1: \documentclass{jfm}
2: %\documentclass[referee]{jfm}
3: \usepackage{epsf}
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: % begin JFM caption
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7:
8: % See if the author has AMS Euler fonts installed: If they have, attempt
9: % to use the 'upmath' package to provide upright math.
10:
11: \ifCUPmtlplainloaded \else
12: \checkfont{eurm10}
13: \iffontfound
14: \IfFileExists{upmath.sty}
15: {\typeout{Found AMS Euler Roman fonts on the system,
16: using the 'upmath' package.}%
17: \usepackage{upmath}}
18: {\typeout{Found AMS Euler Roman fonts on the system, but you
19: dont seem to have the}%
20: \typeout{'upmath' package installed. JFM.cls can take advantage
21: of these fonts, if you use 'upmath' package.}%
22: \providecommand\upi{\pi}%
23: }
24: \else
25: \providecommand\upi{\pi}%
26: \fi
27: \fi
28:
29: % See if the author has AMS symbol fonts installed: If they have, attempt
30: % to use the 'amssymb' package to provide the AMS symbol characters.
31:
32: \ifCUPmtlplainloaded \else
33: \checkfont{msam10}
34: \iffontfound
35: \IfFileExists{amssymb.sty}
36: {\typeout{Found AMS Symbol fonts on the system, using the
37: 'amssymb' package.}%
38: \usepackage{amssymb}%
39: \let\le=\leqslant \let\leq=\leqslant
40: \let\ge=\geqslant \let\geq=\geqslant
41: }{}
42: \fi
43: \fi
44:
45: % See if the author has the AMS 'amsbsy' package installed: If they have,
46: % use it to provide better bold math support (with \boldsymbol).
47:
48: \ifCUPmtlplainloaded \else
49: \IfFileExists{amsbsy.sty}
50: {\typeout{Found the 'amsbsy' package on the system, using it.}%
51: \usepackage{amsbsy}}
52: {\providecommand\boldsymbol[1]{\mbox{\boldmath $##1$}}}
53: \fi
54:
55:
56: %%% Example macros (some are not used in this sample file) %%%
57:
58: % For units of measure
59: \newcommand\dynpercm{\nobreak\mbox{$\;$dynes\,cm$^{-1}$}}
60: \newcommand\cmpermin{\nobreak\mbox{$\;$cm\,min$^{-1}$}}
61:
62: % Various bold symbols
63: \providecommand\bnabla{\boldsymbol{\nabla}}
64: \providecommand\bcdot{\boldsymbol{\cdot}}
65: \newcommand\biS{\boldsymbol{S}}
66: \newcommand\etb{\boldsymbol{\eta}}
67:
68: % For multi-letter symbols
69: \newcommand\Real{\mbox{Re}} % cf plain TeX's \Re and Reynolds number
70: \newcommand\Imag{\mbox{Im}} % cf plain TeX's \Im
71: \newcommand\Rey{\mbox{\textit{Re}}} % Reynolds number
72: \newcommand\Pran{\mbox{\textit{Pr}}} % Prandtl number, cf TeX's \Pr product
73: \newcommand\Pen{\mbox{\textit{Pe}}} % Peclet number
74: \newcommand\Ai{\mbox{Ai}} % Airy function
75: \newcommand\Bi{\mbox{Bi}} % Airy function
76: \newcommand\Ei{\mbox{Ei}} % Exponential integral
77: \newcommand\Ci{\mbox{Ci}} % Cosine integral
78: \newcommand\Si{\mbox{Si}} % Sine integral
79: \newcommand\ue{\mathrm{e}} % e
80: \newcommand\uexp{\mathrm{exp}} % "exp"
81: \newcommand\ui{\mathrm{i}} % I
82: \newcommand\ucc{\mathrm{c.c.}} % complex conjugate
83: \newcommand\uhot{\mathrm{h.o.t.}} % H.O.T.
84: \newcommand\ud{\mathrm{d}} % differential sign
85: \newcommand\usign{\mathrm{sign}} % sign function
86: \newcommand\usin{\mathrm{sin}} % sine function
87: \newcommand\ucos{\mathrm{cos}} % cosine function
88: \newcommand\bu{\boldsymbol{u}}
89: \newcommand\p{\ensuremath{\partial}}
90: \newcommand\D{\ensuremath{\Delta}}
91: \newcommand\eps{\epsilon}
92: \newcommand\Om{\Omega}
93: \newcommand\om{\omega}
94: %
95: % For sans serif characters:
96: % The following macros are setup in JFM.cls for sans-serif fonts in text
97: % and math. If you use these macros in your article, the required fonts
98: % will be substituted when you article is typeset by the typesetter.
99: %
100: % \textsfi, \mathsfi : sans-serif slanted
101: % \textsfb, \mathsfb : sans-serif bold
102: % \textsfbi, \mathsfbi : sans-serif bold slanted (doesn't exist in CM fonts)
103: %
104: % For san-serif Roman use \textsf and \mathsf as normal.
105: %
106: \newcommand\ssC{\mathsf{C}} % for sans serif C
107: \newcommand\sfsP{\mathsfi{P}} % for sans serif sloping P
108: \newcommand\slsQ{\mathsfbi{Q}} % for sans serif bold-sloping Q
109:
110: % Hat position
111: \newcommand\hatp{\skew3\hat{p}} % p with hat
112: \newcommand\hatR{\skew3\hat{R}} % R with hat
113: \newcommand\hatRR{\skew3\hat{\hatR}} % R with 2 hats
114: \newcommand\doubletildesigma{\skew2\tilde{\skew2\tilde{\Sigma}}}
115: % italic Sigma with double tilde
116:
117: % array strut to make delimiters come out right size both ends
118: \newsavebox{\astrutbox}
119: \sbox{\astrutbox}{\rule[-5pt]{0pt}{20pt}}
120: \newcommand{\astrut}{\usebox{\astrutbox}}
121:
122: \newcommand\GaPQ{\ensuremath{G_a(P,Q)}}
123: \newcommand\GsPQ{\ensuremath{G_s(P,Q)}}
124: %\newcommand\p{\ensuremath{\partial}}
125: \newcommand\tti{\ensuremath{\rightarrow\infty}}
126: \newcommand\kgd{\ensuremath{k\gamma d}}
127: \newcommand\shalf{\ensuremath{{\scriptstyle\frac{1}{2}}}}
128: \newcommand\sh{\ensuremath{^{\shalf}}}
129: \newcommand\smh{\ensuremath{^{-\shalf}}}
130: \newcommand\squart{\ensuremath{{\textstyle\frac{1}{4}}}}
131: \newcommand\thalf{\ensuremath{{\textstyle\frac{1}{2}}}}
132: \newcommand\Gat{\ensuremath{\widetilde{G_a}}}
133: \newcommand\ttz{\ensuremath{\rightarrow 0}}
134: \newcommand\ndq{\ensuremath{\frac{\mbox{$\partial$}}{\mbox{$\partial$} n_q}}}
135: \newcommand\sumjm{\ensuremath{\sum_{j=1}^{M}}}
136: \newcommand\pvi{\ensuremath{\int_0^{\infty}%
137: \mskip \ifCUPmtlplainloaded -30mu\else -33mu\fi -\quad}}
138:
139: \newcommand\etal{\mbox{\textit{et al.}}}
140: \newcommand\etc{etc.\ }
141: \newcommand\eg{e.g.\ }
142:
143:
144: \newtheorem{lemma}{Lemma}
145: \newtheorem{corollary}{Corollary}
146:
147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
148: % end JFM caption
149: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
150:
151: %\def\baselinestretch{2}
152:
153: \title[Flow induced by a randomly ...]
154: {Flow induced by a randomly vibrating boundary}
155: \author[D. Volfson and J. Vi\~nals]{DMITRI VOLFSON$^{1}$ AND JORGE
156: VI\~NALS$^{1,2}$}
157:
158: \affiliation{$^1$Supercomputer Computations Research Institute,
159: Florida State University, Tallahassee, Florida 32306-4130, USA\\
160: $^{2}$ Department of Chemical Engineering,
161: FAMU-FSU College of Engineering, Tallahassee, Florida
162: 31310-6046, USA}
163: \pubyear{1999}
164: \volume{000}
165: \pagerange{000--000}
166: \date{\today}
167: \setcounter{page}{1}
168:
169: \begin{document}
170: \maketitle
171:
172: \begin{abstract}
173: We study the flow induced by random vibration of a solid
174: boundary in an otherwise quiescent fluid. The analysis is motivated by
175: experiments conducted under the low level and random effective
176: acceleration field that is
177: typical of a microgravity environment. When the boundary is planar and
178: is being vibrated along its own plane, the variance of the velocity field
179: decays as a power law of
180: distance away from the boundary. If a low frequency cut-off is introduced in
181: the power spectrum of the boundary velocity, the variance
182: decays exponentially for distances larger than a Stokes layer thickness
183: based on the cut-off frequency. Vibration of a gently curved boundary
184: results in steady streaming in the ensemble average of the tangential
185: velocity. Its amplitude diverges logarithmically with distance away
186: from the boundary, but asymptotes to a constant value instead
187: if a low frequency cut-off is considered. This steady component of the
188: velocity is shown to depend
189: logarithmically on the cut-off frequency. Finally, we consider the case
190: of a periodically modulated solid boundary that is being randomly vibrated.
191: We find steady streaming in the ensemble average of the first order velocity,
192: with flow extending up to a characteristic distance of the order of the
193: boundary wavelength. The structure of the flow in the vicinity of the
194: boundary depends strongly on the correlation time of the boundary velocity.
195: \end{abstract}
196:
197: \section{Introduction}
198: \label{sec:introduction}
199:
200: This paper examines the formation and separation of viscous layers
201: in a fluid which is in contact with a
202: solid boundary that is vibrated randomly. The analysis is motivated by the low
203: level and random acceleration field that affects a number of microgravity
204: experiments. We first study the case of a planar boundary to
205: generalize the classical result of \cite{re:stokes51} who considered
206: a boundary vibrated periodically along its own plane.
207: We next consider a slightly curved boundary, and show that steady streaming
208: appears in the ensemble average at first order in the perturbed flow
209: variables. There are
210: several qualitative similarities and differences with the classical result by
211: \cite{re:schlichting79} for the case of periodic vibration. Finally, we
212: address the case of a modulated boundary that is vibrated randomly.
213:
214: Our study is motivated by the significant levels of residual accelerations
215: (g-jitter) that have been detected during space missions in which microgravity
216: experiments have been conducted
217: (\cite{re:walter87,re:nelson91,re:delombard97}).
218: Direct measurement of these residual accelerations has shown that they have a
219: wide frequency spectrum, ranging approximately from $10^{-4} Hz$ to $10^{2}
220: Hz$. Amplitudes range from $10^{-6}g_{E}$ at the lowest end of
221: the frequency spectrum, and increase roughly linearly for high
222: frequencies, reaching values of $10^{-4}g_{E}\;-\;10^{-3}g_{E}$
223: at frequencies of around $10 Hz$ ($g_{E}$ is the intensity of
224: the gravitational field on the Earth's surface). Despite the efforts of a
225: number of researchers over the last decade, there remain areas of uncertainty
226: about the potential effect of such a residual acceleration field on
227: typical microgravity fluid experiments, especially in quantitative terms.
228: A better understanding of the response of a fluid to such
229: disturbances would enable improved experiment design to minimize
230: or compensate for their influence. In addition, it would also be useful to
231: have error estimates of quantities measured in the presence of residual
232: accelerations, including whenever possible some methodology for extrapolation
233: to ideal zero gravity.
234:
235: The formation of viscous layers around solid boundaries
236: when the flow amplitude has a random component has not been addressed yet
237: despite its potential relevance for a number of microgravity experiments.
238: Among them we
239: mention the dynamics of colloidal suspensions, coarsening studies of
240: solid-liquid mixtures in which purely diffusive controlled transport
241: is desired,
242: or the interaction between the viscous layer produced by bulk flow of random
243: amplitude and the morphological instability of a crystal-melt interface.
244: Our study represents the first step in this direction, and focuses on simple
245: geometries in order to elucidate those salient features of the flow that arise
246: from the random nature of the vibration.
247:
248: Previous theoretical work on the influence of g-jitter on fluid flow
249: ranges from order of magnitude estimates to
250: detailed numerical calculations. For example, the order of magnitude of
251: the contributions to fluid flow from the residual acceleration field
252: may be estimated from the length and time scales of a particular
253: experiment, and the values of the relevant set of dimensionless numbers
254: (\cite{re:alexander90}). Such studies are of interest as a first approximation,
255: but are not very accurate. Other studies have modeled the residual
256: acceleration field by some simple analytic function in which the acceleration
257: is typically decomposed into steady and time dependent components, the latter
258: being periodic in time
259: (\cite{re:gershuni76,re:kamotani81,re:alexander91,re:farooq94,re:grassia98a,re:grassia98b,
260: re:gershuni98}).
261: A few studies have also addressed the consequences of isolated pulses
262: of short duration (\cite{re:alexander97}).
263:
264: \cite{re:zhang93} and \cite{re:thomson97} adopted a statistical
265: description of the residual acceleration field onboard spacecraft, and modeled
266: the acceleration time series as a stochastic process in time.
267: The main premise of this approach is that a statistical
268: description is necessary in those cases in which the characteristic
269: time scales of the physical process under investigation are long
270: compared with the correlation time of $g$-jitter, $\tau$ (the acceleration
271: amplitudes and orientations at two different times are statistically
272: independent if separated by an interval larger than $\tau$).
273: Progress has been achieved through the consideration of a specific stochastic
274: model according to which each Cartesian component of the residual acceleration
275: field $\vec{g}(t)$ is modeled as a narrow band noise. This noise is a Gaussian
276: process defined by three independent parameters: its intensity
277: $\left<g^2\right>$, a dominant angular frequency $\Omega$, and a
278: characteristic spectral width $\tau^{-1}$. Each realization of narrow band
279: noise can be viewed as a temporal sequence of periodic functions of angular
280: frequency $\Omega$ with amplitude and phase that remain constant only for a
281: finite amount of time ($\tau$ on average). At random intervals, new values of
282: the amplitude and phase are drawn from prescribed distributions. This model is
283: based on the following mechanism underlying the residual acceleration field:
284: one particular natural frequency of vibration of the spacecraft structure
285: ($\Omega$) is excited by some mechanical disturbance inside the spacecraft,
286: the excitation being of random amplitude and taking place at a sequence of
287: unknown (and essentially random) instants of time.
288:
289: Narrow band noise has been
290: shown to describe reasonably well many of the features of g-jitter time
291: series measured onboard Space Shuttle by \cite{re:thomson97}. Actual
292: $g$-jitter data collected during the SL-J mission were analyzed, and a time
293: series of roughly six hours was studied in detail. A scaling analysis revealed
294: the existence of both deterministic and stochastic components in the time
295: series. The deterministic contribution appeared at a frequency of $17$ Hz,
296: with an amplitude $\left<g^2\right>^{1/2}= 3.56 \times 10^{-4} g_{E}$.
297: Stochastic components included two well defined spectral features with a
298: finite correlation time; one at 22 Hz with $\left<g^2\right>^{1/2}= 3.06
299: \times 10^{-4}g_{E}$ and $\tau = 1.09$ s, and one at 44 Hz with
300: $\left<g^2\right>^{1/2}= 5.20 \times 10^{-4} g_{E}$ and $\tau = 0.91$ s. White
301: noise background is also present in the series with an intensity
302: $D = 8.61 \times 10^{-4} {\rm cm}^{2}/{\rm s}^{3}$.
303:
304: A further theoretical advantage of narrow band noise is that it provides a
305: convenient way of
306: interpolating between monochromatic noise (akin to studies involving a
307: deterministic and periodic gravitational field), and white noise (in which
308: no frequency component is preferred). In the limit $ \tau\rightarrow 0$ with
309: $D = \left< g^{2}\right> \tau$ finite, narrow band noise reduces to white noise
310: of intensity $D$; whereas, for $\tau\rightarrow\infty$ with $\left< g^{2}
311: \right>$ finite, monochromatic noise is recovered.
312:
313: We discuss in this paper the flow induced in an otherwise quiescent fluid by
314: the random vibration of a solid boundary. The velocity of the boundary
315: $u_{0}(t)$ is assumed prescribed, and modeled as a narrow band stochastic
316: process. First, we consider an infinite planar boundary that is being vibrated
317: along its own plane to generalize the classical problem studied
318: by \cite{re:stokes51}. In the monochromatic limit, the variance of the
319: velocity field decays exponentially away from the wall, with a
320: characteristic decay
321: length given by the Stokes layer thickness $\delta_s=(2\nu/\Om)^{1/2}$,
322: where $\nu$ is the kinematic viscosity of the fluid, and $\Om$ is the angular
323: frequency of vibration of the boundary. Since the equations governing the flow
324: are linear, we are able to obtain an analytic solution describing transient
325: layer formation in the stochastic case, but only in the neighborhood of the
326: white and monochromatic noise limits. We then show that for any finite
327: correlation time the stationary variance of the tangential velocity
328: asymptotically decays as
329: the inverse squared distance from the wall, in contrast with the exponential
330: decay in the deterministic case. This asymptotic behavior originates from
331: the low frequency range of the power spectrum of the boundary velocity. The
332: crossover from power law to exponential decay is explicitly computed by
333: introducing a low frequency cut-off in the power spectrum.
334:
335: We next investigate two additional geometries in which the equations governing
336: fluid flow are not linear, and show that several of the generic features
337: obtained for the case of a
338: planar boundary still hold. In the first case, we generalize the analysis of
339: \cite{re:schlichting79} concerning secondary steady streaming.
340: He found that the oscillatory motion of the boundary
341: induces a steady secondary flow outside of the viscous boundary layer
342: even when the velocity of the boundary averages to zero. If the thickness of
343: the Stokes layer, $\delta_s$, and the amplitude of oscillation, $a$, are small
344: compared with a characteristic length scale of the boundary $L$
345: ($\delta_s \ll L$, $a\ll L$), then the generation of secondary steady
346: streaming may be described as follows. Vibration of the rigid boundary gives
347: rise to an oscillatory and nonuniform motion of the fluid. The flow is
348: potential in the bulk, and rotational in the boundary layer because of
349: no-slip conditions
350: on the boundary. The bulk flow applies pressure at the outer edge of boundary
351: layer, which does not vary across the layer. The non uniformity
352: of the flow leads to vorticity convection in the boundary layer through
353: nonlinear terms. Both convection and the applied pressure drive
354: vorticity diffusion, and thus induce secondary steady motion which does not
355: vanish outside of the boundary layer. In the simplest case in which the far
356: field velocity is a standing wave $U(x,t)=U(x)\ucos(\Om t)$, the
357: tangential component of the secondary steady velocity is,
358: \begin{equation}
359: u^{(s)}=-\frac{3}{4\Om}U\frac{\ud U}{\ud x},
360: \label{sch0}
361: \end{equation}
362: where $x$ is a curvilinear coordinate along the boundary. In fact, Eq.
363: (\ref{sch0}) serves as the boundary condition for the stationary part of the
364: flow in the bulk. Similar conclusions have been later reached by
365: \cite{re:batchelor67} who studied sinusoidal oscillations of
366: nonuniform phase, and by \cite{re:gershuni98} who studied
367: monochromatic oscillations of a general form.
368:
369: The second geometry that we address is the so called wavy wall
370: (\cite{re:lyne71}). The deterministic limit in which a wavy boundary is being
371: periodically vibrated has been studied by a number of authors, mainly to
372: address the interaction between the flow above the sea bed and ripple
373: patterns on it (\cite{re:lyne71,re:kaneko79,re:vittori89,re:blondeaux94} and
374: references therein). \cite{re:lyne71} deduced the existence of steady
375: streaming in the limit in which the amplitude of the wall deviation
376: from planarity is small compared with the thickness of the Stokes
377: layer. He introduced a conformal transformation and obtained an explicit
378: solution in the limit of small $k\Rey$, where $k$ is the wavenumber of
379: the wall profile scaled by the thickness of the Stokes layer, and
380: $\Rey$ is the Reynolds number. The detailed structure of the secondary
381: flow depends on the ratio between the wavelength of the boundary
382: profile and the thickness of the Stokes layer.
383:
384: In Sections \ref{sec:sch} and \ref{sec:wavy_wall}, we discuss how the results
385: for these two geometries generalize to the case of stochastic vibration.
386: Section \ref{sec:sch} addresses the flow created by a gently curved solid
387: boundary that is being vibrated randomly. The perturbation parameter that we use
388: is the ratio between the amplitude of vibration and the characteristic inverse curvature
389: of the wall. The ensemble average of the stream function is not zero,
390: and hence there exists stationary streaming in the
391: stochastic case as well. The average velocity diverges logarithmically away
392: from the boundary because of the low frequency range of the power spectrum.
393: We again introduce a low frequency cut-off $\om_c$ in the spectrum, and study the
394: dependence of the stationary streaming on the cut-off frequency.
395: We compute the stationary tangential velocity as a function of $\om_c \ll 1$
396: and arbitrary $\beta$, and find a weak (logarithmic) singularity as
397: $\om_c \rightarrow 0$.
398:
399: Section \ref{sec:wavy_wall} discusses the formation of a boundary layer around
400: a wavy boundary that is vibrated randomly. Positive and negative vorticity
401: production in adjacent regions of the boundary introduces a natural
402: decay length
403: in the solution, thus leading to exponential decay of the flow away from the
404: boundary, even in the absence of a low frequency cut-off in the power spectrum
405: of the boundary velocity. Steady streaming is found at second order comprising
406: two or four recirculating cells per period of the boundary profile. The number
407: of cells depends on the scaled correlation time $\Om \tau$.
408:
409: \section{Randomly vibrating planar boundary}
410: \label{sec:plane}
411:
412: We first examine the case of a planar boundary that is being vibrated along
413: its own plane. In this case the governing equations are considerably simpler
414: then in the more general geometries discussed in Sections \ref{sec:sch}
415: and \ref{sec:wavy_wall}. In particular, the Navier-Stokes equation is linear,
416: fact that allows a complete solution of the flow. Nevertheless, this simple
417: solution still exhibits several of the qualitative features that are present in
418: the case of random forcing by a curved boundary, namely asymptotic power law
419: decay of the velocity field away from the boundary, and sensitive dependence
420: on the low frequency range of the power spectrum of the boundary velocity.
421:
422: Consider an infinite solid boundary located at $z=0$, and an
423: incompressible fluid that
424: occupies the region $z > 0$. The Navier-Stokes equation, and boundary
425: conditions are,
426: \begin{equation}
427: \partial_t{u} = \nu\partial_z^2{u},
428: \label{plane1}
429: \end{equation}
430: \begin{equation}
431: u(0,t)=u_0(t), \quad u(\infty,t) < \infty,
432: \label{plane2}
433: \end{equation}
434: where $z$ is the coordinate normal to the boundary, $u(z,t)$ is the $x$
435: component of the velocity, and $u_{0}(t)$ is the prescribed velocity of the
436: boundary. The solution for harmonic vibration $u_0(t)=u_0\ucos(\Om t)$
437: was given
438: by \cite{re:stokes51}. It is a transversal wave that propagates into the bulk
439: fluid with an exponentially decaying amplitude,
440: \begin{equation}
441: u(z,t)=u_0\ue^{-z/\delta_s}\ucos(\Om t - z/\delta_s),
442: \label{stokes}
443: \end{equation}
444: where $\delta_s=(2\nu/\Om)^{1/2}$ is the Stokes layer thickness.
445:
446: \subsection{Narrow band noise}
447:
448: As discussed in the introduction, the main topic of this paper is to
449: examine how the nature of the bulk flow changes when the boundary velocity
450: $u_{0}(t)$ is a random process. Specifically, we consider a Gaussian process
451: defined by
452: \begin{equation}
453: \left< u_0(t)\right>=0, \quad
454: \left<u_{0}(t)u_{0}(t')\right>=\left<{u_0}^{2}\right>
455: \ue^{-|t-t'|/\tau}\ucos\Om(t-t').
456: \label{nbn2}
457: \end{equation}
458: This process is known as narrow band noise (\cite{re:stratonovich67}).
459: It is defined by three independent parameters: its variance
460: $\left<u_0^2\right>$, its dominant angular frequency $\Om$, and the
461: correlation time $\tau$. Each realization of this random process can be viewed
462: as a sequence periodic functions of frequency $\Om$, with amplitude and phase
463: that remain constant for a time interval $\tau$ on average. White noise is
464: recovered when $\Om\tau\rightarrow 0$ while $D = \left<{u_0}^{2}\right>\tau$
465: remains finite, whereas the monochromatic noise limit corresponds to
466: $\Om\tau\rightarrow \infty$, with $\left<{u_0}^{2}\right>$ finite.
467: Monochromatic noise is akin to a single frequency periodic signal of the same
468: frequency, but with randomly drawn amplitude and phase. The relationship
469: between the two can be illustrated by considering a deterministic function
470: $x(t)=x_0\ucos(\Om t)$ and defining the temporal average as,
471: \begin{equation}
472: \left<x(t)x(t')\right> =
473: \lim_{T\rightarrow\infty}\frac{1}{T}\int_0^T\ud t\,x(t)x(t')=
474: \frac{x_0^2}{2}\ucos(\Om(t-t')).
475: \label{nbn21}
476: \end{equation}
477: This average coincides with the ensemble average of the noise when
478: $\left<u_0^2\right>=x_0^2/2$. The power spectrum corresponding to the
479: correlation function (\ref{nbn2}) is
480: \begin{equation}
481: P(\omega)=\frac{\left<u_0^2\right>}{2\upi}\left[\frac{\tau}
482: {1+\tau^2(\omega-\Omega)^2}+\frac{\tau}{1+\tau^2(\omega+\Omega)^2}\right].
483: \label{nbn3}
484: \end{equation}
485: We will also use the spectral density of the process $u_0(t)$,
486: \begin{equation}
487: \hat{u}_0(\om)=\frac{1}{2\upi}\int_{-\infty}^{\infty}
488: \ud t\,u_0(t)\ue^{-\ui\om t},
489: \label{spec_dens}
490: \end{equation}
491: so that its ensemble average and correlation function are respectively given by
492: \refstepcounter{equation}
493: $$
494: \left<\hat{u}_0(\om)\right>=0,\quad
495: \left<\hat{u}_0(\om)\hat{u}^{*}_{0}(\om')\right>=\delta(\om-\om')P(\om).
496: \eqno{(\theequation{\mathit{a},\mathit{b}})}
497: \label{spec_dens_cor}
498: $$
499: We will often use dimensionless variables in which $\left<u_0^2\right>/\Om$
500: is the scale of $P(\om)$, and $\Om$ is the angular frequency scale. In
501: dimensionless form,
502: \begin{equation}
503: P(\omega,\beta)=\frac{1}{2\upi}\left[\frac{\beta}{1+\beta^2(\omega-1)^2}+
504: \frac{\beta}{1+\beta^2(\omega+1)^2}\right],
505: \label{nbn6}
506: \end{equation}
507: where $\beta=\Omega\tau$.
508: We have $\int_{-\infty}^{\infty}\ud\om P(\om,\beta)=1$, independent of $\beta$,
509: and also $\lim_{\beta \rightarrow \infty} P(\omega)
510: = \left[\delta(\omega-1)+\delta(\omega+1)\right]/2$.
511: Note that the power spectrum does not vanish at small frequencies. Instead,
512: $P(0,\beta)=\beta/(\upi(1+\beta^2)$, which for large and small $\beta$
513: behaves as $P(0,\beta)\sim 1/(\upi \beta)$ and $P(0,\beta)\sim \beta/\upi$
514: respectively. We will discuss separately the effect of this
515: low frequency contribution on the results presented in the remainder of the
516: paper.
517:
518: \subsection{Transient layer formation}
519: \label{subsec:transient}
520:
521: In the two limiting cases of white and monochromatic noise, it is possible
522: to find an analytic solution for the transient flow starting from an initially
523: quiescent fluid. The solution can be found, for example, by introducing
524: the retarded, infinite space Green's function corresponding to
525: equation (\ref{plane1}), with boundary conditions (\ref{plane2}),
526: \begin{equation}
527: G(z,t|z',t')=\frac{1}{(4\upi\nu(t-t'))^{1/2}}\left[
528: \ue^{-(z-z')^2/4\nu(t-t')}-\ue^{-(z+z')^2/4\nu(t-t')}\right],
529: \quad t>t',
530: \label{greenplane}
531: \end{equation}
532: and $G(z,t|z',t')=0$ for $t<t'$. If the fluid is initially quiescent,
533: $u(z,0)=0$, we find
534: \begin{equation}
535: u(z,t)=\nu\int_{0}^{t}\ud t'\,u_0(t')\left(\p_{z'}G\right)_{z'=0},
536: \label{quis_sol}
537: \end{equation}
538: with
539: \begin{equation}
540: \left(\p_{z'}G\right)_{z'=0}=\frac{z}{[4\upi\nu^3(t-t')^3]^{1/2}}\,
541: \ue^{-z^2/4\nu(t-t')}.
542: \label{dif_green}
543: \end{equation}
544: Equations (\ref{quis_sol}-\ref{dif_green}) determine the transient behavior for
545: any given $u_0(t)$.
546:
547: If $u_0(t)$ is a Gaussian, white noise process, the ensemble average of
548: Eq.(\ref{quis_sol}) yields $\left<u(z,t)\right>=0$. The corresponding equation
549: for the variance reads
550: \begin{equation}
551: \left<u^2(z,t)\right>=2D\nu^2\int_0^t\ud t'\,\left[\left(\p_{z'}G\right)_{z'=0}
552: \right]^2=
553: \frac{2D\nu}{\upi z^2}\left(1+\frac{z^2}{2\nu t}\right)\,\ue^{-z^2/2\nu t}.
554: \label{wn_var}
555: \end{equation}
556: The variance of the induced fluid velocity propagates into the fluid
557: diffusively.
558: Saturation occurs for $t\gg z^2/2\nu$, at which point the variance does
559: not decay exponentially far away from the wall, but rather as a power law.
560: \begin{equation}
561: \left<u^2(z,\infty)\right>=\frac{2D\nu}{\upi z^2}.
562: \label{wn_var_longtime}
563: \end{equation}
564: Ascertaining whether random vibration can induce flows far away from
565: the boundary in more general geometries is one of the main
566: motivations for this paper.
567:
568: Consider now the opposite limit of monochromatic noise with
569: correlation function
570: \begin{equation}
571: \left<u_0(t)u_0(t')\right>=\left<u_0^2\right>\ucos\left[\Om(t-t')\right] .
572: \label{monocf}
573: \end{equation}
574: Now using Eqs. (\ref{quis_sol}) and (\ref{monocf}) we find,
575: (\cite{re:carslaw59}),
576: \begin{equation}
577: \frac{\left<u^2(z,t)\right>}{2\left<u_0^2\right>}=\frac{2}{\upi}
578: \int_{\kappa}^{\infty}\ud\sigma\,\ue^{-\sigma^2}
579: \int_{\kappa}^{\infty}\ud\mu\,\ue^{-\mu^2}
580: \ucos\left[\frac{z^2}{2\delta_s^2}\left(\frac{1}{\mu^2}-\frac{1}{\sigma^2}
581: \right)\right],
582: \label{monovar}
583: \end{equation}
584: with $\kappa = z/(4\nu t)^{1/2}$. A closed form solution can only be obtained
585: for long times. We find,
586: \begin{equation}
587: \frac{\left<u^2(z,t)\right>}{2\left<u_0^2\right>}=\frac{\ue^{-2z/\delta_s}}{2}
588: +\frac{2\kappa^3\delta_s^2}{\upi^{1/2}z^2}\ue^{-z/\delta_s}\usin(\Om\,t-z/
589: \delta_s)+{\cal O}(\kappa^5(t)).
590: \label{monovarcor}
591: \end{equation}
592: At long times, the variance propagates into the bulk with phase velocity
593: $(2\nu\Om)^{1/2}$, while its amplitude decays exponentially in space over the
594: scale of the Stokes layer, and as $t^{-3/2}$ in time. In summary, the flow
595: created by the vibration of the boundary propagates diffusively for
596: white noise $(z^2\propto 2\nu t)$, and as a power law
597: $(z^2\propto \upi\nu\Om^2 t^3)$ in the monochromatic limit.
598:
599: \subsection{Stationary variance for narrow band noise}
600:
601: We have been unable to obtain a closed analytic solution for the transient
602: evolution of the variance $\left< u(z,t)^{2} \right>$ when the
603: vibration of the
604: boundary is given by a general narrow band process. It is possible,
605: however, to obtain the stationary variance of the velocity. Equation
606: (\ref{plane1}) can be rewritten in Fourier space as
607: \begin{equation}
608: \ui\om\hat{u}(z,\om)=\nu\p^2_z\hat{u}(z,\om)
609: \label{fourier2}
610: \end{equation}
611: with
612: \begin{equation}
613: u(z,t)=\int_{-\infty}^{\infty}\ud\om\,\hat{u}(z,\om)\ue^{\ui\om t},
614: \label{fourier1}
615: \end{equation}
616: The boundary conditions are,
617: $\hat{u}(0,\om)=\hat{u}_0(\om)$, and $\hat{u}(z,\om)<\infty$ at $z \rightarrow
618: \infty$. The solution of Eq. (\ref{fourier2}) with these boundary conditions is
619: \begin{equation}
620: \hat{u}(z,\om)=\ue^{-\alpha z}\hat{u}_0(\om),
621: \quad \alpha=(1+\ui\,\usign(\om))(\om/2\nu)^{1/2}
622: \label{fourier4}
623: \end{equation}
624:
625: We next choose $1/\Om$ as the time scale, and $\delta_s$ as the length scale,
626: and after some straightforward algebra we find,
627: \begin{equation}
628: \frac{\left<u^2(z,\beta)\right>}{2 \left<u_{0}^{2}\right>} = I(z,\beta)=
629: \int_{0}^{\infty} \ud\om P(\om,\beta)e^{-z\om^{1/2}}.
630: \label{plane3}
631: \end{equation}
632: We have also used the fact that the power spectrum (\ref{nbn6}) is even
633: in frequency.
634:
635: We next analyze the asymptotic dependence of $I(z,\beta)$ at large $z$.
636: In this limit, $I(z,\beta)$ mainly depends on the low frequency region of the
637: power spectrum; higher frequencies are suppressed by the exponential factor.
638: By using Watson's lemma (\cite{re:nayfeh81}), we find,
639: \begin{equation}
640: I(z,\beta)=\frac{2\beta}{\upi(1+\beta^2)}\frac{1}{z^2} +
641: \frac{240\beta^3(3\beta^2-1)}{\upi(1+\beta^2)^3}\frac{1}{z^6} +
642: {\cal O}(z^{-10}).
643: \label{plane4}
644: \end{equation}
645: This asymptotic form at large $z$ is valid for all $\beta$.
646: In particular, the dominant behavior for small and large $\beta$ is
647: \begin{displaymath}
648: I(z,\beta) \sim \frac{2\beta}{\upi}\frac{1}{z^2},\quad
649: I(z,\beta) \sim \frac{2}{\upi\beta}\frac{1}{z^2},
650: \end{displaymath}
651: respectively. Hence we recover the power law decay of Eq.
652: (\ref{wn_var_longtime}).
653:
654: We can also find the asymptotic behavior at large $\beta$ that is
655: uniformly valid in $z$,
656: \begin{equation}
657: I(z,\beta)=\frac{\ue^{-z}}{2}-\frac{z}{2\upi\beta}\left(
658: \Ci(z)\sin(z)-\Si(z)\cos(z)-\frac{1}{2}(\ue^{-z}\Ei(z)-\ue^z\Ei(-z))
659: \right) + {\cal O}(\beta^{-3}),
660: \label{plane5}
661: \end{equation}
662: where $\Ci$ and $\Si$ denote the integral sine and cosine functions,
663: and $\Ei$ stands for the exponential integral function
664: (\cite{re:gradshteyn80}).
665: For $z\lesssim 1$, the variance decreases exponentially. At larger $z$,
666: the exponential terms in (\ref{plane5}) become small, so that the remaining
667: asymptotic dependence for large $z$ is given by Eq. (\ref{plane4}).
668: The quantity $I(z,\beta)\,z^2$ computed both from (\ref{plane3}) and the
669: uniform expansion (\ref{plane5}) is presented in Fig.\ref{fig:plane_var}.
670: For fixed $\beta$, $I(z,\beta)$ asymptotes to $2\beta/\upi(1+\beta^2)$
671: outside of Stokes layer. This value is the coefficient of the
672: leading term in the asymptotic expansion (\ref{plane4}). The expansion
673: (\ref{plane5}) is a good approximation even for moderate values of $\beta$.
674:
675: To summarize, the variance of the velocity field does not decay
676: exponentially away from the wall for finite $\beta$, but rather as the
677: inverse squared distance. The crossover length separating exponential
678: and power law decay increases with increasing $\beta$.
679:
680: \subsection{Low frequency cut-off in the power spectrum}
681: \label{subsec:cutplane}
682:
683: The coefficient of the leading term in (\ref{plane4}) is in fact
684: twice the value of $P(0,\beta)=\beta/\upi(1+\beta^2)$. The algebraic decay of
685: $\left<u_0^2(z,t)\right>$ follows from the diffusive nature of Eq.
686: (\ref{plane1}), and a non vanishing value of $P(\om,\beta)$ as $\beta
687: \rightarrow 0$. Before we analyze in Sections \ref{sec:sch} and
688: \ref{sec:wavy_wall} how this behavior is modified by nonlinearities in the
689: governing equations, we explicitly address here the consequences of a low
690: frequency cut-off in the power spectrum. Of course, there always exists in
691: practice a low frequency cut-off because of limited observation time.
692: Furthermore, the low frequency range of the power spectrum of the residual
693: acceleration field in microgravity ($\Om/2 \upi < 10^{-3}$Hz) is
694: fairly difficult to measure reliably. We therefore introduce an
695: effective cut-off frequency in the power spectrum, $\om_c\ll1$, and
696: study the dependence of $\left<u_0^2(z,t)\right>$
697: on $\om_c$. The stationary value of variance of the velocity is now given by,
698: \begin{equation}
699: \frac{\left<u^2(z,\beta,\om_c)\right>}{2\left<u_0^2\right>}
700: = I_c(z,\beta,\om_c)=\int_{\om_c}^{\infty}
701: \ud\om P(\om,\beta)e^{-z\om^{1/2}}.
702: \label{plane1_cut}
703: \end{equation}
704: By using Watson's lemma, we find for large $z$,
705: \begin{equation}
706: I_c(z,\beta,\om_c)=\frac{2\beta\ue^{-z\om_c^{1/2}}}{\upi(1+\beta^2)}
707: \left(\frac{1}{z^2}+\frac{\om_c^{1/2}}{z}+\uhot\right),
708: \label{plane2_cut}
709: \end{equation}
710: where $\uhot$ stands for terms which are of higher order than terms
711: retained under the assumption that both $1/z$ and $\om_c^{1/2}$
712: are small but independent. For $z \gg 1$, but $z \om_c^{1/2} \ll 1$
713: the dominant term in (\ref{plane2_cut}) is
714: \begin{equation}
715: I_c(z,\beta,\om_c)\sim
716: \frac{2\beta}{\upi(1+\beta^2)}\frac{1}{z^2},\quad z\om_c^{1/2}<<1.
717: \label{plane3_cut}
718: \end{equation}
719: On the other hand, if $z\om_c^{1/2}\geq 1$, the leading order term is now
720: a function of $\zeta=z\om_c^{1/2}$
721: \begin{equation}
722: I_c(z,\beta,\om_c)\sim
723: \frac{2\beta\om_c\ue^{-\zeta}}{\upi(1+\beta^2)}
724: \left(\frac{1}{2\zeta}+\frac{1}{\zeta^2}\right),\quad \zeta\geq 1
725: \label{plane4_cut}
726: \end{equation}
727: Equations (\ref{plane3_cut}) and (\ref{plane4_cut}) show that at
728: distances that are large compared with the thickness of the Stokes
729: layer based on the dominant frequency $\Om$, $\left<u^2(z,t)\right>$
730: decays algebraically with $z$. There exists, however, a length scale
731: $z\om_{c}^{1/2}$ beyond which the decay is exponential.
732: This new characteristic length scale is the thickness of the Stokes
733: layer based on the cut-off frequency. This conclusion appears natural
734: given the principle of superposition for the linear differential
735: equation (\ref{plane1}).
736:
737: \section{Streaming due to random vibration}
738: \label{sec:sch}
739:
740: Next we investigate to what extent the results of Section \ref{sec:plane} hold
741: in configurations in which the governing equations are not linear. We examine
742: in this section the flow induced by a gently curved solid boundary
743: that is being randomly vibrated. The boundary velocity is assumed to be
744: described by a narrow band stochastic process, and hence our results will
745: reduce to Schlichting's in the limit of infinite correlation time.
746: However, for finite values of $\beta$ the results are qualitatively different.
747: The mechanism of secondary steady streaming generation described in the
748: introduction is no longer valid because there is no boundary layer solution at
749: zeroth order. Vorticity produced at the vibrating boundary penetrates
750: into the bulk fluid, at least perturbatively for small curvature.
751: This results in a logarithmic divergence of the ensemble average of
752: the first order velocity with distance away from the wall. A cut-off
753: analysis is also presented, and similarly to that of Section
754: \ref{subsec:cutplane}, it reveals the existence of an effective boundary
755: layer of thickness based on the cut-off frequency. We then find
756: an expression analogous to Eq. (\ref{sch0}) as a function of $\beta$ and
757: $\om_c$.
758:
759: Define the following dimensionless quantities,
760: \begin{equation}
761: \left.
762: \begin{array}{l}
763: \displaystyle
764: z=\tilde{z}[(\nu/\Om)^{1/2}],\quad x=\tilde{x}[L],\quad
765: t=\tilde{t}[\Om^{-1}],\quad\psi=\tilde{\psi}[(2\left<u_0^2\right>
766: \nu/\Om )^{1/2}] , \\[16pt]
767: \displaystyle
768: \quad \eps=(2\left<u_0^2\right>)^{1/2}/\Om L,\quad \Rey_p=2\left<u_0^2\right>/
769: \Om\nu,
770: \quad\tilde{\Delta}=\p_{\tilde{z}}^2+\eps^2/\Rey_p\p_{\tilde{x}}^2
771: \end{array}
772: \right\}
773: \end{equation}
774: Assume now that the characteristic scale of the boundary $L$ is large so that
775: $\eps$ is a small quantity. If the Reynolds number $Re_{p}$ is assumed to
776: remain finite, both conditions imply $\nu / \Om L^{2} \ll 1$. We next write the
777: governing equations and boundary conditions in the frame of reference co-moving
778: with the solid boundary and obtain for a two dimensional geometry (tildes are
779: omitted),
780: \begin{equation}
781: \partial_t{\Delta\psi}+\epsilon\frac{\p(\psi,\Delta\psi)}{\p(z,x)}=
782: \Delta^2\psi
783: \label{sch1}
784: \end{equation}
785: \refstepcounter{equation}
786: $$
787: \psi=0, \quad \p_z\psi=0 \quad \mbox{at}\quad y=0,\qquad
788: \eqno{(\theequation{\mathit{a},\mathit{b}})}
789: \label{sch2}
790: $$
791: \begin{equation}
792: \p_z\psi=2^{-1/2}U(x)u_0(t) \quad \mbox{at}\quad z=\infty,
793: \label{sch3}
794: \end{equation}
795: where $x,z$ are the tangential and normal coordinates along the boundary,
796: and $\psi$ is the stream function $\bu=(\p_z\psi,-\p_x\psi)$. We have also
797: used the notation $\p(a,b)/\p(z,x) = (\p_za)(\p_xb)-(\p_xa)(\p_zb)$ for the
798: nonlinear term. The far field boundary condition is a nonuniform and random
799: velocity field, of the order of $\left<u_0^2\right>^{1/2}$, with a spatially
800: nonuniform amplitude $U(x)$, and a stochastic modulation $u_0(t)$ which is a
801: Gaussian stochastic process with zero mean, and narrow band power spectrum.
802: We first expand the stream function as a power series of $\eps$,
803: $ \psi=\psi_0+\eps\psi_1+\ldots $ and solve (\ref{sch1}) order by
804: order. At order $\eps^0$ we obtain the following equation,
805: \begin{equation}
806: (\p_t{\p_z^2}-\p_z^4)\psi_0=0
807: \label{sch5}
808: \end{equation}
809: with boundary conditions,
810: \refstepcounter{equation}
811: $$
812: \psi_0=0, \quad \p_z\psi=0 \quad \mbox{at}\quad z=0,\qquad
813: \eqno{(\theequation{\mathit{a},\mathit{b}})}
814: \label{sch6}
815: $$
816: \begin{equation}
817: \p_z\psi=2^{-1/2}U(x)u_0(t) \quad \mbox{at}\quad z=\infty .
818: \label{sch7}
819: \end{equation}
820: At this order, the equations effectively describe the flow induced
821: above a planar boundary with a far field velocity boundary condition
822: that is not uniform. The solution can be found by Fourier
823: transformation. We define,
824: \begin{eqnarray}
825: \psi_0(x,z,t)=\int_{-\infty}^{\infty}\ud\om\,\hat{\psi}_0(x,z,\om)
826: \ue^{\ui\om t}.
827: \label{sch8}
828: \end{eqnarray}
829: The transformed Eq. (\ref{sch5}) and the transformed boundary conditions
830: (\ref{sch6}) allow separation of variables. We define,
831: $$
832: \hat{\psi}_0(x,z,\om)=2^{-1/2}U(x)\hat{u}_0(\om)\hat{\zeta}_0(z,\om),
833: $$
834: so that Eq. (\ref{sch5}) leads to,
835: \begin{equation}
836: (\ui\om{\p_z^2}-\p_z^4)\hat{\zeta}_0=0,
837: \label{sch51}
838: \end{equation}
839: with boundary conditions
840: $\hat{\zeta}_0=0$, $\p_z\hat{\zeta}_0=0$, at $z=0$ and
841: $\p_z\hat{\zeta}_0=1$ at $z=\infty$. The solution is,
842: \begin{equation}
843: \hat{\zeta}_0(z,\om)=-1/\alpha+z+1/\alpha\ue^{-\alpha z},\quad
844: \alpha(\om) = (1+\ui\,\usign(\om))(|\om|/2)^{1/2}).
845: \label{sch81}
846: \end{equation}
847: %Note that $\hat{\zeta}_0$ grows linearly with $z$ at large $z$.
848:
849: At order $\eps$ we find,
850: \begin{equation}
851: (\p_t{\p_z^2}-\p_z^4)\psi_1=\p_x\psi_0\p_z^3\psi_0 -
852: \p_z\psi_0\p_x\p_z^2\psi_0
853: \label{sch9}
854: \end{equation}
855: with boundary conditions $ \psi_1=0, \; \p_z\psi_1=0$ at $y=0$.
856: The remaining boundary condition for $\psi_1$ needs to be discussed separately.
857: Consider first the classical deterministic limit which can be
858: formally obtained by setting $\beta=\infty$. Then, the right hand side of
859: Eq. (\ref{sch9}) involves stationary terms (of zero frequency), and sinusoidal
860: terms with twice the frequency of the far field flow. Since the equation is
861: linear, the solution $\psi_1$ has exactly the
862: same temporal behavior. In this case, it is known that it is not possible to
863: find a solution for $\p_z\psi_{1}$ that vanishes at large $z$, but
864: only one that simply remains bounded as $z \rightarrow \infty$. By
865: analogy, we introduce a similar requirement in the stochastic case of
866: $\beta < \infty$. Since the
867: zeroth order stream function diverges linearly, this condition simply
868: amounts to requiring that the expansion in powers of $\eps$ remains consistent.
869:
870: We now take the ensemble average of Eq. (\ref{sch9}), and consider the long
871: time stationary limit of the average
872: ($\psi_1^{(s)}=\lim_{t\rightarrow\infty}\left<\psi_1\right>$), to find,
873: \begin{equation}
874: \p_z^4\psi_1^{(s)}=\left<\p_z\psi_0\p_x\p_z^2\psi_0-\p_x\psi_0\p_z^3\psi_0
875: \right>.
876: \label{sch101}
877: \end{equation}
878: The right hand side of this equation can be integrated from infinity
879: to $z$. We obtain,
880: \begin{equation}
881: \p_z^3\psi_1^{(s)}=\frac{U}{2}\frac{\ud U}{\ud x}F(z,\beta),
882: \label{sch11}
883: \end{equation}
884: where
885: $$
886: F(z,\beta)=\int_{0}^{\infty}\ud\om P(\om,\beta)Q(\om,z),
887: $$
888: and
889: $$
890: Q(\om,z)= (-2+2\hat{\zeta}_0\p_z\hat{\zeta}_0^* -
891: \hat{\zeta}_0\p_z^2\hat{\zeta}_0^* -
892: \hat{\zeta}_0^*\p_z^2\hat{\zeta}_0).
893: $$
894: The power spectrum $P(\om,\beta)$ is defined in Eq. (\ref{nbn6}). The constant
895: that appears in the expression for $Q(\om,z)$ comes from the pressure gradient
896: imposed at infinity.
897:
898: We now proceed to solve Eq.(\ref{sch11}) subject to the boundary conditions
899: $ \psi_1^{(s)}=\p_z\psi_1^{(s)}=0$ on the solid boundary, and
900: $\p_z\psi_1^{(s)}<\infty$ as $z\rightarrow \infty$.
901: This is a boundary value problem for $\psi_1^{(s)}$.
902: In the limit $\beta \rightarrow \infty$, it can be solved
903: analytically, and the result obtained by Schlichting is recovered, namely that
904: the solution may be bounded at infinity simply by setting the homogeneous
905: part of the solution equal to zero to satisfy the principle of minimal
906: singularity (\cite{re:vandyke64}). Otherwise $\p_z\psi_1^{(s)}$ is
907: singular a fortiori. Since we cannot find an complete analytic
908: solution for finite $\beta$, we proceed as follows. We recast the
909: boundary value problem as an initial value
910: problem, and search for a boundary condition on $\p_z^2\psi_1^{(s)}$ at $z=0$
911: so that the homogeneous part of the solution remains bounded. This boundary
912: condition can be found analytically by integrating (\ref{sch11}) from
913: $0$ to $z$. We find,
914: $$
915: \p_z^2\psi_1^{(s)}(z)-\p_z^2\psi_1^{(s)}(0)=
916: \frac{U}{2}\frac{\ud U}{\ud x}\int_0^{z}\ud z'
917: \int_{0}^{\infty}\ud\om P(\om,\beta)Q(\om,z'),
918: $$
919: or after changing the order of integration,
920: $$
921: \p_z^2\psi_1^{(s)}(z)-\p_z^2\psi_1^{(s)}(0)=
922: \frac{U}{2}\frac{\ud U}{\ud x}
923: \int_{0}^{\infty}\ud\om P(\om,\beta)
924: \left[
925: \int_z \ud z' Q(\om,z')-\left(\int_z \ud z' Q(\om,z')\right)_{z=0}
926: \right].
927: $$
928: The second integral within brackets equals $(1/2\om)^{1/2}$. In order to
929: avoid a linear divergence of $\p_z\psi_1^{(s)}(z)$ we equate the
930: constant terms on both sides, thus obtaining the third initial condition
931: \begin{eqnarray}
932: \p_z^2\psi_1^{(s)}(0,\beta) & = &
933: U\frac{\ud U}{\ud x}
934: \frac{2^{1/2}}{4}\int_0^{\infty}\ud\om\,\om^{-1/2}P(\om,\beta) =
935: \nonumber\\
936: & = & U\frac{\ud U}{\ud x}\frac{(2\beta)^{1/2}}{8q}
937: ((2(q-\beta))^{1/2}+(q+1)^{1/2}+(q-1)^{1/2}),
938: \label{sch12}
939: \end{eqnarray}
940: where $q = (1+\beta^2)^{1/2}$, and the dependence of initial condition
941: on $\beta$ is shown explicitly. Equation (\ref{sch11}) with its
942: original boundary
943: conditions, supplemented with Eq. (\ref{sch12}) is an initial
944: value problem, which we have solved numerically.
945:
946: Before presenting the numerical results, we study the
947: asymptotic behavior of the solution for large $z$ which
948: is determined by the asymptotic form of $F(z,\beta)$ at large $z$.
949: By explicit substitution of the zeroth order solution we find
950: \begin{equation}
951: Q(z,\om)=(-4\ucos(Z)-2Z(\ucos(Z)+\usin(Z))+2\usin(Z))\ue^{-Z}+2\ue^{-2Z},
952: \label{a2}
953: \end{equation}
954: where $Z = z(\om/2)^{1/2}$. The leading contribution to $F(z,\beta)$
955: as $z\rightarrow\infty$ originates from the zero frequency limit of
956: $P(\om,\beta)$. Thus $ F(z,\beta)\sim
957: P(0,\beta)\int_{0}^{\infty}\ud\om Q(\om,z)$.
958: The remaining integral may be easily calculated to yield the
959: asymptotic form,
960: \begin{equation}
961: F(z,\beta)\sim \frac{6\beta^2}{\upi(1+\beta^2)}\frac{1}{z^2} .
962: \label{sch13}
963: \end{equation}
964: Therefore the stationary mean
965: first order velocity $u_1^{(s)}=\p_z\psi_1^{(s)}$ has a logarithmic
966: asymptotic form (see Eq. (\ref{sch11})). This is to be contrasted with
967: the power law decay of the stationary variance for the case of a
968: randomly vibrating planar boundary.
969:
970: Finally, we have numerically calculated $\partial_{z}\psi_{1}^{(s)}(z)$ for a
971: range of values of $\beta$. The results are presented in Fig.
972: \ref{fig:sc_u_prof}. Equation (\ref{sch11}) was integrated numerically
973: with no-slip boundary conditions for $\psi_1^{(s)}$, and Eq.
974: (\ref{sch12}). The numerical results support our conclusion
975: about the logarithmic divergence of $u_1^{(s)}$ with distance for any
976: finite $\beta$. As $\beta$ increases the stationary mean velocity
977: profile approaches a limiting
978: form that corresponding to the monochromatic limit of
979: $\beta \rightarrow \infty$.
980: In this limit we recover the Schlichting result, according to which
981: $\partial_{z} \psi_{1}^{(s)}(z)$ asymptotes to a constant value over a
982: distance of the order of the deterministic Stokes layer.
983:
984: \subsection{Low frequency cut-off in the power spectrum}
985: \label{subsec:cut_sch}
986:
987: In Section \ref{sec:plane}, we showed that a low frequency cut-off in
988: $P(\om,\beta)$ led
989: to an exponential decay of the velocity outside an effective boundary
990: layer of thickness determined by the cut-off frequency. We therefore
991: examine here the consequences of a low frequency cut-off on the
992: divergent behavior of the stationary average of the first order stream
993: function. In order to find the asymptotic form of $\p_z\psi_1^{(s)}$,
994: we first integrate (\ref{sch11}) twice
995: from $z=0$ to $z$. By using the low frequency cut-off defined in Section
996: (\ref{subsec:cutplane}), we write,
997: \begin{equation}
998: \p_z\psi_1^{(s)}(z,\beta,\om_c)=\frac{U}{2}\frac{\ud U}{\ud x}
999: \left[G_c(z,\beta,\om_c)-G_c(0,\beta,\om_c)\right] ,
1000: \label{sch_cut_asym}
1001: \end{equation}
1002: with,
1003: $$
1004: G_c(z,\beta,\om_c)=\int \ud z'\int \ud z' F_c(z',\beta,\om_c),
1005: $$
1006: and,
1007: $$
1008: F_c(z,\beta,\om_c)=\int_{\om_c}^{\infty}\ud\om P(\om,\beta)Q(\om,z).
1009: $$
1010: If we set $\om_c = 0$ and consider the monochromatic limit of $\beta
1011: \rightarrow \infty$, we find that $G_c(z,\infty,0)$ contains an exponential
1012: factor that vanishes at $z\sim{\cal O}(1)$, and that $G_c(0,\infty,0) = 3/4$.
1013: Therefore the Schlichting result is recovered. An explicit form of
1014: $G_c(z,\beta,\om_c)$ for any $\beta$ and $\om_c$ can be obtained
1015: analytically, but it is far too complicated and we do not quote it here.
1016: It has a similar functional dependence as in the cut-off analysis of
1017: the planar boundary, and contains exponential terms involving
1018: $(-z\om_c^{1/2})$. We find that $\p_z\psi_1^{(s)}(\infty,\beta,\om_c)=
1019: -U\ud U/2\ud x\, G_c(0,\beta,\om_c).$ For finite but small $\om_c$, we find,
1020: \begin{equation}
1021: \p_z\psi_1^{(s)}(\infty,\beta,\om_c)=-\frac{3}{4}U\frac{\ud U}{\ud x}
1022: \frac{\beta}{\upi(1+\beta^2)}
1023: \left(2\beta\arctan(\beta)-\ln\left(\frac{\beta^2\om_c^2}{1+\beta^2}\right)
1024: \right) + {\cal O}(\om_c^2)
1025: \label{sch_cut_asym2}
1026: \end{equation}
1027: This asymptotic formula is compared with the numerically computed
1028: value of the tangential mean stationary velocity at large distances
1029: in Figure \ref{fig:usc_cut}. Computations were done as described in the
1030: previous section, and for different values of $\om_c\ll 1$ and
1031: $\beta$. In all cases
1032: $\p_z\psi_1^{(s)}$ reached constant values at large enough $z$ (the numerical
1033: value of infinity, $z_{\infty}$, was chosen so that the change in velocity for
1034: $z\geq z_{\infty}$ was less than a prescribed tolerance). We also checked
1035: that any change in the boundary condition
1036: Eq.(\ref{sch12}) leads to a linear divergence in the
1037: tangential mean stationary velocity, thus confirming the adequacy of this
1038: boundary condition. The figure also shows that the computed values
1039: of $\p_z\psi_1^{(s)}(z=\infty,\beta,\om_c)$ for small $\om_c$ are in a good
1040: agreement with (\ref{sch_cut_asym2}).
1041:
1042: In summary, Eq. (\ref{sch_cut_asym2}) shows that the tangential velocity away
1043: from the boundary asymptotes to a constant that is a function of
1044: $\beta$, and
1045: has a weak (logarithmic) dependence on the cut-off frequency $\om_c$.
1046: Therefore the asymptotic dependence in the stochastic case (with a cut-off)
1047: and the deterministic case is
1048: qualitatively similar, although the value of the asymptotic velocity of the
1049: former depends on $\beta$. Note also that this asymptotic behavior only
1050: sets in for distances larger than $(\nu/\om_{c}\Om)^{1/2}$, a value that can be
1051: quite large in practical microgravity conditions.
1052:
1053: \section{Randomly vibrating wavy boundary}
1054: \label{sec:wavy_wall}
1055:
1056: In the two previous sections we considered cases in which the characteristic
1057: longitudinal length scale of the solid boundary was either infinite or large
1058: compared with
1059: the characteristic amplitude of boundary vibrations. We examine here
1060: the case of a wavy boundary and study how comparable length scales in
1061: both directions influence the flow away from the boundary. In contrast
1062: with the Schlichting problem, the external applied flow is now uniform
1063: or, alternatively, the length scale over which the flow is not uniform
1064: is much larger that the wavelength of the boundary. Thus one
1065: anticipates that the normal component of the flow that appears
1066: is caused by the wall profile. This flow interacts through
1067: nonlinear terms with the externally forced flow that is parallel
1068: to the average profile of the boundary and gives rise
1069: to stationary streaming. Even for stochastic vibration we show that positive
1070: and negative vorticity production in
1071: adjacent regions of the boundary introduces a natural decay length in the
1072: zeroth order solution, thus leading to exponential decay of the flow
1073: away from the boundary, even in the absence of a low frequency cut-off in the
1074: power spectrum of the boundary velocity.
1075:
1076: Consider a rigid wavy wall being washed by a uniform oscillatory flow parallel
1077: to the wall wave vector,
1078: \begin{equation}
1079: \mathbf{u}(x,z=\infty)=(u_0(t),0).
1080: \label{rvwb_uinf}
1081: \end{equation}
1082: We now assume that $u_0(t)$ is a narrow band Gaussian process.
1083: Assume also that the amplitude of the boundary profile $l$ is small
1084: compared with both the Stokes layer $\delta_s$ and the
1085: wavelength $L$, with $\delta_{s}/L$ finite. The following dimensionless
1086: quantities are introduced,
1087: \begin{equation}
1088: \left.
1089: \begin{array}{l}
1090: \displaystyle
1091: z=\tilde{z}[(\nu/\Om)^{1/2}],\quad x=\tilde{x}[(\nu/\Om)^{1/2}],\quad
1092: t=\tilde{t}[\Om^{-1}],\quad\psi=\tilde{\psi}[(2\left<u_0^2\right>
1093: \nu/\Om )^{1/2}], \\[16pt]
1094: \displaystyle
1095: \eps=l/(\nu/\Om)^{1/2},\quad \Rey=\left[2\left<u_0^2\right>/
1096: \Om\nu\right]^{1/2},
1097: \quad k=2\upi(\nu/\Om)^{1/2}/L,\quad\tilde{\Delta}=\p_{\tilde{z}}^2+
1098: \p_{\tilde{x}}^2
1099: \end{array}
1100: \right\}
1101: \end{equation}
1102: referred to the Cartesian coordinate system sketched in
1103: Fig. \ref{fig:wavy_geom}. The solid boundary is located at
1104: \begin{equation}
1105: \tilde{\eta}(\tilde{x})=\hat{\eta}\epsilon\uexp(\ui k \tilde{x})+\ucc
1106: \label{profile}
1107: \end{equation}
1108: with constant complex amplitude $\hat{\eta}$ so that $|\hat{\eta}|=1/2$.
1109: The dimensionless (and two dimensional) Navier-Stokes equation
1110: (tildes are omitted in what follows) reads,
1111: \begin{equation}
1112: \partial_t{\Delta\psi}+\Rey\frac{\p(\psi,\Delta\psi)}{\p(z,x)}=
1113: \Delta^2\psi ,
1114: \label{rvwb1}
1115: \end{equation}
1116: with no-slip conditions at the boundary,
1117: \refstepcounter{equation}
1118: $$
1119: \psi=0, \quad \p_z\psi=0 \quad \mbox{at}\quad y=\eta(x),\qquad
1120: \eqno{(\theequation{\mathit{a},\mathit{b}})}
1121: \label{rvwb2}
1122: $$
1123: and the imposed uniform flow at infinity,
1124: \refstepcounter{equation}
1125: $$
1126: \p_x\psi=0, \quad \p_z\psi=2^{-1/2}u_0(t) \quad \mbox{at}\quad z=\infty.
1127: \eqno{(\theequation{\mathit{a},\mathit{b}})}
1128: \label{rvwb3}
1129: $$
1130: These equations depend only on three dimensionless parameters: $\eps$,
1131: the ratio of the amplitude of the wavy wall to the boundary layer
1132: width; $\Rey$, the Reynolds number (the square root of $\Rey_p$ used in
1133: Section \ref{sec:sch}); and $k$, the wavenumber of the wall profile in units
1134: of boundary layer width. We assume $\eps \ll 1$ and expand the
1135: stream function in a power series of $\eps$,
1136: \begin{equation}
1137: \psi=\psi_0+\eps\psi_1+\ldots
1138: \label{rvwb5}
1139: \end{equation}
1140: The boundary conditions are likewise expanded in power series of
1141: $\eps$,
1142: \begin{equation}
1143: \psi(z)|_{z=\eta}=\psi_0(z)|_{z=0}+\eps(\psi_1(z)|_{z=0}+
1144: \eta\p_z\psi_0(z)|_{z=0})+\ldots
1145: \label{rvwb6}
1146: \end{equation}
1147: We now solve (\ref{rvwb1}) order by order in $\eps$.
1148:
1149: At zeroth order the wall is effectively planar. We decompose the Fourier
1150: transform of the stream function as,
1151: $ \hat{\psi}_0(x,z,\om)=2^{-1/2}\hat{u}_0(\om)\hat{\zeta}_0(z,\om)$.
1152: The function $\hat{\zeta}_0(z,\om)$ is given in Eq. (\ref{sch81}). At
1153: this order, the solution is identical to that found for a planar boundary.
1154:
1155: At first order we seek a solution of the form,
1156: \begin{equation}
1157: \psi_1=\hat{\eta} \; \uexp(\ui kx)\phi(z,t)+\ucc
1158: \label{rvwb7}
1159: \end{equation}
1160: so that the amplitude $\phi(z,t)$ satisfies the Orr-Sommerfeld equation,
1161: \begin{equation}
1162: (\p_t{\cal D}-{\cal D}^2)\phi=\ui k\Rey (\p_z^3\psi_0-\p_z\psi_0{\cal D})
1163: \phi,\quad {\cal D}=\p_z^2-k^2
1164: \label{rvwb8}
1165: \end{equation}
1166: with non-homogeneous boundary conditions,
1167: \refstepcounter{equation}
1168: $$
1169: \phi=0, \quad \p_z\phi=-\p_z^2\psi_0 \quad \mbox{at}\quad z=0,\quad
1170: \eqno{(\theequation{\mathit{a},\mathit{b}})}
1171: \label{rvwb9}
1172: $$
1173: \refstepcounter{equation}
1174: $$
1175: \phi=0, \quad \p_z\phi=0 \quad \mbox{at}\quad z=\infty,\quad
1176: \eqno{(\theequation{\mathit{a},\mathit{b}})}
1177: \label{rvwb10}
1178: $$
1179:
1180: The linear operator in the left hand side of Eq. (\ref{rvwb8})
1181: contains a significant difference with respect to that of Eq.
1182: (\ref{sch9}), the equation governing the first order stream function
1183: for the case of a slightly curved boundary. Both equations describe
1184: vorticity diffusion, but the biharmonic equation (\ref{rvwb8}) contains
1185: a cut-off through the parameter $k$. Is is precisely this term that
1186: will lead to an asymptotic exponential decay of the velocity field
1187: sufficiently far away from the boundary for any finite $\beta$. The
1188: exponential decay at long distances arises from the screening
1189: introduced by the simultaneous positive and negative vorticity
1190: produced at the troughs and crests of the wavy wall.
1191:
1192: In order to obtain a solution of the Orr-Sommerfeld equation (Eq.
1193: (\ref{rvwb8})),
1194: we further expand the amplitude $\phi(z,t)$ in power series of
1195: $k\Rey=2\upi\left(2 \left<u_0^2\right>\right)^{1/2}/L\Om$. This is the
1196: ratio between the amplitude of oscillation of the flow at infinity and
1197: the wall wavelength. We write,
1198: \begin{equation}
1199: \phi=\phi_0+\ui k \Rey\phi_1+\ldots
1200: \label{rvwb11}
1201: \end{equation}
1202: The function $\phi_0$ obeys the linearized Eq. (\ref{rvwb8}) with
1203: boundary conditions as in Eqs. (\ref{rvwb9}-\ref{rvwb10}) with
1204: $\phi$ replaced by $\phi_{0}$. The Fourier transform of $\phi_{0}$ is
1205: given by,
1206: \begin{equation}
1207: \hat{\phi}_0(z,\om)=(2)^{-1/2}\hat{u}_0(\om)
1208: \frac{\alpha}{\rho-k}\left(\ue^{-\rho z}-\ue^{-kz}\right),
1209: \label{rvwb_phi01}
1210: \end{equation}
1211: with $\rho\equiv(\alpha^2+k^2)^{1/2}$, and the principal branch of the
1212: square root is assumed $(\Re\{\rho\}>0)$. Recall that
1213: $\alpha = ( 1 + \ui{\rm sign}(\om))(\om/2\nu)^{1/2}$. The field
1214: $\phi_0$ describes
1215: vorticity diffusion near the wavy wall caused by the uniform but
1216: oscillatory far field flow. Both the spatial and ensemble averages of
1217: this flow are zero. However, the flow non-uniformity at this order
1218: induces mean flow at the next order, as it is readily apparent from the
1219: equation for $\phi_1$,
1220: \begin{equation}
1221: (\p_t{\cal D}-{\cal D}^2)\phi_1=(\p_z^3\psi_0-\p_z\psi_0{\cal D})\phi_0,
1222: \label{rvwb_phi1}
1223: \end{equation}
1224: with boundary conditions $\phi_1=0, \; \p_z\phi_1=0$ at $z=0,\infty$.
1225: The field $\phi_1$ describes vorticity diffusion forced by the
1226: nonlinear interaction between $\phi_0$ and $\psi_0$. As was the case
1227: in Sec. \ref{sec:sch}, we focus on the long time limit of the ensemble
1228: average of Eq. (\ref{rvwb_phi1}) $\phi_1^{(s)} = \lim_{t \rightarrow
1229: \infty} \left<\phi_1\right> = \chi + \ucc$, where $\chi$ is given by,
1230: \begin{equation}
1231: {\cal D}^2\chi=-\frac{1}{2}G(z,\beta),
1232: \label{rvwb12}
1233: \end{equation}
1234: with,
1235: \begin{eqnarray}
1236: G(z,\beta)=\int_{0}^{\infty}\ud\om\,P(\om,\beta)\,Q(z,\om,\beta),\nonumber\\
1237: Q(z,\om,\beta)=\alpha(\rho+k)(2\ue^{-(\alpha^*+\rho)z}-
1238: \ue^{-(\alpha^*+k)z}-\ue^{-\rho z}).
1239: \label{rvwb13}
1240: \end{eqnarray}
1241: The corresponding boundary conditions are homogeneous,
1242: $ \chi=0, \; \p_z\chi=0$ at $z=0,\infty$. The solution is,
1243: \begin{eqnarray}
1244: \chi(z,\beta) & = &\int_{0}^{\infty}\ud\om\,P(\om,\beta)\hat{\chi}(z,\om),
1245: \nonumber\\
1246: \hat{\chi}(z,\om) & = &
1247: A_1\ue^{-\rho z}+A_2\ue^{-(\alpha^{*}+\rho)z}+A_3\ue^{-(\alpha^*+k)z}+
1248: (B_1+z B_2)\ue^{-kz}
1249: \label{rvwb_chi}
1250: \end {eqnarray}
1251: where the functions $A_i,\,B_i$ depend on frequency and wavenumber,
1252: \begin{eqnarray}
1253: A_1=D/\alpha^4,\quad A_2=D/(2\alpha^2\rho^2),\quad
1254: A_3=-D/(\alpha^2(\alpha^*+2k)^2,\quad D=\alpha(\rho+k)/2,\qquad\nonumber\\
1255: B_1=-(A_1+A_2+A_3),\quad B_2=(\rho-k)A_1+(\rho+\alpha^*-k)A_2 +
1256: \alpha^*A_3.\qquad
1257: \label{rvwb_chi_const}
1258: \end{eqnarray}
1259: Therefore the stationary part of the averaged first order stream
1260: function is given by,
1261: \begin{equation}
1262: \psi_1^{(s)}=\ui k \Rey\hat{\eta}\uexp(\ui kx)(\chi+\chi^*)+\ucc
1263: \label{rvwb17}
1264: \end{equation}
1265: This solution shows that $\psi_1^{(s)}$
1266: has a phase advance of $\upi/2$ with respect to the wall profile, and hence
1267: the flow in the vicinity of the boundary is directed from trough to crest
1268: ($\p_z^2(\chi+\chi^*)|_{z=0} > 0 )$.
1269: By shifting the coordinate system along the $x$ axis we can change the
1270: phase of the wall profile so as to make it a simple cosine function.
1271: We consider $\eta(x)=\eps\ucos(kx)$ in what follows.
1272:
1273: Following \cite{re:lyne71}, we now proceed to study the limits of
1274: $k$ large and small, while $k\Rey \ll 1$. For $k \gg 1$ the wavelength of the
1275: boundary profile is much smaller that
1276: the thickness of the viscous layer. In this case, a boundary layer
1277: appears of characteristic thickness $1/k$. Screening between regions
1278: producing positive and negative vorticity occurs over a distance
1279: much smaller that the Stokes thickness based on the frequency of
1280: oscillation. The net vorticity does not diffuse even to
1281: distances of order $z\sim\cal {O}$(1), hence
1282: giving rise to exponential decay with an $\uexp(-kz)$
1283: factor. The region in which the stream function is not exponentially
1284: small depends on $Z = kz$. The explicit form of $\psi_1^{(s)}$ may be
1285: obtained by direct expansion of the solution (\ref{rvwb17}) in power
1286: series of $1/k$, keeping $Z$ fixed.
1287: The leading contribution to the steady part of
1288: the tangential component of the velocity is given by,
1289: \begin{equation}
1290: u_1^{(s)}=k\p_Z\psi_1^{(s)}\sim-\frac{\Rey}{24k^2}\usin(kx)\,\ue^{-Z}Z(6-Z^2)
1291: \label{wavy_klarge}
1292: \end{equation}
1293: The boundary layer comprises two recirculating cells per wall
1294: period, located within $0<z\lesssim1/k$. The separation point is given
1295: by $z^2=6/k^2$.
1296:
1297: In the opposite limit of $k \ll 1$ one formally recovers the Schlichting
1298: problem in that the characteristic longitudinal length scale is much
1299: larger than the Stokes layer thickness. There is one fundamental
1300: difference, however, which can be seen from the solution,
1301: Eq. (\ref{rvwb_chi}). It has two contributions: the first one is
1302: proportional $A_i$, arises from the particular solution,and serves to
1303: balance the non-homogeneity in Eq. (\ref{rvwb12}). This contribution
1304: decays within the Stokes layer. The second one is proportional to
1305: $B_i$, and arises from the general solution of the homogeneous part of
1306: the equation. This contribution decays over the stretched scale
1307: $Z$. It turns out that this second
1308: contribution introduces an additional separation of the composite
1309: boundary layer when $0<k\leq0.23$. The lines of zero value of the
1310: longitudinal component of the steady velocity profile
1311: are shown Fig. \ref{fig:wavy_uzero}. The location of the second
1312: separation point is entirely determined by that part of the solution
1313: that is proportional to $B_{i}$, and occurs at $z \sim {\cal{O}} (1/k) \gg 1$.
1314:
1315: The steady velocity may be obtained by expanding the
1316: exact solution (Eq. (\ref{rvwb_chi})) in power series of $k$, first
1317: keeping $z$ fixed (inner solution, $u_{1i}^{(s)}$),
1318: and second keeping $Z$ fixed (outer solution $u_{1o}^{(s)}$).
1319: To leading order, we find,
1320: \begin{eqnarray}
1321: u_{1i}^{(s)}(z') & \sim & -\Rey k^2\usin(kx)
1322: \left\{ \big[\frac{z'}{2}(\usin(z')-\ucos(z'))+2\usin(z')+\frac{1}{2}
1323: \ucos(z')\big]\ue^{-z'}
1324: \right. \nonumber\\
1325: && \left.\mbox{}\hspace{70pt}+
1326: \frac{\ue^{-2z'}}{4} -\frac{3}{4}\astrut\right\},\quad
1327: z' = z/2^{1/2}\sim {\cal{O}}(1),\\
1328: u_{1o}^{(s)}(Z) & \sim & -\Rey k^2\usin(kx)\big[\frac{3}{4}(Z-1)\ue^{-Z}\big],
1329: \qquad Z\sim {\cal{O}}(1) .
1330: \label{wavy_io_sols}
1331: \end{eqnarray}
1332: The solution for the inner and outer steady velocities was already obtained by
1333: \cite{re:lyne71} by a conformal transformation technique. We further note that
1334: the inner and outer solutions can now be matched by requiring that
1335: $u_{1i}^{(s)}(\infty)=u_{1o}^{(s)}(0)=3/4\Rey k^2\usin(kx)$. Hence it
1336: is possible to construct a uniformly valid solution by adding the
1337: inner and outer solutions, and subtracting the first term of the
1338: inner expansion of the outer solution. We find,
1339: \begin{eqnarray}
1340: u_{1c}^{(s)}(z') & \sim & -\Rey k^2\usin(kx)
1341: \left\{ \big[\frac{z'}{2}(\usin(z')-\ucos(z'))+2 \usin(z') +
1342: \frac{1}{2}
1343: \ucos(z')\big]\ue^{-z'} \right. \nonumber\\
1344: && \left.\mbox{} \hspace{70pt}+
1345: \frac{\ue^{-2z'}}{4} + \frac{3}{4}(2^{1/2}kz'-1)\ue^{-2^{1/2}kz'}
1346: \astrut \right\}
1347: \label{wavy_csol}
1348: \end{eqnarray}
1349:
1350: We now turn to a numerical study of the case of finite $\beta$.
1351: The boundary value problem (\ref{rvwb12}) has been solved numerically
1352: by using a multiple shooting method for non stiff and linear boundary
1353: value problems (\cite{re:mattheij84}). The method has the advantage
1354: that the necessary intermediate shooting points are determined by the method itself,
1355: and that it can give the solution on a preset and nonuniform grid of points.
1356: The code was tested on the analytically known solution of the deterministic
1357: limit, Eqs. (\ref{rvwb_chi}) and (\ref{rvwb_chi_const}). Our results are
1358: summarized in Figs. \ref{fig:wavy_uzero}, \ref{fig:wavy_ks} and \ref{fig:wavy_kl}.
1359:
1360: Figure \ref{fig:wavy_uzero} shows the separation points of the stationary
1361: velocity as a function of the boundary wavenumber $k$ for a range of values of
1362: $\beta$, including for reference the deterministic limit of $\beta = \infty$
1363: (separation points are defined to be the zeros of the stationary tangential velocity).
1364: This figure shows that for fixed $k$, the mean stationary velocity field may
1365: be comprised
1366: of two or four recirculating cells per wall period depending on $\beta$.
1367: The first separation point is largely independent of $\beta$, whereas
1368: the deviation of the second relative to its value in the monochromatic
1369: limit is proportional to $\beta/\upi(1+\beta^2)$, the value of $P(0,\beta)$.
1370:
1371: Our results in the limit $k\ll1$ are presented in Fig. \ref{fig:wavy_ks}, where
1372: profiles of tangential component of the mean stationary velocity
1373: are plotted for $k=0.1$ and different values of $\beta$. At large
1374: $\beta$ (close to the monochromatic limit) the flow is comprised of four
1375: recirculating cells per boundary period. Upon decresing $\beta$, the
1376: second separation point moves to infinity (see also Fig.
1377: \ref{fig:wavy_uzero}), so that beyond some critical value of $\beta$,
1378: only two recirculating cells remain. Further decrease in $\beta$ results
1379: in the reappearance of the second separation point at infinity, which
1380: then continues to move towards decreasing $z$. The intensity of the
1381: recirculating modes does not change monotonically with $\beta$ as we
1382: further discuss below. In the opposite limit of $k \gg 1$
1383: (Fig. \ref{fig:wavy_kl} shows the case $k=10.0$; note that
1384: $u_1^{(s)}$ is now normalized by $\Rey/k^2$), the qualitative structure of
1385: the flow is largely independent of $\beta$. The streaming flow has
1386: two recirculating cells per wall wavelength, and their intensity increases
1387: monotonically with decreasing $\beta$.
1388:
1389: The complex dependence of the flow on $\beta$ and $k$ can be
1390: qualitatively understood from the interplay between the width
1391: of the power spectrum (given by $1/\beta$), the viscous damping of each
1392: elementary excitation that depends on its frequency, and the penetration
1393: depth of the flow field which is primarily dictated by the boundary wavelength.
1394: For small $k$, large frequency modes are damped close to the boundary and
1395: do not penetrate much into the recirculating layers. Reducing $\beta$
1396: introduces high frequency components into the driving terms at first
1397: order, but they are dynamically damped.
1398: At the same time, the power in the dominant frequency components (around
1399: $\Omega$) decreases. Overall, a decrease in $\beta$ then leads to a decrease in
1400: recirculation strength. As $k$ increases, larger
1401: frequencies contribute to the flow over the entire range of the
1402: recirculating cells. Decreasing $\beta$ decreases the strength of the
1403: dominant components, but increases the range of high frequencies that
1404: can contribute to the flow. From Eq. (\ref{rvwb12}) one can show that the
1405: driving contribution from higher frequencies which is contained in $Q$
1406: increases faster
1407: with frequency than the decreasing weight given to them by the power
1408: spectrum $P(\omega,\beta)$. Consequently, decreasing $\beta$ (which
1409: amounts to moving towards the white noise limit) leads to increasing
1410: amplitude of the recirculation.
1411:
1412: In summary, for any value of $\beta$, finite or infinite, the vorticity
1413: produced by vibration of the wavy boundary does not penetrate into the bulk
1414: farther than a distance of order of the wavelength of the boundary.
1415: However, there are qualitative differences with the deterministic
1416: limit in the character of the flow within that layer. In particular,
1417: the structure and the intensity of the stationary secondary flow
1418: strongly depend on $\beta$.
1419:
1420: \section{Summary}
1421:
1422: We have addressed the flow induced by a randomly vibrating solid
1423: boundary in an otherwise quiescent fluid. This analysis has been
1424: motivated by the random residual acceleration field in which microgravity
1425: experiments are conducted. The salient features of the flow are
1426: summarized below.
1427:
1428: When the solid boundary is planar, the flow field averages to zero
1429: (the average velocity of the boundary has been taken to be zero in all
1430: cases investigated), but its variance decays algebraically with
1431: distance away from the wall. This dependence follows from a
1432: non vanishing power spectrum of the boundary velocity at zero
1433: frequency. Introducing a low frequency cut-off in the power spectrum
1434: leads back to the classical exponential decay, with a rate that is
1435: determined by the cut-off frequency, Eq. (\ref{plane4_cut}).
1436: The amplitude of the decaying variance depends explicitly on the
1437: correlation time of the boundary velocity, $\beta = \Omega \tau$,
1438: where $\Omega$ is the dominant angular frequency of the power spectrum
1439: of the boundary velocity, and $\tau$ is inverse spectral width ($\tau$
1440: is the correlation time of the boundary velocity).
1441:
1442: If the solid boundary is curved, steady streaming is generated in
1443: analogy with the classical analysis of Schlichting. The stationary part
1444: of the ensemble average of the secondary velocity is nonzero, even though
1445: the boundary velocity averages to zero. In this case, we find that the
1446: leading contribution to the average stationary velocity diverges
1447: logarithmically with distance away from the boundary.
1448: In analogy to the planar case, the introduction of a low frequency cut-off
1449: in the power spectrum of the boundary velocity changes the asymptotic behavior
1450: qualitatively. The average stationary velocity asymptotes now to a constant,
1451: given by Eq. (\ref{sch_cut_asym2}). The asymptotic velocity explicitly
1452: depends on $\beta$ and logarithmically on the cut-off frequency. This asymptotic
1453: behavior is not reached until a length scale of the order of the Stokes layer
1454: thickness that is based on the cut-off frequency.
1455:
1456: We have finally analyzed the case of a periodically modulated solid
1457: boundary in the limit in which the scale of the wall modulation is
1458: small compared to the thickness of the Stokes layer, and also when the
1459: spatial amplitude of the boundary oscillation is small compared with
1460: the wavelength of the wall profile. Cancellation of vorticity
1461: production over the wall boundary leads to exponential decay of the
1462: fluid velocity away from the boundary, with a decay length which is
1463: proportional to the wall wavelength, even if the zero
1464: frequency value of the power spectrum of the boundary velocity is
1465: nonzero. If the boundary wavelength is much larger than the Stokes
1466: layer thickness, we find steady streaming in secondary flow with two or
1467: four recirculating cells per wall period depending on $\beta$.
1468: On the other hand, if the wavelength is
1469: much smaller than the Stokes layer thickness, only two recirculating
1470: cells are formed regardless of the value of $\beta$. Somewhat unexpectedly,
1471: the intensity of the recirculation can both increase or decrease with $\beta$.
1472:
1473: \begin{acknowledgments}
1474: This research has been supported by the Microgravity Science and Applications
1475: Division of the NASA under contract No. NAG3-1885, and also in part
1476: by the Supercomputer Computations Research Institute, which is
1477: partially funded by the U.S. Department of Energy, contract No.
1478: DE-FC05-85ER25000.
1479: \end{acknowledgments}
1480:
1481: \newpage
1482: \bibliographystyle{$HOME/tex/jfm}
1483: \bibliography{dv1}
1484:
1485: \newpage
1486: \begin{figure}
1487: \vspace{1.5pc}
1488: \centerline{\epsffile{plane_var.eps}}
1489: \caption{Normalized variance of the tangential velocity for the case
1490: of a planar boundary computed by numerical integration of Eq. (\ref{plane3})
1491: (symbols), and its uniform asymptotic expansion, Eq. (\ref{plane5}),
1492: (solid lines). The function $I(z,\beta)\,z^2$ asymptotes to a constant
1493: value outside of the classical Sokes layer based on $\Om$. The uniform expansion remains
1494: a good approximation even for moderate $\beta$.}
1495: \label{fig:plane_var}
1496: \end{figure}
1497: %\vskip 5cm
1498: \newpage
1499: \newpage
1500: %
1501: \begin{figure}
1502: \vspace{1.5pc}
1503: \centerline{\epsffile{sc_u_prof.eps}}
1504: \caption{Stationary first order velocity as a function of distance for a
1505: range of values of $\beta$. All curves diverge logarithmically at large $z$,
1506: except for $\beta=\infty$ (monochromatic limit), in which the velocity
1507: asymptotes to a constant within the Stokes layer. This latter behavior
1508: reproduces the classical result of Schlichting.}
1509: \label{fig:sc_u_prof}
1510: \end{figure}
1511: %\vskip 5cm
1512: \newpage
1513: %
1514: \begin{figure}
1515: \vspace{1.5pc}
1516: \centerline{\epsffile{sc_cut_prof.eps}}
1517: \caption{Stationary first order velocity as a function of distance for a
1518: range of values of $\beta$. The power spectrum of the boundary velocity
1519: has a low frequency cut-off at $\om_c=0.05$. The velocity asymptotes to
1520: a constant that depends on the value of $\beta$.}
1521: \label{fig:usc_cut_prof}
1522: \end{figure}
1523: %\vskip 5cm
1524: \newpage
1525: %
1526: \begin{figure}
1527: \vspace{1.5pc}
1528: \centerline{\epsffile{sc_cut.eps}}
1529: \caption{Asymptotic dependence of the stationary velocity as a function of
1530: the cut-off frequency $\om_c$. We show the case $\beta=10$ given by
1531: Eq. (\ref{sch_cut_asym2}) along with the numerically obtained solution.}
1532: \label{fig:usc_cut}
1533: \end{figure}
1534: %\vskip 5cm
1535: \newpage
1536: %
1537: \begin{figure}
1538: \vspace{1.5pc}
1539: \centerline{\epsffile{wavy_geom.eps}}
1540: \caption{Schematic view of the geometry of the wavy wall studied
1541: in Section \ref{sec:wavy_wall}.}
1542: \label{fig:wavy_geom}
1543: \end{figure}
1544: %\vskip 5cm
1545: \newpage
1546: %
1547: \begin{figure}
1548: \vspace{1.5pc}
1549: \centerline{\epsffile{wavy_uzero.eps}}
1550: \caption{Separation points of the boundary layer over the wavy boundary
1551: as a function of its wavenumber $k$, for different values of $\beta$.
1552: The separation points are the loci of zero tangential velocity. At fixed
1553: $k$, the location of the second separation point depends strongly on $\beta$.}
1554: \label{fig:wavy_uzero}
1555: \end{figure}
1556: \newpage
1557: %
1558: \begin{figure}
1559: \vspace{1.5pc}
1560: \centerline{\epsffile{wavy_ks.eps}}
1561: \caption{Tangential component of the mean stationary velocity as a function
1562: of $z$ for $k=0.1$ and a range of values of $\beta$. The case $\beta=\infty$
1563: corresponds to analytic solution obtained by Lyne. The other
1564: curves are the numerical solutions of the boundary value problem
1565: defined by Eq. (\ref{rvwb12}) and corresponding boundary conditions.}
1566: \label{fig:wavy_ks}
1567: \end{figure}
1568: \newpage
1569: %
1570: \begin{figure}
1571: \vspace{1.5pc}
1572: \centerline{\epsffile{wavy_kl.eps}}
1573: \caption{Tangential component of mean stationary velocity as a function
1574: of $z$ for $k=10$ and a range of values of $\beta$. The case $\beta=\infty$
1575: corresponds to analytic solution obtained by Lyne. The other
1576: curves are the numerical solutions of the boundary value problem
1577: defined by Eq. (\ref{rvwb12}) and corresponding boundary conditions.}
1578: \label{fig:wavy_kl}
1579: \end{figure}
1580:
1581: \end{document}
1582:
1583: