary.
834: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
835: \begin{figure}
836: \epsfxsize=7.5cm \epsfbox{Ramaampdep.eps}
837: \caption{Dependence of the growth rate of the secondary mode on the
838: amplitude of the primary disturbance. Solid line: stratified 
839: viscosity; dashed line: constant viscosity. $k_p=1, k_+=0.5, \beta=1,
840: \R=6000$.}
841: \label{f:ampdep}
842: \end{figure}
843: %%%%%%%%%%%%%%%%%%%%%%%%%
844: \section{Concluding Remarks}
845: \label{conclude}
846: 
847: We addressed the primary and secondary instability of simple channel
848: flows, and examined the effects of small viscosity variations. We find
849: dramatic effects of stabilization when the viscosity variations exist
850: in the vicinity of the critical layers, in which the speed of
851: propagation of the modes coincided with the mean velocity of the basic
852: flow. With about 10\% viscosity changes we can have very large
853: increases in the threshold Reynolds numbers for instability. In all
854: cases we find that the main mechanism for the large effects is the
855: reduction of the intake of energy from the mean flow to the putative
856: unstabale modes, which therefore become stable. For the same Reynolds
857: numbers in Newtonian fluids there is no such mechanism for
858: stabilization and these flows will become turbulent. We would like to
859: propose that similar effects should be examined in the case of
860: turbulent drag reduction by polymer additives.
861: 
862: 
863: We recognize that in a turbulent flow there are many more modes that
864: interact, but we propose that a similar mechanism operates for each
865: mode at its critical layer, where both elastic and viscous effects
866: determine the mean flow. The advantage of the present calculation is
867: that we can consider all the putative unstable modes, and conclude
868: that with a viscosity gradient similar to that seen in polymeric
869: turbulent flows the linear threshold $\R_{\rm th}$ goes up five times
870: (to 31000). We note in passing that this effect had not been put to an
871: experimental test, and it would be exciting to have a confirmation of
872: our predictions by future experiments. For actual turbulent flows we
873: will need first to identify what are the main modes that interact
874: between themselves and with the mean flow. A significant numerical
875: effort is required, but appears worthwhile due to the importance of
876: the phenomenon of drag reduction, and its relative lack of
877: understanding.
878: 
879: We have demonstrated that the exact form of the viscosity profile is
880: immaterial; a continuous profile of viscosity in the critical region
881: behaves exaclty like a thin mixed layer. We have shown that the secondary
882: three dimensional modes of instability are ``slaved'' to the primary
883: linear mode of instability: the mechanism which stabilizes the primary
884: mode indirectly ensures that the secondary is damped out quickly.
885: 
886: Finally we note that a linear disturbance can rear its head either in
887: the form of the fastest growing (or slowest decaying) mode as
888: considered here; or in non-modal form with a transient growth followed
889: by long-term decay \cite{02Chap}. The former situation will correspond
890: to relatively high Reynolds numbers, or cleaner set-ups. We expect
891: similar conclusions in the latter situation as well.
892: %%%%%%%%%%%%%%%%%
893: \acknowledgments This work was supported by the German-Israeli
894: Foundation, the European Commission under a TMR grant, the Israeli
895: Science Foundation and the Naftali and Anna Backenroth-Bronicki Fund
896: for Research in Chaos and Complexity.  RG thanks the Defence R\&D
897: Organisation, India, for financial support.
898: 
899: 
900: \begin{references} 
901: 
902: \bibitem{95NH}
903: R.H. Nadolink and W.W. Haigh,  ASME Appl. Mech. Rev. {\bf 48}, 351 (1995).
904: 
905: \bibitem{69Lum}
906: J. L. Lumley, Ann. Rev. Fluid Mech. {\bf 1}, 367 (1969).
907: 
908: \bibitem{00SW}
909: K.R. Sreenivasan and C. M. White, J. Fluid Mech., {\bf 409}, 149 (2000).
910: 
911: \bibitem{86TG}
912: M. Tabor and P.G. de Gennes, Europhys. Lett., {\bf 2}, 519 (1986).
913: 
914: \bibitem{90Gennes}
915: P.G. de Gennes, {\em Introduction to Polymer Dynamics}, (Cambridge, 1990).
916: 
917: \bibitem{01GLP}
918: R. Govindarajan, V.S. L'vov and I. Procaccia, Phys. Rev. Lett.
919: {\bf 87}, 174501 (2001)
920: 
921: \bibitem{97SBH}
922: R. Sureshkumar, A. N.  Beris and R.A. Handler, Phys. Fluids {\bf 9}, 743
923: (1997).
924: 
925: \bibitem{white}
926: F.M White, {\em Viscous Fluid Flow} (McGraw Hill, 1991).
927: 
928: \bibitem{Rama} 
929: B.T. Ranganathan and R. Govindarajan, Phys. Fluids {\bf 13}, 1 (2001).
930: 
931: \bibitem{97THKN}
932: J.M.J den Toonder, M.A. Hulsen,G.D.C Kuiken and F.T.M Niewstadt,
933: J. Fluid Mech. {\bf 337}, 193 (1997).
934: 
935: \bibitem{01DSBH}
936: C.D. Dimitropoulos, R. Sureshkumar, A.N. Beris and R.A. Handler,
937: Phys. Fluids, {\bf 13}, 1016 (2001).
938: 
939: \bibitem{00Ang}
940: E. De Angelis, ``The Influence of Polymer Additives on the Structure
941: of Wall Turbulence", Thesis, Universita' di Roma, La Sapienza, 2000.
942: 
943: \bibitem{45Lin}
944: C.C. Lin, Quarterly of Appl. Math., {\bf 3}, 117-142 (1945).
945: 
946: \bibitem{80OK}
947: S.A. Orszag and L.C. Kells, J. Fluid Mech., {\bf 96} 159 (1980).
948: 
949: \bibitem{93TTRD}
950: L.N. Trefethen, A.E.Trefethen, S.C. Reddy and T.A. Driscoll, Science, {\bf
951: 261} 578 (1993).
952: 
953: \bibitem{88Herb}
954: T. Herbert, Ann. Rev. Fluid Mech., {\bf 20}, 487 (1988).
955: 
956: \bibitem{83Herb}
957: T. Herbert, Phys. Fluids, {\bf 26} (4), 871 (1983).
958: 
959: \bibitem{02SG}
960: A. Sameen and R. Govindarajan, preprint, 2002.
961: 
962: \bibitem{02Chap}
963: See for example: S.J. Chapman,  J. Fluid Mech., {\bf 451}, 35 (2002).
964: 
965: \end{references}  
966: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
967: 
968: \end{document}
969: 
970: \bibitem{33Squi}
971: H.B. Squire, Proc Roy. Soc. Lond. A, {\bf 142} 621-628 (1933).
972: