physics0607270/STD.tex
1: \documentclass[notitlepage,pre,11pt,showkeys,showpacs,nofootinbib]{revtex4}
2: %\documentclass[notitlepage,aps,12pt,tightenlines]{revtex4}
3: 
4: % This package handles embedded postscript figures.
5: \usepackage{graphicx}
6: \graphicspath{{figs/}}
7: % The amsmath and amssymb packages provide some extra commands/fonts.
8: \usepackage{amsmath,amssymb}
9: \usepackage{bm}
10: \numberwithin{equation}{section}
11: 
12: \bibliographystyle{elsart-num}
13: 
14: %
15: % Definitions
16: %
17: 
18: %\newcommand{\comment}[1]{}
19: \newcommand{\comment}[1]{({\it *** #1 ***})}
20: 
21: \newcommand{\ie}{\textit{i.e.}}
22: \newcommand{\etal}{\textit{et~al.}}
23: \newcommand{\eg}{\textit{e.g.}}
24: 
25: \newcommand{\mathnotation}[2]{\newcommand{#1}{\ensuremath{#2}}}
26: 
27: % gradient
28: \newcommand{\grad}{\nabla}
29: % Laplacian
30: \newcommand{\lapl}{\Delta}
31: \renewcommand{\l}{\left} % \left
32: \renewcommand{\r}{\right} % \right
33: \mathnotation{\ldef}{\mathrel{\raisebox{.069ex}{:}\!\!=}}% Left define
34: \mathnotation{\rdef}{\mathrel{=\!\!\raisebox{.069ex}{:}}}% Right define
35: % Partial derivative
36: \def\pd{\partial}
37: % Real numbers
38: \def\reals{{\mathbb R}}
39: % 1st partial derivative
40: \newcommand {\pdd}[2]{\frac{\partial #1}{\partial #2}}
41: % 2nd partial derivative
42: \newcommand {\pdTwo}[2]{\frac{\partial^{2} #1}{\partial #2^{2}}}
43: % reference style
44: %\newcommand {\nref}[1]{~(\ref{#1})}
45: \newcommand {\nref}[1]{~\ref{#1}}
46: % unit vector notation
47: \newcommand {\unitv}[1]{\bm{\hat{#1}}}
48: % the subscript for the horizontal component
49: \newcommand {\horiz}{_{_{H}}}
50: % the symbol for the velocity
51: \newcommand {\vel}{\bm{u}}
52: % the symbol for the horizontal velocity
53: \newcommand {\uh}{\vel\horiz}
54: % the symbol for the vertical unit vector
55: \newcommand {\khat}{\unitv{k}}
56: % the symbol for the unit vector in th y direction
57: \newcommand {\jhat}{\unitv{j}}
58: % the symbol for the unit vector in the direction \hnabla h
59: \newcommand {\nhat}{\unitv{n}}
60: % the symbol for the horizontal nabla
61: \newcommand {\hnabla}{\nabla\!\horiz}
62: % the symbol for the horizontal laplacian
63: \newcommand {\hlap}{\hnabla^{\,2}}
64: % the subscript denoting top
65: \newcommand {\up}{_{_{T}}}
66: % the subscript denoting bottom
67: \newcommand {\down}{_{_{B}}}
68: % degree symbol
69: \newcommand {\degree} {^{\circ}}
70: % Cross product
71: \def\cross{\times}
72: 
73: \newcommand{\Perm}{\rm{Pe}}
74: \newcommand{\Pe}{\mathrm{Pe}}
75: \newcommand{\Pec}{{{\mathcal{P}}e}}
76: \mathnotation{\ks}{k_s}
77: \mathnotation{\ku}{k_u}
78: \newcommand{\fracs}[2]{#1/#2}
79: 
80: \begin{document}
81: 
82: \title{Stirring up trouble:\\
83: Multi-scale mixing measures for steady scalar sources\\
84: }
85: \author{Tiffany A. Shaw}
86: \email{tshaw@physics.utoronto.ca}
87: \affiliation{Department of Physics, University of Toronto,
88: Toronto, M5S 1A7, Canada}
89: \author{Jean-Luc Thiffeault}
90: \email{jeanluc@imperial.ac.uk}
91: \affiliation{Department of Mathematics, Imperial College
92:   London, SW7 2AZ, United Kingdom}
93: \author{Charles R. Doering}
94: \email{doering@umich.edu}
95: \affiliation{Department of Mathematics and Michigan Center
96: for Theoretical Physics\\
97: University of Michigan, Ann Arbor, MI 48109-1043, USA}
98: 
99: 
100: \begin{abstract}
101: The mixing efficiency of a flow advecting a passive scalar sustained by steady
102: sources and sinks is naturally defined in terms of the suppression of
103: bulk scalar variance in the presence of stirring, relative to the
104: variance in the absence of stirring.  These variances can be weighted at
105: various spatial scales, leading to a family of multi-scale mixing measures and
106: efficiencies.  We derive {\it a priori} estimates on these efficiencies from
107: the advection--diffusion partial differential equation, focusing on a broad
108: class of statistically homogeneous and isotropic incompressible flows.  The
109: analysis produces bounds on the mixing efficiencies in terms of the P\'eclet
110: number, a measure the strength of the stirring relative to molecular
111: diffusion.  We show by example that the estimates are sharp for particular
112: source, sink and flow combinations.  In general the high-P\'eclet number
113: behavior of the bounds (scaling exponents as well as prefactors) depends on
114: the structure and smoothness properties of, and length scales in, the scalar
115: source and sink distribution.  The fundamental model of the stirring of a
116: monochromatic source/sink combination by the random sine flow is investigated
117: in detail via direct numerical simulation and analysis.  The large-scale
118: mixing efficiency follows the upper bound scaling (within a logarithm) at high
119: P\'eclet number but the intermediate and small-scale efficiencies are
120: qualitatively less than optimal.  The P\'eclet number scaling exponents of the
121: efficiencies observed in the simulations are deduced theoretically from the
122: asymptotic solution of an internal layer problem arising in a quasi-static
123: model.
124: \end{abstract}
125: 
126: \keywords{stirring, mixing, advection, diffusion, eddy diffusion, turbulent diffusion}
127: \pacs{47.27.Qb, 92.10.Lq, 92.60.Ek, 94.10.Lf }
128: 
129: \maketitle
130: 
131: \section{Introduction}
132: \label{sec:intro}
133: 
134: Mixing processes in fluids play a key role in a wide variety of engineering
135: applications and for natural systems such as the ocean and atmosphere.  Their
136: theoretical study has been a major focus of research, as indicated by the
137: large number of review articles~\cite{Ottino1990,Majda1999,Warhaft2000,%
138: Shraiman2000,Sawford2001,Falkovich2001,Aref2002,Wiggins2004}.  At the smallest
139: scales mixing is achieved by molecular diffusion processes, but it may be
140: facilitated greatly by stirring.  The result of stirring is usually to enhance
141: the effect of molecular diffusion and increase the mixing
142: rate~\cite{Batchelor1949,Batchelor1952a,Batchelor1959,Kraichnan1968,%
143: Kraichnan1974}.  Quantitative understanding of the fundamental features of
144: stirring and its influence on mixing processes is important for the effective
145: modeling, simulation and design or control of these systems.
146: 
147: The ``efficiency'' of mixing means different things in different contexts.
148: For example the dispersion of an initial distribution by an imposed flow is a
149: transient problem where the temporal approach to the final fully mixed state,
150: rather than the final state itself, is of central interest.  Consider for
151: definiteness the homogeneous advection--diffusion equation for a passive
152: scalar field $\theta(\bm{x},t)$ stirred by a divergence-free velocity field
153: $\bm{u}(\bm{x},t)$,
154: \begin{equation}
155:   \frac{\partial \theta}{\partial t} + \bm{u} \cdot \nabla \theta = \kappa
156:   \lapl \theta
157:   \label{eq:ADhomo}
158: \end{equation}
159: where $\kappa$ is the molecular diffusivity.  If this equation is supplied
160: with initial concentration $\theta(\bm{x},0)$ and applied in an appropriate
161: domain without sources, sinks or scalar flux at the boundaries, then the
162: integral of $\theta$ is conserved, so without loss of generality it may be
163: taken to vanish from the start.  But the $L_{2}$-norm
164: $\|\theta(\cdot,t)\|_{2}$, proportional to the scalar variance in finite
165: volume domains, decreases with time.  Indeed, multiplying (\ref{eq:ADhomo}) by
166: $\theta$ and integrating by parts,
167: \begin{equation}
168:   \frac{d}{dt}  \|\theta\|_{2}^{2} = - 2 \kappa\|\nabla \theta\|_{2}^{2},
169:   \label{eq:ADhomodecay}
170: \end{equation}
171: indicating an inexorable decay of the variance.  Efficient mixing in this
172: transient decay problem means faster decay of the scalar variance.  The mixing
173: efficiency of a particular flow could be defined, for example, in terms of its
174: ability to reduce the variance from the initial value to a prescribed value
175: within a specific period of time~\cite{Constantin2005}.  Because the
176: right-hand side of (\ref{eq:ADhomodecay}) is proportional to $\kappa$, it is
177: evident that molecular processes are ultimately responsible for mixing by this
178: criterion.  Even though the stirring field does not appear explicitly in
179: (\ref{eq:ADhomodecay}), the conventional intuition is that material line
180: stretching in the flow can amplify scalar gradients thereby enhancing the
181: molecular mixing rate.  Indeed, the velocity's rate-of-strain matrix serves as
182: the local growth rate of the scalar gradient field. These issues are of extreme interest for both theory and
183: applications, but in this paper we are interested in a distinct scenario where
184: different effects are at work.  
185: 
186: Mixing a scalar field whose fluctuations are constantly replenished by steady
187: but spatially inhomogeneous sources and sinks is a problem with a long
188: history.  Early on, Townsend~\cite{Townsend1951,Townsend1954} was concerned
189: with the effect of turbulence and molecular diffusion on a line-source of
190: temperature, a heated filament.  The spatial localization of the source,
191: imposed by experimental constraints, enhanced the role of molecular
192: diffusivity.  Saffman~\cite{Saffman1960} also found that molecular diffusion
193: and turbulent diffusion were not simply additive and that higher-order
194: corrections were needed.  Durbin~\cite{Durbin1980} and
195: Drummond~\cite{Drummond1982} introduced stochastic particle models to
196: turbulence modeling, and these allowed more detailed studies of the effect of
197: the source on diffusion.  Sawford and Hunt~\cite{Sawford1986} pointed out that
198: small sources, such as heated filaments, lead to an explicit dependence of the
199: variance on molecular diffusivity.  Many refinements to these models followed,
200: see for instance~\cite{Thomson1990,Borgas1994} and the review by
201: Sawford~\cite{Sawford2001}.  Chertkov
202: \etal~\cite{Chertkov1995,Chertkov1995b,Chertkov1997,Chertkov1997b,%
203: Chertkov1998} and Balkovsky \& Fouxon~\cite{Balkovsky1999} treated the case of
204: a random, statistically-steady source.  Our goal in the present paper is to
205: make the source-dependence of the concentration variance more precise by
206: working directly from the advection--diffusion equation, without specifying
207: the underlying turbulent statistics other than basic stationarity and
208: homogeneity assumptions.
209: 
210: %\section{Advection--Diffusion with a Source}
211: %\label{eq:ADS}
212: 
213: When a source of scalar concentration is present, the transient kinetics are
214: of less immediate interest and instead the properties of the (statistical)
215: steady state are of greater relevance.  As will be seen, this sustained
216: steady-state dynamics highlights other features of stirring and mixing
217: processes.  In comparing the steady-state problem to the transient problem
218: defined by Eq.~\eqref{eq:ADhomodecay}, it is important to remember that the
219: long-time asymptotic behavior of the decaying problem is usually irrelevant
220: to the corresponding long-time behavior of the steady-state problem.  This is
221: because the continuous replenishing of concentration overwhelms
222: small-amplitude effects observed for long-time decay, such as the `strange
223: eigenmode'~\cite{%
224: Pierrehumbert1994,Antonsen1996,Rothstein1999,Fereday2002,Sukhatme2002,%
225: Wonhas2002,Pikovsky2003,Thiffeault2003d,Thiffeault2004b,Schekochihin2004,%
226: Vanneste2005,Gilbert2006}.
227: 
228: In this paper we consider the stirring and mixing of a passive scalar
229: sustained by a steady source-sink function~$s(\bm{x})$.  Given a prescribed
230: divergence-free velocity field $\bm{u}(\bm{x},t)$ and a molecular
231: diffusivity~$\kappa$, the scalar concentration $\theta(\bm{x},t)$ obeys the
232: inhomogeneous advection--diffusion equation
233: \begin{equation}
234:   \frac{\partial \theta}{\partial t} + \bm{u} \cdot \nabla \theta =
235:   \kappa \lapl \theta + s(\bm{x})
236:   \label{eq:AD}
237: \end{equation}
238: supplemented with initial concentration field $\theta(\bm{x},0)$.  We consider
239: a domain without any net scalar flux at the boundaries: the
240: periodic box of size $L$, i.e., $\bm{x} \in \mathbb{T}^d$, the $d$-dimensional
241: torus of volume $L^{d}$.  The spatial mean of $\theta$ is computed
242: immediately,
243: \begin{equation}
244:    \frac{1}{L^{d}} \int \theta(\bm{x},t) \, d^{d}x \ = \
245:    \frac{1}{L^{d}} \int \theta(\bm{x},0) \, d^{d}x \ + \
246:    t \times \frac{1}{L^{d}} \int s(\bm{x}) \, d^{d}x \, ,
247: \end{equation}
248: and deviations from the spatial mean satisfy (\ref{eq:AD}) with
249: $s(\bm{x})$ replaced by $s(\bm{x}) - L^{-d} \int s \, d^{d}y$.  So to study
250: the fluctuations we may assume without loss of generality that
251: $\theta(\bm{x},0)$ and $s(\bm{x})$, and thus also
252: $\theta(\bm{x},t)$ have spatial mean zero.
253: 
254: Fluctuations in the scalar concentration are natually measured in terms of the
255: steady-state variance $\langle \theta^2 \rangle$, where we introduce the
256: space-time average.  The two averaging operations we use are the time average
257: \begin{equation}
258:   \overline{F}(\bm{x})\ldef\lim_{t \rightarrow \infty}
259:   \frac{1}{t} \int_0^t F(\bm{x},t') \, dt',
260: \end{equation}
261: assuming as necessary that the limit exists, and the space-time average
262: \begin{equation}
263:   \langle F \rangle \ldef \frac{1}{L^{d}} \int \overline{F}(\bm{x})
264:   \, d^{d}x \, .
265: \end{equation}
266: Effective stirring makes the scalar field more spatially uniform, lowering the
267: variance, and this is the basic mixing effect that we set out to study.  Many
268: investigations have been concerned with other statistical properties of the
269: scalar field for this kind of model, such as details of the tails of the
270: probability distribution of $\theta$~\cite{Majda1999, Falkovich2001}.  While
271: these studies present fascinating mathematical and physical issues, in terms
272: of applications they are most likely to be of ultimate use in designing
273: closure approximations, i.e., models of the model, in order to accurately
274: estimate bulk measures of mixing like variance reduction.  In this work we
275: focus directly on the supression of the scalar fluctuations as a primary
276: indicator of mixing.
277: 
278: In terms of the Fourier decomposition of the scalar field,
279: \begin{equation}
280:   \hat{\theta}_{\bm{k}}(t) = \frac{1}{L^d}
281:   \int \theta({\bm{x}},t)\,e^{-i\bm{k}\cdot \bm{x}}\,
282:   d^{d}x
283: \end{equation}
284: where ${\bm{k}}=(2\pi/L){\bm{n}}$ for ${\bm{n}}=(n_1,\ldots,n_d)
285: \in \mathbb{Z}^d$, the steady-state variance is
286: \begin{equation}
287:    \langle \theta^2 \rangle = \sum_{\bm{k}}
288:    \overline{|\hat{\theta}_{\bm{k}}|^{2}}.
289: \end{equation}
290: The magnitude $k$ of the wavenumbers naturally index spatial scales, so
291: information about the fluctuations at different scales may be obtained
292: by weighting the sums of the Fourier coefficients~\cite{Mathew2005}.  The
293: simplest indicators of the fluctuations on small and large scales are,
294: respectively,
295: \begin{equation}
296:    \langle |\nabla \theta|^2 \rangle = \sum_{\bm{k}} k^{2}
297:    \overline{|\hat{\theta}_{\bm{k}}|^{2}}
298: \end{equation}
299: and 
300: \begin{equation}
301:    \langle |\nabla^{-1} \theta|^2 \rangle = \sum_{\bm{k}} 
302:    \overline{|\hat{\theta}_{\bm{k}}|^{2}}/k^{2}
303: \end{equation}
304: where the inverse gradient $\nabla^{-1}$ is defined in Fourier space as
305: multiplication by $- i {\bm k}/k^{2}$, a well-defined operator on these
306: functions with spatial mean zero.  For the purposes of this study the
307: effectiveness of stirring on mixing at relatively small and large scales
308: will be gauged in terms of these norms.  Efficient stirring decreases the
309: variances on all scales, although it can be expected that any particular
310: stirring may be more effective on some scales than on others.
311: 
312: It is appropriate here to point out a fundamental and elementary distinction
313: between transient and steady-state stirring.  For efficient transient mixing
314: the goal is to decrease the scalar variance $\sim \|\theta(\cdot,t)\|_{2}^{2}$
315: as quickly as possible by {\it increasing} the gradient variance $\sim
316: \|\nabla \theta(\cdot,t)\|_{2}^{2}$ via stirring.  However in the steady-state
317: problem stirring can only {\it reduce} the mean scalar gradient variance---and
318: thus the mean rate of variance decay---from its purely diffusive value in the
319: absence of stirring.
320: 
321: The proof of this (perhaps unexpected) fact is easy.
322: Multiplying (\ref{eq:AD}) by $\theta$ and averaging over space and time
323: with appropriate integrations by parts produces the well-known balance
324: \begin{equation}
325:    \kappa \langle |\nabla \theta|^2 \rangle = 
326:    \langle \theta s \rangle.
327:    \label{elphaba}
328: \end{equation}
329: Inserting the gradient and its inverse on the right-hand side, integrating by
330: parts again, and then employing the Cauchy--Schwarz inequality yields
331: \begin{equation}
332:    \kappa \langle |\nabla \theta|^2 \rangle \ = \
333:    \langle \theta  \nabla \cdot \nabla^{-1}s \rangle \ = \
334:    - \langle \nabla \theta \cdot \nabla^{-1}s \rangle \ \le \
335:    \langle |\nabla \theta|^{2}\rangle^{1/2}
336:    \langle |\nabla^{-1}s| \rangle^{1/2}.
337:    \label{glinda}
338: \end{equation}
339: 
340: Note that the long-time solution of (\ref{eq:AD}) in the absence of stirring
341: is the steady-state solution $\theta_{0}(\bm{x})$ of the diffusion equation
342: with source $s(\bm{x})$,
343: \begin{equation}
344:   \theta_{0} \ = \ - \frac{1}{\kappa} \lapl^{-1} s \, ,
345:   \label{zero}
346: \end{equation}
347: where the inverse Laplacian is multiplication by $-k^{-2}$ in Fourier space.
348: Solving for the steady-state scalar gradient variance in (\ref{glinda}) and
349: noting that $\nabla \theta_{0} = -\kappa^{-1} \nabla^{-1} s$, we conclude that
350: \begin{equation}
351:   \langle |\nabla \theta|^2 \rangle \le 
352:   \langle |\nabla \theta_{0}|^2 \rangle.
353:   \label{boc}
354: \end{equation}
355: 
356: This relationship is uniform in the advecting velocity field
357: $\bm{u}(\bm{x},t)$ implying that no clever stirring on any scales can increase
358: the mean scalar gradient variance over its baseline unstirred value.  In light
359: of this observation we may anticipate that stirring strategies designed to
360: maximize mixing efficiency in the sustained source problem are likely to be
361: different from those employed in the transient decay scenario.
362: 
363: The variances $\langle |\nabla^{p} \theta_{0}|^2 \rangle$, defined above for
364: $p \in \{-1,0,1\}$, are both dimensional and dimensionally distinct quantities.
365: In order to compare them with each other or to compare different
366: physical systems we need sensible nondimensional measures.
367: Hence we define the dimensionless {\it multi-scale mixing efficiencies}, denoted
368: ${\cal E}_{p}$ for $p \in \{-1,0,1\}$, via
369: \begin{equation}
370:   {\cal E}_{p}^{2} = \frac {\langle |\nabla^{p} \theta_{0}|^2 \rangle}
371:   {\langle |\nabla^{p} \theta|^2 \rangle}
372: \end{equation}
373: where $\theta_{0}$ is the steady solution of the unstirred problem defined in
374: (\ref{zero}).  Effective stirring decreases scalar variances relative to those
375: due to diffusion alone, increasing these mixing efficiences.  The calculation
376: resulting in (\ref{boc}) has established that ${\cal E}_{1} \ge 1$.
377: 
378: Intensifying the stirring often increases the mixing efficiencies and it is
379: important to characterize this property in terms of the forcefulness of the
380: flow.  The simplest bulk measure of the vigor of the velocity field is its
381: mean kinetic energy, or equivalently the {\it rms} speed $U$ defined by
382: \begin{equation}
383:   U^{2} = \langle \lvert\bm{u}\rvert^2 \rangle.
384: \end{equation}
385: The nondimensional measure of the strength of the stirring relative to the
386: effect of molecular viscosity is the P\'eclet number $\Pe$ that we define as
387: \begin{equation}
388:   \Pe = \frac{UL}{\kappa}
389: \end{equation}
390: using the domain length scale $L$ for simplicity here.  It will become
391: apparent that this domain length scale may not be the most appropriate one for
392: this purpose; determining the relevant length scales is one of the central
393: points of this study.
394: 
395: For many applications it is useful to know how fluctuations at various length
396: scales may be suppressed as functions of $\kappa$, $U$, and other features of
397: the problem such as details of the flow and source-sink structures.  Toward
398: this end it is desirable to know how the multi-scale mixing efficiencies depend
399: on the P\'eclet number, and the notion of ``eddy diffusivity'' provides a
400: conceptual benchmark for this dependence.
401: 
402: A flow with velocity scale $U$ and ``persistence length'', ``mixing length''
403: or ``eddy size" $\ell$ that characterizes the typical distance a particle
404: travels before changing direction can disperse particles diffusively on
405: appropriate space and/or time scales.  This suggests that when advection
406: dominates molecular diffusion, an effective diffusion with coefficient
407: $\kappa_{\mathrm{eff}} \sim U\ell$ might replace the advection to determine
408: some gross statistical features of the scalar field.\footnote{Indeed, the
409: situation where $\ell$ is much smaller than any length scales in the initial
410: data $\theta(\bm{x},0)$ or the source $s(\bm{x})$ is the setting for
411: homogenization theory.}  If this is so, then the steady-state scalar variances
412: are all $\sim \kappa_{\mathrm{eff}}^{-2}$ and the efficiencies are all $\sim
413: \kappa_{\mathrm{eff}}/\kappa$.  According to this argument
414: \begin{equation}
415:    {\cal E}_{p} \sim \frac{U\ell}{\kappa} = \frac{\ell}{L} \times \Pe.
416: \end{equation}
417: 
418: The linear scaling ${\cal E}_{p} \sim \Pe$ at high P\'eclet numbers, which we
419: will refer to as the ``classical'' scaling, provides a baseline reference for
420: the multi-scale mixing efficiencies.  Flows that generate this classical
421: scaling asymptotically as $\Pe \rightarrow \infty$ produce a truly effective
422: ``residual" molecular-like diffusion as far as the suppresion of variances at
423: the various spatial scales is concerned, even in the singular $\kappa
424: \rightarrow 0$ limit of vanishing molecular diffusivity.  And as this
425: discussion suggests, if the efficiency scales classically then the prefactor
426: provides a precise prediction of the length scale with which an eddy
427: diffusivity might be meaningfully identified.
428: 
429: The principal purpose of this paper is to determine limits on the mixing
430: efficiencies, i.e., to derive {\it a priori} bounds on ${\cal E}_{p}$ as a
431: function of $\Pe$, and to investigate what sort of flows might realize those
432: limits.  Upper bounds on ${\cal E}_{p}(\Pe)$ are of particular interest
433: because they characterize the most efficient stirring strategies that can
434: possibly be hoped for.  If the bounds are to be useful then they should be
435: realizable or approachable, or at the very least indicative of the kind of
436: behavior that is possible---such as a scaling like ${\cal E}_{p} \sim
437: \Pe^{\alpha}$ at high P\'eclet numbers.
438: 
439: Upper bounds on the mixing efficiencies follow from lower bounds on the
440: variances and the first study in this direction was apparently by Thiffeault,
441: Doering, and Gibbon \cite{Thiffeault2004} who focused on estimates and
442: simulations for ${\cal E}_{0}$.  They adapted an approach that had been used
443: to bound turbulent dissipation in the Navier--Stokes equations
444: \cite{Doering2002,Doering2003} for application to inhomogeneous
445: advection--diffusion equations.  For the steady source model of interest here
446: they showed that if $\langle \theta_{0}^{2} \rangle < \infty$ then ${\cal
447: E}_{0} \le a + b \Pe$ where the coefficients $a$ and $b$ are homogeneous scale
448: invariant functionals of the source $s$, i.e., invariant under the
449: transformation $s({\bm x}) \rightarrow c s({c' \bm x})$ for any constants $c$,
450: and $c' \ne 0$.
451: 
452: That result showed that very generally the classical scaling is an upper limit
453: to the mixing efficiency in this most basic sense.  Moreover, the coefficient
454: in the high P\'eclet number scaling ${\cal E}_{0} \lesssim b \, \Pe$ puts a
455: limit on any reasonable value for a mixing length: $\ell \lesssim b L$.  It is
456: especially notable that this rigorous estimate of $\ell$ (really of the
457: prefactor $b$) is uniform in the stirring field $\bm{u}$ and independent of
458: any length scales it exhibits.  It also is independent of $\kappa$.  It
459: emerges as a length scale in the source-sink distribution, which is seen to
460: play a more important role in the mixing process than the conventional eddy
461: diffusion picture anticipates.  When the relevant length scale in the
462: sustaining source and sink is small then the variance suppression by advection
463: is necessarily limited by this no matter what spectrum of scales is present in
464: the stirring process.
465: 
466: These observations serve as the starting point for this study.  Here we carry
467: forward the investigation of stirring and variance suppression by extending
468: the analysis to multi-scale mixing measures while focusing on a broad but
469: specific class of statistically stationary homogeneous and isotropic flows.
470: The restriction to this class of flows---a class that includes but is not
471: limited to homogeneous isotropic high Reynolds number turbulence---allows
472: for the exact solution of some variational problems for bounds on the mixing
473: efficiencies.  For certain sources and sinks these bounds yield anomalous
474: sub-classical exponents for the $\Pe$ scaling of some of the ${\cal E}_{p}$.
475: Thus anomalous scaling is inevitable for some source-sink distributions.
476: In those cases it cannot be avoided by manipulating details of the flow or
477: the spectrum of the stirring.
478: We study the case of a monochromatic source 
479: $s({\bm x}) \sim \sin({\bm k}_{s} \cdot {\bm x})$ in detail.  
480: The estimates are particularly simple in this case and they are
481: sharp: we will exhibit a statistically stationary homogeneous and
482: isotropic stirring strategy that saturates the upper bounds.
483: 
484: The rest of this paper is organized as follows.  We introduce the relevant
485: class of statistically stationary, homogeneous and isotropic flows in
486: Section~\ref{sec:SHIF}, and present some specific examples.  We formulate
487: variational problems for bounds on the mixing efficiencies in
488: Section~\ref{sec:BOUNDS}, and derive general estimates.  In
489: Section~\ref{sec:saturation} we evaluate the bounds explicitly for some
490: particular sources and sinks and show that classical scaling estimates may
491: be sharp.  In Section~\ref{sec:SARSAS} we focus on the high-$\Pe$ behavior of
492: the mixing efficiency bounds for a variety of source-sink distributions,
493: and show that anomalous sub-classical scaling is sometimes unavoidable.
494: 
495: In Section~\ref{sec:SSSSSF} we are concerned with the fundamental example of a
496: monochromatic source stirred by a flow with a single length scale, the
497: so-called random sine flow, a type of renewing flow.  Measuring the mixing
498: efficiencies in direct numerical simulations, we find that the large-scale
499: efficiency ${\cal E}_{-1}$ scales (nearly) classically with respect to $\Pe$,
500: like its upper bound, but the intermediate and small-scale efficiencies ${\cal
501: E}_{0}$ and ${\cal E}_{1}$ scale sub-classically, i.e., with powers of $\Pe$
502: less than 1.  We show that the anomalous exponents for ${\cal E}_{0}$ and
503: ${\cal E}_{1}$ can be deduced from the asymptotic analysis of a static flow
504: problem.  The concluding Section~\ref{sec:SandD} contains a summary of the
505: results along with a discussion of open problems and compelling future
506: challenges.  Some technical details are relegated to appendices.
507: 
508: 
509: \section{Statistically Stationary Homogeneous Isotropic Flows}
510: \label{sec:SHIF}
511: 
512: Our approach to estimating mixing efficiencies is kinematic: the stirring
513: vector field ${\bm u}({\bm x},t)$ is assumed to be given.  It could be a
514: solution of the Navier--Stokes equations, or it could be a stochastic process
515: with convenient or interesting spectral properties, or it could be a regular
516: time-periodic field.  The mixing efficiency bounds obtained in this paper will
517: apply so long as a few generic statistical conditions are satisfied.  Of
518: course not every stirring field satisfying these conditions will saturate the
519: bounds, but they are all limited by them.
520: 
521: This analysis in this paper is concerned with velocity fields that satisfy the
522: following three conditions:
523: \begin{itemize}
524: \item{The field is divergence free, $\nabla \cdot {\bm u} = 0$,
525: everywhere and at all times.
526: %We want to consider incompressible volume-preserving flows.
527: }
528: \item{The field has finite mean kinetic energy, $U^{2} = \langle |{\bm u}|^{2}
529: \rangle < \infty$, so that the P\'eclet number $\Pe = UL/\kappa$ is finite.
530: We will also presume more regularity for the velocity whenever necessary to
531: carry out formal calculations.  This will be apparent in the course of the
532: applications if the ultimate estimates depend on other norms of the field.}
533: \item{The velocity field is statistically stationary, homogeneous and
534: isotropic.  For the purposes of this work these qualities are defined by the
535: one- and two-point equal time statistics (presuming that these time averages
536: exist pointwise in space)
537: \begin{equation}
538:  \overline{u_i(\bm{x},\cdot)} = 0
539:  \label{mean1}
540: \end{equation}
541: \begin{equation}
542:   \overline{u_i(\bm{x},\cdot)u_j(\bm{y},\cdot)} \ = \ C_{ij}(\bm{x}-\bm{y})
543:   \ = \ \sum_{\bm{k} \ne 0} \frac{\hat{C}(k)}{d-1}
544:   \ \l(\delta_{ij} - \frac{k_{i}k_{j}}{k^{2}}\r)
545:   \ e^{i\bm{k}\cdot(\bm{x}-\bm{y})} \ + \ \frac{\hat{C}(0)}{d}\delta_{ij}
546:   \label{eq:SHIT}
547: \end{equation}
548: where $\hat{C}(k)$ depends only on the magnitude $k$ of the wavenumber
549: $\bm{k}$.  }
550: \end{itemize}
551: The conditions of finite energy and incompressiblity are familiar and
552: straightforward.  In the remainder of this section we discuss these
553: stationarity, homogeneity and isotropy conditions and their relevant
554: implications for the calculation of bounds on the various multi-scale mixing
555: efficiencies.  We also provide explicit examples, i.e., we describe several
556: flows with these properties that will be useful for considerations
557: in subsequent sections.
558: 
559: First, setting $\bm{x} = \bm{y}$ in (\ref{eq:SHIT}) produces the single point
560: component-by-component covariance
561: \begin{equation}
562:   \overline{u_i(\bm{x},\cdot)u_j(\bm{x},\cdot)} \ = \ C_{ij}(0)
563:   \ = \ \sum_{\bm{k} \ne 0} \frac{\hat{C}(k)}{d-1}
564:   \ \l(\delta_{ij} - \frac{k_{i}k_{j}}{k^{2}}\r)
565:   \ + \ \frac{\hat{C}(0)}{d}\delta_{ij}.
566: \end{equation}
567: Note that because $\hat{C}(k)$ depends only on the magnitude of the wavenumber,
568: \begin{equation}
569: \sum_{\bm{k \ne 0}} \hat{C}(k) \ \frac{k_{i}k_{j}}{k^{2}}
570:  \ = \ \frac{1}{d} \sum_{\bm{k} \ne 0} \hat{C}(k) \delta_{ij}.
571: \end{equation}
572: Thus
573: \begin{equation}
574:   \overline{u_i(\bm{x},\cdot)u_j(\bm{x},\cdot)} \ = \ 
575:   \frac{1}{d} \, \sum_{\text{all} \ \bm{k}} \hat{C}(k) \ \delta_{ij} \ = \
576:   \frac{U^2}{d} \delta_{ij}
577:   \label{eq:HIT1}
578: \end{equation}
579: where we identify the mean square velocity $U^{2} = \langle |{\bm u}|^{2}
580: \rangle = \sum \hat{C}(k)$.
581: 
582: Then if the velocity field is sufficiently regular (or equivalently if
583: $\hat{C}(k)$ decays sufficiently fast as $k \rightarrow \infty$) we may also
584: deduce some correlations of the derivatives of ${\bm u}$.  For example
585: differentiating $C_{ij}(\bm{x}-\bm{y})$ in (\ref{eq:SHIT}) by $x_{k}$ and
586: setting $\bm{y}=\bm{x}$ leads to
587: \begin{equation}
588:  \overline{\frac{\partial u_i(\bm{x},\cdot)}
589:  {\partial x_k}u_i(\bm{x},\cdot)} \ = \ 
590:  \sum_{\bm{k \ne 0}} \frac{i \,\hat{C}(k)}{d-1}
591:   \ \l(\delta_{ij} k_{k} \,  - \frac{k_{i}k_{j} k_{k}}{k^{2}}\r)
592:  \ = \ 0
593:  \label{eq:HIT2}
594: \end{equation}
595: because $\hat{C}(k)$ summed against an odd number of orthogonal components
596: vanishes when the sum is absolutely convergent.  Differentiating
597: $C_{ij}(\bm{x}-\bm{y})$ by $x_{k}$ and $y_{l}$, setting $\bm{y}=\bm{x}$, and
598: subsequently contracting over $k$ and $l$ gives
599: \begin{equation}
600:    \overline{ \frac{u_i(\bm{x},\cdot)}{\partial x_k}\frac{\partial u_j(\bm{x},\cdot)}
601:    {\partial x_k}} \ = \
602:    \frac{1}{d} \, \sum_{\bm{k} \ne 0} k^{2}\, \hat{C}(k) \, \delta_{ij}
603:    \ = \
604:    \frac{\Omega^2}{d} \delta_{ij}
605:    \label{eq:HIT3}
606: \end{equation}
607: where we identify the enstrophy 
608: $\Omega^{2} \ldef \langle |\nabla {\bm u}|^{2} \rangle = \sum k^{2}\hat{C}(k)$.
609: In the context of statistical turbulence theory the ratio
610: $\lambda = U/\Omega$ is (proportional to) the Taylor microscale. 
611: 
612: An example of such a statistically stationary homogeneous isotropic flow is
613: the solution ${\bm u}({\bm x},t)$ of the incompressible Navier-Stokes
614: equations
615: \begin{equation}
616:    \frac{\partial \bm u}{\partial t} + {\bm u} \cdot \nabla {\bm u} + \nabla p
617:    \ = \ \nu \Delta {\bm u} + {\bm f}({\bm x},t)\, , \quad 
618:    \nabla \cdot {\bm u} \ = \ 0,
619:      \label{eq:NS}
620: \end{equation}
621: where $\nu$ is the kinematic viscosity and ${\bm f}({\bm x},t)$ is a spatially
622: periodic body force applied to maintain a statistical steady state.  Of course
623: the forcing would have to be capable of producing the homogeneous and isotropic
624: statistics described above, not an altogether trivial task.  
625: This approach would be taken to study mixing properties of high-Reynolds 
626: number turbulent flows.
627: 
628: Other flows may be easier and more convenient to implement in simulations or
629: to utilize in the analysis of specific models.  Some features of the
630: stirring---for instance a stationary energy spectrum consistent with developed
631: turbulence---can be realized by specifying the modal amplitudes $\hat{{\bm
632: u}}_{\bm{k}}(t)$ as mean-zero stochastic processes with $\bm{k} \cdot
633: \hat{{\bm u}}_{\bm{k}}(t) = 0$ and appropriately uncorrelated for distinct
634: wavenumbers, to produce any desired energy spectrum $\hat{C}(k) =
635: \overline{|\hat{{\bm u}}_{\bm{k}}|^{2}}$.  This is possible
636: when $\overline{|\hat{{\bm u}}_{\bm{k}}|^{2}}$ depends only on the magnitude
637: $k$ of the wavenumber $\bm{k}$. 
638: 
639: One model of interest is a flow involving just a single wavenumber $k \ne 0$,
640: the random sine flow, a.k.a.\ the renewing wave
641: flow~\cite{Pierrehumbert1994,Antonsen1996,Thiffeault2004,Tsang2005}.  (The
642: term `renewing' or `renovating' has been used to refer to flows that are
643: piecewise-constant in time but change randomly at regular
644: intervals~\cite{Dittrich1984,Zeldovich1984,Gilbert1992}.
645: Zeldovich~\cite{Zeldovich1982} introduced a similar single-mode flow, but it
646: was oscillatory rather than renewing and thus had poor mixing properties.)  In
647: this example the velocity vector field switches periodically among steady
648: shearing flows of the form
649: \begin{equation}
650:   \bm{u}(\bm{x}) \ = \ \sqrt{2} \, \bm{U} \,
651:   \sin{(\bm{k}\cdot\bm{x} + \varphi)}
652:    \label{eq:ZSF}
653: \end{equation}
654: where $\bm{k}\cdot\bm{U}=0$, $|\bm{U}| = U$ and the phase $\varphi$ is
655: selected independently and uniformly from $[0,2\pi)$ upon each switch.
656: The switching may be strictly periodic or random in time, and
657: in either case the characteristic persistence $\tau$
658: is another parameter of this flow.
659: Of course an appropriate selection of the directions for $\bm{k}$ and
660: $\bm{U}$ must be made in order to realize statistical homogeneity
661: and isotropy by the definition used here.
662: 
663: The kinetically simplest possible example of a statistically homogeneous and
664: isotropic flow is one where only $\hat{C}(0) \ne 0$.  At each instant of time
665: the flow is then spatially uniform, a steady wind $\bm{u}(\bm{x}) = \bm{U}$
666: with the direction switching periodically or randomly in time so that the time
667: average of $\bm{u}$ vanishes and the component-component correlation satisfies
668: $\overline{u_{i}u_{j}} \sim \delta_{ij}$.  This wind could sample many
669: directions, or as few as $2 \times d$ in the $\pm$ directions along an
670: orthogonal set of coordinate axes.
671: 
672: 
673: %===================================================
674: 
675: \section{Bounds on the Mixing Efficiencies}
676: \label{sec:BOUNDS}
677: 
678: In this section we derive bounds on the multi-scale mixing efficiencies ${\cal
679: E}_{p}$ for~$p=0,1,-1$, corresponding to intermediate, small, and large
680: scales, respectively, for flows that satisfy the statistical homogeneity and
681: isotropy conditions in \eqref{eq:HIT1} and \eqref{eq:HIT2}.  We first briefly
682: address lower bounds, but focus for the most part on upper estimates.  The
683: variational formulation and solution for upper bounds on intermediate and
684: large length scales proceeds along similar lines, so we shall treat the
685: variance in detail and give a more brisk derivations for the large-scale
686: mixing measure.  Two different lower estimates on the small-scale variance,
687: corresponding to upper bounds on the mixing efficiency at small scales, are
688: derived.  One of these small-scale results depends on the spectral
689: distribution of energy in the stirring field while the other depends only on
690: the total bulk energy.
691: 
692: \subsection{Lower bounds on the mixing efficiencies}
693: \label{sec:bound_{lower}}
694: 
695: Lower bounds on the mixing efficiencies follow from upper bounds on the
696: corresponding variances.  We already derived a lower bound for 
697: the small-scale mixing efficiency ${\cal E}_{1}$ in the introduction by
698: considering the
699: steady-state variance dissipation-production balance in (\ref{elphaba}),
700: \begin{equation}
701: \kappa \langle |\nabla \theta|^{2} \rangle = 
702: \langle s \theta \rangle.
703: \label{elphaba2}
704: \end{equation}
705: The subsequent result in (\ref{boc}), valid for any incompressible flow even
706: without invoking any statistical assumptions, is precisely the statement that
707: ${\cal E}_{1} \ge 1$.
708: 
709: We expect that this estimate can be sharp in the sense that there may exist a
710: flow field with arbitrary P\'eclet number that may realize it.  Certainly this
711: is true in $2d$: given the source-sink function $s(\bm{x})$ with unstirred
712: scalar distribution $\theta_{0}(\bm{x})$, just define a flow field with stream
713: function $\psi(\bm{x}) \sim \theta_{0}(\bm{x})$.  The streamlines of such a
714: flow are along level sets of $\theta_{0}$ so the flow has no effect.  Indeed,
715: for this flow $\bm{u}\cdot \nabla \theta_{0} = 0$ no matter what the magnitude
716: of $U$ is, so $\theta_{0}$ is the stationary solution of the
717: advection--diffusion equation for any value of $\Pe$.  However, this perfectly
718: ``non-mixing'' flow is not statistically stationary, homogeneous and
719: isotropic, and it is not clear whether further constraints derived from the
720: full advection--diffusion equation might be implemented to raise this lower
721: estimate for such fluctuating flows.  We leave that question for a future
722: study.
723: 
724: We can follow the same line of reasoning to derive lower estimates on the
725: other mixing efficiencies, although perhaps with less satisfaction.  A lower
726: bound on ${\cal E}_{0}$ requires an upper estimate on $\langle \theta^{2}
727: \rangle$.  Starting from (\ref{elphaba2}), recalling that $\theta$ has spatial
728: mean zero and invoking Poincar\'e's inequality on the left and Cauchy--Schwarz
729: on the right,
730: \begin{equation}
731: \kappa \frac{2\pi}{L} \langle \theta^{2} \rangle^{1/2}
732: \langle |\nabla \theta|^{2} \rangle^{1/2} \le
733: \kappa \langle |\nabla \theta|^{2} \rangle = \langle s \theta \rangle
734: \le \langle |\nabla^{-1}s|^{2} \rangle^{1/2} \langle |\nabla \theta|^{2}
735: \rangle^{1/2}.
736: \label{elphaba3}
737: \end{equation}
738: Hence we deduce the upper estimate on the scalar variance,
739: \begin{equation}
740: \langle \theta^{2} \rangle \le
741: \frac{L^{2}}{4 \pi^{2} \kappa^{2}} \langle |\nabla^{-1}s|^{2} \rangle.
742: \label{elphaba4}
743: \end{equation}
744: The unstirred variance is
745: \begin{equation}
746: \langle \theta_{0}^{2} \rangle =
747: \frac{1}{\kappa^{2}} \langle (\Delta^{-1} s)^{2} \rangle,
748: \label{elphaba5}
749: \end{equation}
750: so we have the lower estimate
751: \begin{equation}
752: {\cal E}_{0}^{2} =
753: \frac{\langle \theta_{0}^{2} \rangle}{\langle \theta^{2} \rangle} \ge
754: \frac{4 \pi^{2}}{L^{2}}
755: \frac{\langle (\Delta^{-1} s)^{2} \rangle}{\langle |\nabla^{-1}s|^{2} \rangle}
756: = \frac{\sum_{{\bm{k}}} (Lk/2\pi)^{-4}|\hat{s}_{\bm{k}}|^{2}}
757: {\sum_{{\bm{k}}} (Lk/2\pi)^{-2}|\hat{s}_{\bm{k}}|^{2}}.
758: \label{elphaba6}
759: \end{equation}
760: This lower estimate is strictly positive but because $Lk/2\pi \ge 1$ it is
761: $\le 1$.  Hence it does not rule out the existence of flows that might {\it
762: increase} scalar variance.  Note as well that this estimate depends
763: explicitly on the functional ``shape'' of the source-sink distribution, a
764: feature that we will find for many estimates---upper and lower---on the mixing
765: efficiencies.
766: 
767: The result can in principle be sharpened by solving the variational problem
768: \begin{equation}
769: \langle \theta^{2} \rangle \le \max_{\vartheta} \,
770: \{ \langle \vartheta^{2} \rangle \, | \,
771: \kappa \langle |\nabla \vartheta|^{2} \rangle = 
772: \langle s \vartheta \rangle
773: \}
774: \label{elphaba7}
775: \end{equation}
776: where the maximization is performed over all $\vartheta(\bm{x})$ satisfying
777: the (periodic) boundary conditions on the domain.
778: The Euler-Lagrange equation for the maximizer $\vartheta_{*}(\bf{x})$ is
779: \begin{equation}
780: 0 = 2 \vartheta_{*} - 2\mu \kappa \Delta \vartheta_{*} - \mu s(\bf{x})
781: \label{elphaba8}
782: \end{equation}
783: where $\mu$ is the Lagrange multiplier enforcing the constraint
784: (\ref{elphaba2}).  In terms of the Fourier coefficients the solution of
785: (\ref{elphaba8}) is straightforward,
786: \begin{equation}
787: \hat{\vartheta_{*}}_{\bm{k}} = \frac{\mu}{2} \, \frac{\hat{s}_{\bm{k}}}
788: {\mu \kappa k^{2} + 1},
789: \label{elphaba9}
790: \end{equation}
791: but $\mu$ is the solution of 
792: \begin{equation}
793: \frac{1}{2} \sum_{{\bm{k}}} \frac{\mu \kappa k^{2}
794: |\hat{s}_{\bm{k}}|^{2}}{(\mu \kappa k^{2}+ 1)^{2}} =
795: \sum_{{\bm{k}}} \frac{|\hat{s}_{\bm{k}}|^{2}}{\mu \kappa k^{2}+ 1}.
796: \label{elphaba10}
797: \end{equation}
798: In general it is difficult to solve (\ref{elphaba10}) for $\mu$,
799: but there is one case where it is easy: if the source is 
800: ``monochromatic'', i.e., involves only a single
801: wavenumber of amplitude $k_{s}$, then $\mu = -2/\kappa k_{s}^{2}$ so
802: \begin{equation}
803: {\cal E}_{0}^{2} \ge \frac{\sum_{{\bm{k}}} |\hat{\theta}_{0\bm{k}}|^{2}}
804: {\sum_{{\bm{k}}} |\hat{\theta}_{\bm{k}}|^{2}}
805: = \frac{4\sum_{{\bm{k}}} |\hat{s}_{\bm{k}}|^{2}/(\mu \kappa k^{2})^{2}}
806: {\sum_{{\bm{k}}} |\hat{s}_{\bm{k}}|^{2}/(\mu \kappa k^{2}+ 1)}
807: = 1.
808: \label{elphaba11}
809: \end{equation}
810: Hence stirring a monochromatic source can never increase the variance.
811: However if the source-sink distribution involves even just two distinct
812: wavenumbers, then the solution of (\ref{elphaba10})---which must be performed
813: numerically---yields a value for $\mu$ that produces a lower bound for
814: ${\cal E}_{0}$ that is strictly less than 1~\cite{ShawGFD2005}.  
815: Further details are relegated to Appendix A; this analysis does not 
816: prove that there actually is some stirring that can
817: increase the scalar variance for such sources, but it leaves open the
818: possibility.  For the purposes of this study we will settle for this lower
819: bound as far as its P\'eclet number scaling is concerned.  That is,
820: \begin{equation}
821: {\cal E}_{0} \ge c \, \Pe^{0},
822: \label{elphaba12}
823: \end{equation}
824: where the positive coefficient $c$ may depend on the source-sink
825: distribution, and we cannot rule out that it might be less than 1.
826: 
827: A lower bound on the large-scale mixing efficiency follows from the constraint
828: (\ref{elphaba2}) via Poincar\'e's and the Cauchy--Schwarz inequalities as well.
829: It follows from (\ref{elphaba4}) that
830: \begin{equation}
831: \frac{16\pi^{4}}{L^{4}} \langle |\nabla^{-1}\theta|^{2} \rangle \le
832: \frac {\langle |\nabla^{-1} s|^{2} \rangle} {\kappa^{2}}
833: \label{elphaba13}
834: \end{equation}
835: so that
836: \begin{equation}
837: {\cal E}_{-1}^{2} \ = \
838: \frac{\langle |\nabla^{-1}\theta_{0}|^{2} \rangle}{\langle |\nabla^{-1}\theta|^{2} \rangle}
839: \ge \frac{16 \pi^{4}}{L^{4}} \frac{\langle |\nabla^{-3} s|^{2} \rangle}{ \langle |\nabla^{-1} s|^{2} \rangle}
840: \ = \ \frac{\sum_{{\bm{k}}} (Lk/2\pi)^{-6}|\hat{s}_{\bm{k}}|^{2}}
841: {\sum_{{\bm{k}}} (Lk/2\pi)^{-2}|\hat{s}_{\bm{k}}|^{2}} \ \le \ 1.
842: \label{elphaba14}
843: \end{equation}
844: We note again that for the special case of a monochromatic source, a
845: variational formulation as in (\ref{elphaba7}) yields the improved lower bound
846: ${\cal E}_{-1} \ge 1$.
847: It remains an open problem to determine if these lower estimates $\sim \Pe^{0}$
848: are sharp for arbitrary sources and sinks stirred by some statistically homogeneous 
849: and isotropic flow.  (We will show that they are sharp in some particular cases.)
850: Certainly it will be necessary to use more than just (\ref{elphaba2}) to answer
851: this question in general.
852: 
853: \subsection{Upper bounds on ${\cal E}_{0}$}
854: \label{sec:bound0}
855: 
856: This analysis begins by multiplying the advection--diffusion equation
857: \eqref{eq:AD} by a smooth, time-independent, spatially periodic ``projector
858: function'' $\varphi(\bm{x})$ and taking the space-time average and integrating
859: by parts to obtain
860: \begin{equation}
861:   0 \ = \ \langle \theta (\bm{u}\cdot \nabla + \kappa \lapl)\varphi \rangle
862:   +\langle \varphi s \rangle.
863:   \label{eq:firststep}
864: \end{equation}
865: Because this constraint holds for all~$\varphi$, a lower bound on the variance
866: is
867: \begin{equation}
868:   \langle \theta^2 \rangle \ \geq \ \max_{\varphi} \min_{\vartheta}
869:   \{ \langle \vartheta^2 \rangle~|~0 = \langle \vartheta
870:    (\bm{u} \cdot \nabla \varphi + \kappa \lapl \varphi) \rangle +
871:    \langle \varphi s \rangle \}
872: \label{lbvar}
873: \end{equation}
874: where $\vartheta(\bm{x},t)$ varies over all spatially periodic function with
875: unconstrained dependence.\footnote{Previously, Thiffeault, Doering
876: \& Gibbon~\cite{Thiffeault2004} derived a
877: bound on this variance without optimizing over $\varphi$.}  
878: This min-max variational formulation is equivalent to
879:  \begin{equation}
880:   \langle \theta^2 \rangle \ \geq \ \min_{\vartheta }
881:   \{ \langle \vartheta^2 \rangle~|~ 
882:   \overline{\bm{u} \cdot \nabla \vartheta } = 
883:   \kappa \lapl \overline{\vartheta} + s\}.
884:  \label{glinda1}
885: \end{equation}
886: The multiplier function $\varphi(\bm{x})$ plays the role of a Lagrange
887: multiplier to impose the ``Reynolds averaged'' advection--diffusion equation
888: $\overline{\bm{u} \cdot \nabla \vartheta } = \kappa \lapl \overline{\vartheta}
889: + s$ as a constraint.  The formulation of the bound as a min-max 
890: problem in (\ref{lbvar}) is, as we will see, convenient for its solution.
891: 
892: The minimization over $\vartheta$ in (\ref{lbvar}) is equivalent to an
893: application of the Cauchy--Schwarz inequality to \eqref{eq:firststep}:
894: \begin{equation}
895:   \langle \theta^2 \rangle \geq \max_{\varphi} \frac{ \langle \varphi s
896:   \rangle^2}{\langle (\bm{u}\cdot \nabla \varphi+ \kappa \lapl\varphi)^2
897:   \rangle} = \max_{\varphi} \frac{ \langle \varphi s \rangle^2}{\langle
898:   \varphi\, \overline{{\cal{L}}{\cal{L}}^\dagger} \varphi \rangle}\,,
899:   \label{eq:varb}
900: \end{equation}
901: where we defined the advection--diffusion operator~$\cal{L}$ and its
902: adjoint~${\cal{L}}^\dagger$,
903: \begin{equation}
904:   {\cal{L}}\ldef \bm{u}\cdot \nabla -\kappa
905:   \lapl \qquad \text{and} \qquad
906:   {\cal{L}}^\dagger\ldef -\bm{u}\cdot \nabla -\kappa
907:   \lapl\,.
908: \end{equation}
909: We explicitly indicate the time average in the denominator
910: of~\eqref{eq:varb} remembering that~$\varphi$ is time-independent.  
911: An important point here is that we can average the time-dependent
912: self-adjoint operator ${\cal{L}}{\cal{L}}^\dagger$ before
913: carrying out the maximization over~$\varphi$.
914: 
915: Maximizing~\eqref{eq:varb} over $\varphi$ is equivalent to minimizing its
916: denominator.  Without loss of generality, since the functional~\eqref{eq:varb}
917: is homogeneous in~$\varphi$, we constrain $\varphi$ to have unit projection
918: onto the source.  Thus we must minimize the functional
919: \begin{equation}
920:   {\cal{F}}\ldef\left \langle \tfrac{1}{2} \varphi\,
921:   \overline{{\cal{L}}{\cal{L}}^\dagger} \varphi -\mu (\varphi s -1) \right
922:   \rangle,
923: \end{equation}
924: leading to the Euler--Lagrange equation
925: \begin{equation}
926:   0 = \frac{\delta{\cal{F}}}{\delta \varphi}=
927:   \overline{{\cal{L}}{\cal{L}}^\dagger}\, \varphi - \mu s
928: \end{equation}
929: where $\mu$ is a Lagrange multiplier to enforce the constraint $\langle \varphi s \rangle = 1$.  
930: The minimizer is then
931: \begin{equation}
932:   \varphi = \frac{(\overline{{\cal{L}}{\cal{L}}^\dagger})^{-1} s}
933:   {\langle s (\overline{{\cal{L}}{\cal{L}}^\dagger})^{-1} s \rangle}.
934:   \label{eq:phiopt0}
935: \end{equation}
936: Inserting~\eqref{eq:phiopt0} into~\eqref{eq:varb}, we obtain the lower bound
937: \begin{equation}
938:   \langle \theta^2 \rangle \geq \langle s
939:   (\overline{{\cal{L}}{\cal{L}}^\dagger})^{-1} s \rangle =
940:   \langle s \,\{ \kappa^2 \lapl^2 - \nabla
941:   \cdot(\overline{\bm{u}\bm{u} })+\kappa(2 \nabla
942:   \overline{\bm{u}}:\nabla \nabla + \lapl \overline{\bm{u}} \cdot
943:   \nabla) \}^{-1} s\rangle.
944:   \label{eq:sM0s}
945: \end{equation}
946: Interestingly, this estimate depends only on the mean and equal-point
947: correlation of the stirring.
948: 
949: Specializing to flows satisfying the assumptions of statistical homogeneity
950: and isotropy in (\ref{mean1}) and (\ref{eq:SHIT})---actually we just use
951: (\ref{mean1}) and (\ref{eq:HIT1}) here---we can carry out the time average in
952: \eqref{eq:sM0s}, yielding
953: \begin{equation}
954:   \langle \theta^2 \rangle \geq
955:   \langle s(\overline{{\cal{L}}{\cal{L}}^\dagger})^{-1}s \rangle
956:   = \langle s\, \{ \kappa^2 \lapl^2 -
957:   (U^2/d)\lapl \}^{-1} s\rangle
958:   = \sum_{\bm{k}} \frac{|\hat{s}_{\bm{k}}|^2}{\kappa^{2} k^4 +U^2 k^2/d}.
959:   \label{eq:var0HITbound}
960: \end{equation}
961: Using $\theta_{0} = (-\kappa \Delta)^{-1}s$ we express this result as an upper
962: bound on the mixing efficiency ${\cal E}_{0}$:
963: \begin{equation}
964: {\cal E}_{0}^2 = \frac{\langle \theta_{0}^{2} \rangle}{\langle \theta^{2} \rangle}
965: \leq \frac{\langle s\, \lapl^{-2}\,s \rangle} 
966: {\langle s\, (\lapl^2 - (\Pe^2/L^2 d)\,\lapl)^{-1}\,s \rangle}
967: = \frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^4}
968: {\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/(k^4 +k^2 \Pe^2 / L^{2}d)}
969: \label{eq:varlubound}
970: \end{equation}
971: where the dimensionless P\'eclet number $\Pe = UL/\kappa$ has been inserted.
972: We observe that like the lower estimate on ${\cal E}_{0}$ in (\ref{elphaba6}),
973: the upper bound on the mixing efficiency depends on the shape of the
974: source-sink distribution.
975: 
976: \subsection{Upper bounds on ${\cal E}_{1}$}
977: \label{sec:bound1}
978: 
979: The gradient variance $\langle |\nabla \theta|^2 \rangle$ can quickly and easily
980: be bounded from below in a similar manner to the variance.
981: Begin with Eq.~\eqref{eq:firststep}, integrate by parts, and apply the
982: Cauchy--Schwarz inequality to obtain
983: \begin{equation}
984:   \langle \varphi s \rangle^2 = \langle ( \bm{u} \varphi + \kappa \nabla
985:   \varphi) \cdot \nabla \theta \rangle^2 \leq \langle | \bm{u} \varphi +
986:   \kappa \nabla \varphi|^2 \rangle \langle | \nabla \theta|^2 \rangle
987:   \label{eq:constr1}
988: \end{equation}
989: so that
990: \begin{equation}
991:   \langle |\nabla \theta |^2 \rangle \geq \max_{\varphi} \frac{\langle \varphi
992:   s \rangle^2}{\langle |\bm{u} \varphi + \kappa \nabla \varphi|^2
993:   \rangle}.
994:   \label{eq:gvbound}
995: \end{equation}
996: The right-hand side of~\eqref{eq:gvbound} is homogeneous in~$\varphi$
997: so we minimize the denominator subject to the constraint that 
998: $\langle \varphi s \rangle = 1$.  
999: Under the homogeneity and isotropy assumptions  (\ref{mean1}) and
1000: (\ref{eq:HIT1}), the challenge becomes to evaluate
1001: \begin{equation}
1002:   \min_{\varphi} ~\{\langle \kappa | \nabla \varphi|^2 + U^2\varphi^2 \rangle
1003:   ~|~ \langle \varphi s \rangle=1\}.
1004: \end{equation}
1005: Following a similar development as in Section~\ref{sec:bound0}, the solution is found to be
1006: \begin{equation}
1007: \langle |\nabla \theta |^2 \rangle \geq
1008: \langle s (-\kappa^2 \lapl  + U^2)^{-1} s \rangle
1009: \end{equation}
1010: and the mixing efficiency at small scales is bounded according to
1011: \begin{equation}
1012:   {\cal E}_{1}^{2} \leq \frac{ \langle s\,(-\lapl)^{-1}\,s
1013:   \rangle}{\langle s (-\lapl  + \Pe^2/L^2)^{-1} s \rangle}
1014:   = \frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^2}
1015: {\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/(k^2 + \Pe^2 / L^{2})}.
1016: \label{E0cheap}
1017: \end{equation}
1018: Like the upper bound for ${\cal E}_{0}$ in (\eqref{eq:varlubound}), this
1019: estimate depends on the functional structure of the sources and sinks, and on
1020: the statistically homogeneous and isotropic stirring only through the P\'eclet
1021: number.
1022: 
1023: We can improve this upper estimate by avoiding the application of the
1024: Cauchy--Schwartz inequality that led to~\eqref{eq:constr1}.  We used that
1025: inequality above for expediency, but in fact we expect the bound to involve
1026: only the gradient (i.e., curl-free) part of the field $\bm{u}\varphi+\kappa
1027: \nabla \varphi$.  This can be seen by solving the full min-max variational
1028: problem
1029: \begin{equation}
1030:   \langle |\nabla \theta |^2 \rangle \geq
1031:   \max_{\varphi} \min_{\vartheta}
1032:   \{ \langle | \nabla \vartheta|^2 \rangle ~|~
1033:   \langle \varphi s \rangle = \langle (\bm{u} \varphi + \kappa \nabla
1034:   \varphi) \cdot \nabla \vartheta \rangle \}.
1035: \label{337}
1036: \end{equation}
1037: The minimization is straightforward:
1038: \begin{equation}
1039:   \langle | \nabla \theta |^2 \rangle  \ge \max_{\varphi}
1040:   \frac{ \langle \varphi s \rangle^2}{\langle (\nabla \cdot \bm{w})
1041:   (-\lapl^{-1})\nabla \cdot \bm{w} \rangle}
1042:   \label{eq:varsolsharper}
1043: \end{equation}
1044: where $\bm{w}=\bm{u}\varphi + \kappa \nabla \varphi$.
1045: The vector field $\bm{w}$ can be decomposed as
1046: \begin{eqnarray}
1047:   &&\bm{w}=\underbrace{\bm{w} -\nabla \lapl^{-1} \nabla \cdot
1048:   \bm{w} }+ \underbrace{\nabla \lapl^{-1} \nabla \cdot \bm{w}}\\
1049:   &&~~~~~~~{{\text{divergence-free}}~~~~~~~{\text{curl-free}}}\nonumber
1050: \end{eqnarray}
1051: where the two components are orthogonal.
1052: Only the curl-free part contributes in (\ref{eq:varsolsharper}).
1053: That is, the denominator is just the norm of the curl-free portion:
1054: \begin{align}
1055:   \langle (\nabla \cdot \bm{w} (-\lapl^{-1}) (\nabla \cdot \bm{w})
1056:   \rangle &= \langle [\lapl \lapl^{-1} (\nabla \cdot
1057:   \bm{w})](-\lapl^{-1}) (\nabla \cdot \bm{w}) \rangle \nonumber\\
1058:   &= \langle \nabla (\lapl^{-1} (\nabla
1059:   \cdot \bm{w}) \cdot \nabla (\lapl^{-1} (\nabla \cdot \bm{w})\rangle
1060:   \nonumber\\
1061:   &= \langle | \nabla \lapl^{-1} \nabla \cdot \bm{w}|^2 \rangle
1062:   \ \le \ \langle |\bm{w}|^2 \rangle.
1063: \end{align}
1064: Hence (\ref{eq:varsolsharper}) generally represents an improvement over the
1065: expression resulting from application of the Cauchy--Schwarz inequality in
1066: (\ref{eq:gvbound}).
1067: 
1068: This improved bound depends on the full two-point correlation function of the
1069: velocity field:
1070: \begin{eqnarray}
1071:   \langle (\nabla \cdot \bm{w}) (-\lapl^{-1})\nabla \cdot \bm{w}
1072:   \rangle &=& \frac{1}{L^d} \int d^{d}x \int d^{d}y \
1073:   \overline {\nabla_{\bm{x}} \cdot \bm{w}(\bm{x},\cdot) 
1074:   G( \bm{x} - \bm{y}) \nabla_{\bm{y}} \cdot
1075:   \bm{w}(\bm{y},\cdot)} \nonumber \\
1076:   &=&\frac{1}{L^d} \int d^{d}x \int d^{d}y \
1077:   (-\nabla_{\bm{x}} \nabla_{\bm{x}} G):
1078:   \overline{\bm{w}(\bm{x},\cdot)\bm{w}(\bm{y},\cdot)}
1079: \end{eqnarray}
1080: where $G(\bm{x}-\bm{y})$ is the Green's function for $-\lapl$ on spatially
1081: mean-zero functions with Fourier coefficients $\hat{G}(k)=1/(L^{d}k^{2})$ for
1082: $\bm{k} \ne 0$.
1083: %Integrating by parts,
1084: %\begin{equation}
1085: %  \langle (\nabla \cdot \bm{w}) (-\lapl^{-1})\nabla \cdot \bm{w}
1086: %  \rangle = \frac{1}{L^d} \int d^{d}x \int d^{d}y \
1087: %  (-\nabla_{\bm{x}} \nabla_{\bm{x}} G):
1088: %  \overline{\bm{w}(\bm{x},\cdot)\bm{w}(\bm{y},\cdot)}.
1089: %\end{equation}
1090: Under the assumptions (\ref{mean1}) and (\ref{eq:SHIT}) 
1091: of statistical homogeneity and isotropy,
1092: \begin{eqnarray}
1093:   \overline{w_{i}(\bm{x},\cdot)w_{j}(\bm{y},\cdot)} &=&
1094:   \varphi(\bm{x}) \varphi(\bm{y})
1095:   \overline{u_{i}(\bm{x},\cdot)u_{j}(\bm{y},\cdot)} +
1096:   \kappa^2 \partial_{i}\varphi(\bm{x}) \partial_{j}\varphi(\bm{y})
1097:   \nonumber \\
1098:   &=& C_{ij}(\bm{x}-\bm{y})\varphi(\bm{x}) \varphi(\bm{y})+
1099:   \kappa^2 \partial_{i}\varphi(\bm{x}) \partial_{j}\varphi(\bm{y}).
1100: \end{eqnarray}
1101: In terms of Fourier transformed variables,
1102: \begin{equation}
1103:   \langle (\nabla \cdot \bm{w}) (-\lapl^{-1})\nabla \cdot \bm{w} \rangle = 
1104:   \sum_{\bm{k},\bm{k'} \ne 0} \frac{\hat{C}(k)}{d-1}
1105:   \left( 1-\frac{(\bm{k} \cdot \bm{k'})^{2}}{k^{2}k'^{2}}\right)
1106:   | \hat{\varphi}_{\bm{k}+\bm{k'}}|^{2} +
1107:   \sum_{\bm{k} \ne 0} \left( \frac{\hat{C}(0)}{d} + \kappa^{2} k^{2} \right)
1108:   | \hat{\varphi}_{\bm{k}}|^{2}.
1109: \end{equation}
1110: Thus the mixing efficiency at small scales is actually bounded from above according to
1111: \begin{eqnarray}
1112: {\cal E}_{1}^{2} &\leq&
1113: \min_{\varphi} \ \frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^2}
1114: {\left[ \sum_{\bm{k}} \hat{s}_{\bm{k}}^{*}\hat{\varphi}_{\bm{k}}\right]^{2} }
1115: \ \times \nonumber \\
1116: &\times& 
1117: \left[
1118:  \sum_{\bm{k},\bm{k'} \ne 0} \frac{\hat{C}(k)}{(d-1) \kappa^{2} }
1119:   \left( 1-\frac{(\bm{k} \cdot \bm{k'})^{2}}{k^{2}k'^{2}}\right)
1120:   | \hat{\varphi}_{\bm{k}+\bm{k'}}|^{2} +
1121:   \sum_{\bm{k} \ne 0} \left( \frac{\hat{C}(0)}{d \kappa^{2}} + k^{2} \right)
1122:   | \hat{\varphi}_{\bm{k}}|^{2}.
1123: \right]
1124: \label{IB}
1125: \end{eqnarray}
1126: This upper bound depends on details of the full spectrum of the stirring
1127: velocity field.
1128: 
1129: We will not perform the optimization over $\varphi$ for the general problem
1130: here; the implications of two-point statistical properties of the stirring on
1131: the P\'eclet number dependence of this bound on ${\cal E}_{1}$ will be left
1132: for future investigations.  There is one case, however, where the optimization
1133: can easily be carried out that shows that the result of (\ref{IB}) may indeed
1134: be a quantitative improvement over (\ref{E0cheap}).  That simple case is when
1135: the spectrum of the velocity field is concentrated at $\bm{k} = 0$, i.e., when
1136: the velocity field is at (almost) every moment of time a uniform ``wind'' in
1137: space.  For such spatially uniform statistically homogeneous and isotropic
1138: flows, $\hat{C}(0) = U^{2}$ while all the other $\hat{C}(k) = 0$ for $\bm{k}
1139: \ne 0$.  Then
1140: \begin{eqnarray}
1141: {\cal E}_{1}^{2} \ \leq \
1142: \min_{\varphi} \ \frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^2}
1143: {\left[ \sum_{\bm{k}} \hat{s}_{\bm{k}}^{*}\hat{\varphi}_{\bm{k}} \right]^{2}}
1144:   \sum_{\bm{k} \ne 0} \left( \frac{U^{2}}{d \kappa^{2}} + k^{2} \right)
1145:   | \hat{\varphi}_{\bm{k}}|^{2}
1146:  \ = \ \frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^2}
1147: {\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/(k^2 + \Pe^2 / dL^{2})}.
1148: \label{IB2}
1149: \end{eqnarray}
1150: In this case the improvement over (\ref{E0cheap}) is just the extra factor of
1151: the spatial dimension $d$ in the denominator of the denominator of the
1152: denominator in the last term.
1153: 
1154: \subsection{Upper bounds on ${\cal E}_{-1}$}
1155: \label{sec:boundm1}
1156: 
1157: To derive a lower bound on the inverse-gradient variance
1158: $\langle |\nabla^{-1} \theta|^2 \rangle$ we begin
1159: again with (\ref{eq:firststep}), insert $\lapl\lapl^{-1}=1$, 
1160: integrate by parts and apply the Cauchy--Schwarz
1161: inequality to obtain
1162: \begin{align}
1163:   \langle \varphi s \rangle \ = \ \langle \nabla (\bm{u} \cdot \nabla \varphi +
1164:   \kappa \lapl \varphi) \cdot \nabla \lapl^{-1} \theta \rangle
1165:   % \nonumber \\
1166:   \ \leq \ \langle | \nabla (\bm{u} \cdot \nabla \varphi + \kappa \lapl
1167:   \varphi) |^2 \rangle ^{\frac{1}{2}} \langle | \nabla^{-1} \theta| ^2 \rangle
1168:   ^{\frac{1}{2}}.
1169: \end{align}
1170: This gives a lower bound on the inverse-gradient variance
1171: \begin{equation}
1172:   \langle | \nabla^{-1} \theta|^2 \rangle \geq\max_{\varphi} \frac{ \langle
1173:   \varphi s \rangle^2 }{ \langle | \nabla(\bm{u}\cdot \nabla
1174:   \varphi+\kappa \lapl \varphi|^2\rangle}\,.
1175:   \label{Oz}
1176: \end{equation}
1177: Recalling that $\varphi$ is time-independent and restricting attention to statistically
1178: homogeneous and isotropic flows---assuming as well that the enstrophy 
1179: $\Omega^{2} = \langle |\nabla \bm{u}|^{2} \rangle$ is finite---the
1180: denominator is
1181: \begin{equation}
1182:   |\nabla \bm{u} \cdot \nabla \varphi+ \bm{u} \cdot \nabla \nabla
1183:   \varphi + \kappa \nabla \lapl \varphi  |^2 =(\Omega^2/d) | \nabla \varphi|^2 + (U^2/d)(\lapl
1184:   \varphi)^2+ \kappa^2
1185:   |\lapl \nabla \varphi|^2 .
1186: \end{equation}
1187: Then the optimization over $\varphi$ in (\ref{Oz}) is straightforward.  Recalling that
1188: \begin{equation}
1189: \langle | \nabla^{-1} \theta_0|^2 \rangle =
1190: \langle  | \nabla^{-1} \lapl^{-1} s |^2 \rangle/\kappa^{2}
1191: \end{equation}
1192: and the definition $\lambda = U/\Omega$,
1193: we conclude that
1194: \begin{eqnarray}
1195: {\cal E}_{-1}^2 &\leq& \frac{ \langle
1196:   | \nabla^{-1} \lapl^{-1} s |^2 \rangle}
1197:        { \langle s \l(-\lapl^3 + (\Pe^2/L^2\,d)
1198: \lapl^2 - (\Pe^2/\lambda^2 L^2\,d) \lapl\r)^{-1} s \rangle}
1199: \nonumber \\
1200: &=& 
1201: \frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^6}
1202: {\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/(k^6 +k^4 \Pe^2 / L^{2}d
1203: +k^2 \Pe^2 / \lambda^{2} L^{2}d)} .
1204: \label{eq:igvarlubound}
1205: \end{eqnarray}
1206: 
1207: 
1208: Compare this to (\ref{eq:varlubound}) for the upper estimate on ${\cal E}_{0}$
1209: and (\ref{E0cheap}) and (\ref{IB}) for the upper estimate on ${\cal E}_{1}$.
1210: Note that the efficiency depends on the enstropy in the flow, i.e., the flow's
1211: ``shear'' or ``strain'' content, directly through $\lambda = U/\Omega$.
1212: Interestingly, the $\Omega$ (or $\lambda$) term allows for an increase in the
1213: bound on the mixing efficiency on large scales via stirring on small scales,
1214: something that is absent in the bounds on the mixing efficiency at
1215: intermediate and small scales.
1216: 
1217: %===================================================
1218: 
1219: \section{Saturating the Multi-scale Mixing Efficiency Bounds}
1220: \label{sec:saturation}
1221: 
1222: The upper bounds on the mixing efficiencies ${\cal E}_{p}$ derived in the
1223: previous section depend on the entire source-sink distribution functions, but
1224: on just a few features ($U$, and for ${\cal E}_{-1}$, $\lambda=U/\Omega$) of
1225: the statistically homogeneous and isotropic stirring field.  In this section
1226: we show that there is at least one combination of sources, sinks and stirring
1227: strategies that saturate the upper estimates exactly at all P\'eclet numbers.
1228: This establishes that on the highest level of generality the upper bound
1229: analysis is absolutely sharp.  We also show that there is at least one
1230: combination of sources, sinks and statistically stationary homogeneous
1231: isotropic stirrings that saturate the scaling (if not necessarily the
1232: prefactor) of the lower estimates on the efficiencies $\sim \Pe^{0}$.
1233: 
1234: Consider the simple monochromatic source function
1235: \begin{equation}
1236: s(\bm{x}) \ = \ \sqrt{2} S \sin(k_{s}x_{1})
1237: \label{MS}
1238: \end{equation}
1239: where $S$ is the root mean square amplitude, $2\pi/k_{s}$ is the wavelength,
1240: and $x_{1}$ is one of the $d$ coordinates.
1241: For this monochromatic source the lower bounds on the mixing efficiencies
1242: at all scales are the same, i.e., for $p = -1, 0, 1$,
1243: \begin{equation}
1244: 1 \ \le \ {\cal E}_{p} \quad
1245: \text{ (monochromatic sources)}.
1246: \end{equation}
1247: 
1248: The upper bounds on ${\cal E}_{0}$, ${\cal E}_{1}$ and ${\cal E}_{-1}$ in,
1249: respectively, (\ref{eq:varlubound}), (\ref{E0cheap}) and (\ref{IB2}) are
1250: generally different for monochromatic source-sink distribution functions.  But
1251: if we restrict attention to ``uniform wind'' flows $\bm{u}(\bm{x},t)$ that are
1252: at each instant of time spatially uniform (i.e., $\nabla \bm{u}(\bm{x},t)=0$),
1253: then $\Omega = 0$, $\lambda \rightarrow \infty$, and we can take advantage of
1254: the improved bound for ${\cal E}_{1}$ in (\ref{IB2}) to see that the upper
1255: estimates are all the same.  For $p = -1, 0, 1$,
1256: \begin{equation}
1257: {\cal E}_{p} \ \leq \ \sqrt{1+\Pe^2/ L^{2}k_{s}^{2}d}
1258: \quad \text{ (monochromatic sources \& uniform winds)}.
1259: \label{UMU}
1260: \end{equation}
1261: 
1262: In order to show that the upper bounds in (\ref{UMU}) are sharp, we construct
1263: a family of statistically homogeneous and isotropic flows that approach these
1264: limits.  The trick is to sustain a uniform wind in a given direction for a
1265: ``long'' time so that the steady state is very nearly achieved before
1266: switching to a uniform wind in another direction.  The dynamic transients
1267: between the changes in flow configurations decay at least at rate $\kappa
1268: k_{s}^{2}$, so when the transitions are sufficiently infrequent the scalar
1269: field is almost always (nearly) in a static configuration.  Then we may solve
1270: the steady flow problem for the scalar field exactly and average the variances
1271: over the wind directions to evaluate the efficiencies in the limit of slow
1272: switches among the directions.  This adiabatic averaging method can be
1273: implemented for other flows, too, as will be considered in
1274: Section~\ref{sec:SSSSSF}.
1275: 
1276: Consider uniform winds of speed $U$ blowing along the ``diagonals'', in the
1277: $2^{d}$ directions given by unit vectors $\frac{1}{\sqrt{d}} \left( \pm
1278: \hat{\bm{e}}_{1} \dots \pm \hat{\bm{e}}_{d} \right)$.  Each of these
1279: directions is equivalent, so we can solve any single problem to evaluate the
1280: variances.  The steady advection--diffusion equation with source
1281: $s(\bm{x})=\sqrt{2}S \sin (\ks x_1)$ and uniform stirring field
1282: $\bm{u}(\bm{x})=\frac{U}{\sqrt{d}}\sum_{j=1}^{d} \hat{\bm{e}}_{j}$ is
1283: \begin{equation}
1284:   \frac{U}{\sqrt{d}} \sum_{j=1}^d \frac{\partial \theta}{\partial x_j} =
1285:   \kappa \sum_{j=1}^d \frac{\partial^2 \theta}{\partial x_j^2} + \sqrt{2} S
1286:   \sin(\ks x_1).
1287: \end{equation}
1288: The solution is of the form $\theta(\bm{x})=\sum_{j=1}^dF^{(j)}(x_j)$ where
1289: the functions $F^{(j)}(x_j)$ satisfy the system of constant coefficient ODEs
1290: \begin{equation}
1291: \begin{split}
1292:   \frac{d^2F^{(1)}}{dx_1^2} - \frac{U}{\sqrt{d}\,\kappa}\, \frac{dF^{(1)}}{dx_1}
1293:   +\frac{\sqrt{2}S}{\kappa}\, \sin(\ks x_1/L)&=0,\\
1294:   \frac{d^2F^{(j)}}{dx_n^2} - \frac{U}{\sqrt{d}\,\kappa}\, \frac{dF^{(j)}}{dx_n}
1295:   &=0,\qquad\text{for $2 \leq j \leq d$,}
1296: \end{split}
1297: \end{equation}
1298: with periodic boundary conditions $F^{(j)}(0)=F^{(j)}(L)$.  Recalling that
1299: $\theta$ has, without loss of generality, spatial mean zero, the solution is
1300: \begin{equation}
1301: \begin{split}
1302:   F^{(1)} &= \frac{\sqrt{2}S}{\kappa k_{s}^{2} + U^2/d\kappa}\left[
1303:     \sin(\ks x_1) - \frac{U}{\kappa k_{s} \sqrt{d}}\, \cos(\ks x_1) \right],\\
1304:   F^{(j)} &= 0,\qquad\text{for\ } 2\leq j \leq d.
1305: \end{split}
1306: \end{equation}
1307: The variance is thus
1308: \begin{equation}
1309:   \langle \theta^2 \rangle = \frac{S^2}{\kappa^{2} k_{s}^{4}
1310:   + k_{s}^{2}U^{2}/d},
1311: \end{equation}
1312: and because the scalar field is monochromatic, the small-scale and large-scale
1313: variances simply satisfy $k_{s}^{2} \langle |\nabla^{-1} \theta|^2 \rangle =
1314: \langle |\nabla \theta|^2 \rangle/k_{s}^{2} =\langle \theta^2 \rangle$.  The
1315: multi-scale mixing efficiencies are all then
1316: \begin{equation}
1317: {\cal E}_{p} = \sqrt{\frac{\langle  |\nabla^{p} \theta_{0}|^2 \rangle}
1318: {\langle  |\nabla^{p} \theta|^2 \rangle}}
1319: = \sqrt{1+ \frac{\Pe^2}{k_{s}^{2}L^{2}d}}
1320: \end{equation}
1321: precisely as in (\ref{UMU}).
1322: 
1323: On one hand this result is fairly intuitive: the most efficient way to reduce the variance 
1324: (on any length scale) is to direct the flow from source regions directly toward the 
1325: closest convenient sink regions and the likewise from sinks toward sources---{\it if} 
1326: this can be accomplished effectively given the constraints of incompressibility 
1327: and statistical homogeneity and isotropy.  
1328: This can be done simply for these monochromatic flows on the torus, 
1329: and we have discovered that such flows actually suppress the scalar
1330: variance at all scales as well as is possible for {\it any} 
1331: statistically homogeneous and isotropic flow field. 
1332: (Note: this example was inspired by Plasting \& Young~\cite{Plasting2006}
1333: who showed that the steady direct flow across the source saturates the variance 
1334: bound for single-wavenumber sources among {\it all} flows regardless
1335: of any statistical considerations.)
1336: On the other hand this type of sweeping flow is somewhat pathological in the 
1337: sense that it simply transports the source onto the sink without ``stirring''
1338: and ``mixing'' by the usual meanings of those words.
1339: In some cases, such as transient mixing as discussed in the introduction, the 
1340: role of effective stirring is to stretch material lines and amplify gradients 
1341: of the passive scalar to accelerate the action of molecular diffusion 
1342: to dissipate variance on small scales.
1343: 
1344: It is not presently clear whether the upper bounds in (\ref{eq:varlubound}),
1345: (\ref{IB}) and (\ref{IB2}) can be saturated for more general
1346: sources and sinks.
1347: Moreover, this particular saturation result depends explicitly on the 
1348: geometry of the domain.
1349: In geophysical and astrophysical applications, for example, it is natural
1350: to consider the domain to be the surface of a sphere.  
1351: While such a domain admits similar concepts of statistical homogeneity
1352: and isotropy for the flow and ``monochromaticity'' for the source-sink 
1353: distribution (eigenfunctions of the Laplacian on the sphere), there is
1354: no analogous uniform sweeping flow as there is on the torus.  
1355: This makes it clear that the optimal stirrer is a function of both
1356: the source shape and the domain.
1357: Formulating the optimization problem for the best stirring field for a given
1358: source-sink distribution remains a problem for future investigation.
1359: 
1360: We close this section by pointing out that a similar statistically homogeneous
1361: and isotropic uniform wind can also be arranged to saturate the high-$\Pe$ scaling
1362: of the {\em lower} bounds on the mixing efficiencies.  For the monchromatic source
1363: (\ref{MS}), consider uniform winds oriented along the coordinate axes in the
1364: $2d$ directions $\pm \hat{\bm{e}}_{j}$, switching only after blowing each way
1365: for a long enough time for the transients to be negligible for the time
1366: averages.  The wind reduces the scalar variances below the variance in
1367: $\theta_{0}$ only when it blows in the $\pm \hat{\bm{e}}_{1}$ directions,
1368: which occurs $1/d$ of the time.  Averaging the scalar variance adiabatically
1369: over the wind directions yields the mixing efficiencies
1370: \begin{equation}
1371: {\cal E}_{p} =
1372: \sqrt{\frac{1+\Pe^{2}/k_{s}^{2}L^{2}}{1+(d-1)\Pe^{2}/dk_{s}^{2}L^{2}}}.
1373: \end{equation}
1374: These efficiencies are monotonically increasing in $\Pe$ but bounded according
1375: to
1376: \begin{equation}
1377: 1 \ \le \ {\cal E}_{p} \ < \ \lim_{\Pe \rightarrow \infty}
1378: \sqrt{\frac{1+\Pe^{2}/k_{s}^{2}L^{2}}{1+(d-1)\Pe^{2}/dk_{s}^{2}L^{2}}}
1379: \ = \ \sqrt{\frac{d}{d-1}} \times \Pe^{0}.
1380: \end{equation}
1381: 
1382: 
1383: \section{High-$\Pe$ behavior of the mixing efficiency bounds}
1384: \label{sec:SARSAS}
1385: 
1386: Here we examine the high P\'eclet number behavior of the upper bounds on the
1387: multi-scale mixing efficiencies for statistically homogeneous and isotropic
1388: flows.  As a point of reference we recall that if there is an effective eddy
1389: diffusion associated with the flow field, so that the stirring suppresses the
1390: scalar variances in the manner of enhanced molecular diffusion in the form of
1391: an equivalent (eddy) diffusion $\sim U\ell$, then the efficiencies would scale
1392: ``classically'', as ${\cal E}_{p} \sim \Pe^{1}$, as $\Pe \rightarrow \infty$.
1393: Classical scaling allows for the precise identification of equivalent
1394: diffusivities $\kappa^{(\mathrm{eq})}_{p} \ldef \kappa {\cal E}_{p} \rdef U
1395: \ell_{p}$ that serve as the definition of associated ``mixing lengths''
1396: $\ell_{p}$ that are independent of the magnitude of $U$ and $\kappa$.  The
1397: upper limits on ${\cal E}_{0}$ in (\ref{eq:varlubound}), ${\cal E}_{1}$ in
1398: (\ref{E0cheap}) and ${\cal E}_{-1}$ (\ref{eq:igvarlubound}), reproduced
1399: immediately below for reference, all depend on the full structure of the
1400: source-sink distribution:
1401: \begin{eqnarray}
1402: {\cal E}_{1} &\leq& \sqrt{\frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^2}
1403: {\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/(k^2 + \Pe^2 / L^{2})}} 
1404: \label{E0cheap2a} \\
1405: {\cal E}_{0} &\le& \sqrt{\frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^4}
1406: {\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/(k^4 +k^2 \Pe^2 / L^{2}d)}}
1407: \label{eq:varlubounda} \\
1408: {\cal E}_{-1} &\leq& \sqrt{\frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^6}
1409: {\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/(k^6 +k^4 \Pe^2 / L^{2}d
1410: +k^2 \Pe^2 / \lambda^{2} L^{2}d)} },
1411: \label{eq:igvarluboundb}
1412: \end{eqnarray}
1413: The high-$\Pe$ behavior, though, can be discerned from just a few features
1414: of the sources and sinks.
1415: 
1416: \subsection{Square-integrable sources and sinks}
1417: 
1418: Consider the case where the source-sink distribution function is square integrable,
1419: $s(\bm{x}) \in L^{2}(\mathbb{T}^d)$, so that the Fourier coefficients are square summable:
1420: \begin{equation}
1421: \sum_{\bm{k}} |\hat{s}_{\bm{k}}|^{2} \ < \ \infty.
1422: \end{equation}
1423: Then the $\Pe \rightarrow \infty$ asymptotic behaviors in (\ref{E0cheap2a}), 
1424: (\ref{eq:varlubounda}) and (\ref{eq:igvarluboundb}) are elementary to evaluate:
1425: \begin{eqnarray}
1426: {\cal E}_{1} \ &\lesssim& \ \Pe \
1427: \sqrt{\frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^2}
1428: {L^{2}\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2}} \ \rdef \
1429: \Pe \times \frac{\ell_{1}^{(\mathrm{max})}}{L}
1430: \label{HighPeL2p1}
1431: \\
1432: {\cal E}_{0} \ &\lesssim& \ \Pe \
1433: \sqrt{\frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^4}
1434: {dL^{2}\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^2}} \ \rdef \
1435: \Pe \times \frac{\ell_{0}^{(\mathrm{max})}}{L}
1436: \label{HighPeL2v0}
1437: \\
1438: {\cal E}_{-1}  \ &\lesssim& \ \Pe \
1439: \sqrt{\frac{\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/k^6}
1440: {dL^{2}\sum_{\bm{k}} |\hat{s}_{\bm{k}}|^2/(k^{4}
1441: +k^2 / \lambda^{2})}} \ \rdef \
1442: \Pe \times \frac{\ell_{-1}^{(\mathrm{max})}}{L}.
1443: \label{HighPeL2m1}
1444: \end{eqnarray}
1445: There are three significant features of these scaling bounds worth noting.
1446: 
1447: The first point is that these upper estimates all scale classically, allowing
1448: us to identify the largest possible values for meaningful mixing lengths that
1449: we have labeled $\ell_{p}^{(\mathrm{max})}$.
1450: 
1451: The second remarkable fact is that the largest possible mixing lengths
1452: relevant to small and intermediate scale fluctuations do {\it not} depend on
1453: the flow field, but rather only on the source-sink distribution.  Indeed,
1454: $\ell_{1}^{(\mathrm{max})} \le \ell_{0}^{(\mathrm{max})} \le L/2\pi$ are particular length
1455: scales in the source-sink function that have nothing to do with the stirring
1456: field or any length scales in the flow.  This is in direct conflict with the
1457: notion that a mixing length should be a characteristic persistence length or
1458: eddy size in the velocity vector field.  Rather, these mixing efficiencies are
1459: ultimately limited by the structure of the sources and sinks.  When
1460: $\ell_{1}^{(\mathrm{max})}$ and $\ell_{0}^{(\mathrm{max})}$ are ``small'' (i.e., when
1461: $\ell_{1}^{(\mathrm{max})} \le \ell_{0}^{(\mathrm{max})} \ll L$) then the mixing efficiencies are
1462: also ``small'' and {\it not} subject to any further improvement by any sort of
1463: clever stirring designed to enhance the variance reduction at intermediate and
1464: small scales.
1465: 
1466: The bound on the mixing length $\ell_{-1}$ for large-scale variance reduction,
1467: however, {\it does} depend on the spectrum of length scales in the flow
1468: through (the Taylor microscale) $\lambda$.  It is interesting to note that the
1469: bound on $\ell_{-1}^{(\mathrm{max})}$ is an increasing function of $\lambda^{-1}$,
1470: allowing for the possibility that small-scale stirring could enhance large
1471: scale mixing.  That is, this analysis does not preclude small-scale stirring from
1472: suppressing large-scale fluctuations in ways that it cannot decrease the variance at
1473: intermediate and small scales.
1474: 
1475: We note as well that the improved upper bound on the small-scale mixing
1476: efficiency ${\cal E}_{1}$ in (\ref{IB}) generally produces an even {\it
1477: smaller} estimate for $\ell_{1}^{(\mathrm{max})}$ that depends on the spectrum, i.e.,
1478: the magnitude and distribution, of the length scales in the flow.  In cases
1479: where the shortest length scale (call it $\ell_{s}$) in the source is much
1480: longer than the longest length scale (call it $\ell_{u}$) in the stirring
1481: field, a thoughtful examination of (\ref{IB}) suggests that $\ell_{1}^{(\mathrm{max})}
1482: \sim \ell_{u} \ll \ell_{s}$.
1483: 
1484: The third point worth noting is that the upper estimates for all three of the
1485: ${\cal E}_{p}$ scale the same, $\sim \Pe^{1}$ as $\Pe \rightarrow \infty$.  So
1486: far all the examples we have considered share this property, but in the next
1487: subsection we will see that this is {\it not} a general feature of the
1488: multi-scale mixing efficiency bounds.
1489: 
1490: \subsection{Measure-valued source-sink distributions}
1491: 
1492: If $s(\bm{x}) \notin L^{2}(\mathbb{T}^d)$ then the sum in the denominator of
1493: (\ref{HighPeL2p1}) diverges and the ratio defining the prefactor
1494: ($\ell_{1}^{(\mathrm{max})}/L$) to the $\Pe^{1}$ scaling could vanish.  This
1495: would violate the lower bound ${\cal E}_{1} \ge 1$, so the high-$\Pe$
1496: asymptotic analysis of the ratios of sums must be revisited.  This issue
1497: arises for measure-valued sources and sinks, i.e., when the distribution
1498: involves singular objects like $\delta$-functions.  
1499: Then ``anomalous'' sub-classical P\'eclet number scalings for some
1500: efficiencies are inevitable.
1501: 
1502: Consider the extreme cases where the Fourier coefficients of $s(\bm{x})$ obey
1503: \begin{equation}
1504: |\hat{s}(\bm{k})| = {\cal O}(1) \quad \text{as} \quad 
1505: |\bm{k}| \rightarrow \infty.
1506: \end{equation}
1507: In this case, in spatial dimensions $d=2$ and $3$, sums in both
1508: (\ref{HighPeL2p1}) and (\ref{HighPeL2v0}) diverge so the high-$\Pe$ behavior
1509: of the bounds on ${\cal E}_{1}$ and ${\cal E}_{0}$ must be re-evaluted
1510: directly from (\ref{E0cheap2a}) and (\ref{eq:varlubounda}).  The classical
1511: high-$\Pe$ scaling (\ref{HighPeL2m1}) for the bound on ${\cal E}_{-1}$ from
1512: (\ref{eq:igvarluboundb}) remains the same in $d=2$ and $3$.
1513: 
1514: To discern the high P\'eclet number behavior of the expressions in
1515: (\ref{E0cheap2a}) and (\ref{eq:varlubounda}) when $|\hat{s}(\bm{k})| \sim C =
1516: {\cal O}(1)$ as $|\bm{k}| \rightarrow \infty$, we analyze integral
1517: approximations to the sums.  This is justified because the major contribution
1518: to these diverging sums comes from the high-$k$ end where the discreteness
1519: of the wavenumbers is negligible.  For example, the denominator for the bound
1520: on ${\cal E}_{0}$ in (\ref{eq:varlubounda}) is
1521: \begin{equation}
1522: \sum_{\bm{k}} \frac{|\hat{s}_{\bm{k}}|^2}
1523: {k^4 + k^{2}\Pe^2 / L^{2}d}
1524: \ \approx \ \left( \frac{L}{2\pi} \right)^{d} \ S_{d}
1525: \int_{\fracs{2\pi}{L}}^{\infty} \frac{C \ k^{d-3} dk}
1526: {k^{2}+\Pe^2 / L^{2}d}
1527: \end{equation}
1528: where $S_{2} = 2\pi$ and $S_{3}=4\pi$.
1529: 
1530: In $d=2$ this is
1531: \begin{equation}
1532: \sum_{\bm{k}} \frac{|\hat{s}_{\bm{k}}|^2}
1533: {k^4 + k^{2}\Pe^2 / L^{2}d}
1534: \ \approx \ \frac{L^{2}}{2\pi} \ C \ \frac{L^{2}}{\Pe^{2}}
1535: \ \log \left[1+\frac{\Pe^{2}}{8\pi^{2}} \right],
1536: \end{equation}
1537: so the upper bound on ${\cal E}_{0}$ in $d=2$ is
1538: \begin{equation}
1539: {\cal E}_{0} \ \lesssim \
1540: \frac{\Pe}
1541: {\sqrt{8\pi^{2}\log \left[1+\frac{\Pe^{2}}{8\pi^{2}} \right]}}
1542: \ \sim \ \frac{\Pe}{(\log{\Pe})^{1/2}}
1543: \quad \text{as} \ \Pe \rightarrow \infty.
1544: \end{equation}
1545: This upper bound exhibits a logarithmic correction to classical scaling.  The
1546: significance of this is perhaps best appreciated as the {\it absence} of any
1547: residual variance suppression in the limit of vanishing molecular diffusivity.
1548: That is, the largest possible effective diffusivity defined by the bulk
1549: variance suppression $\kappa_{\mathrm{eff}} \le \kappa {\cal E}_{0} \sim {\cal
1550: O}(|\log{\kappa}|^{-1/2}) \rightarrow 0$ as $\kappa \rightarrow 0$.  It is
1551: worthwhile stressing that this result does not depend on---and cannot be
1552: circumvented by manipulating---any further details of the statistically
1553: homogeneous and isotropic stirring field.
1554: 
1555: The deviation from classical scaling is more dramatic in $d=3$ where
1556: \begin{equation}
1557: \sum_{\bm{k}} \frac{|\hat{s}_{\bm{k}}|^2}
1558: {k^4 + k^{2}\Pe^2 / L^{2}d}
1559: \ \approx \ \left( \frac{L}{2\pi} \right)^{3} \ 4 \pi C \ \frac{\sqrt{3}L}{\Pe}
1560: \ \left[ \frac{\pi}{2} - \arctan \left(\frac{2\pi\sqrt{3}}{\Pe} \right) \right].
1561: \end{equation}
1562: Then the upper bound on ${\cal E}_{0}$ is
1563: \begin{equation}
1564: {\cal E}_{0} \ \lesssim \  
1565: \sqrt{
1566: \frac{\Pe}{
1567: 2 \pi \sqrt{3} \ \left[ \frac{\pi}{2} 
1568: - \arctan \left(\frac{2\pi\sqrt{3}}{\Pe} \right)\right]
1569: }
1570: } 
1571: \ \sim \ \Pe^{1/2}
1572: \quad \text{as} \ \Pe \rightarrow \infty.
1573: \end{equation}
1574: This upper bound exhibits strictly
1575: sub-classical scaling at high P\'eclet numbers.
1576: 
1577: Both the numerator and the denominator diverge in the bound in
1578: (\ref{E0cheap2a}) for ${\cal E}_{1}$ for such measure-valued source-sink
1579: distributions, so those bounds must be evaluated as the limit of ratios for a
1580: sequence of appropriately mollified sources and sinks.  This is
1581: straightforward if we simply truncate the Fourier series for $s(\bm{x})$ at
1582: small length scale $\ell_{s} \ll L$ and study the ideal case where
1583: $|\hat{s}(\bm{k})| = C$ for $2\pi/L \le |\bm{k}| < 2\pi/\ell_{s}$ and
1584: $|\hat{s}(\bm{k})| = 0$ for $|\bm{k}| > 2\pi/\ell_{s}$, and then take the
1585: limit $\ell_{s}/L \rightarrow 0$.
1586: 
1587: Again we analyze integral approximations to the sums;
1588: the numerator in (\ref{E0cheap2a}) is
1589: \begin{equation}
1590: \sum_{\bm{k}} \frac{|\hat{s}_{\bm{k}}|^2}{k^2}
1591: \ \approx \ \left( \frac{L}{2\pi} \right)^{d} \ S_{d}
1592: \int_{\fracs{2\pi}{L}}^{\fracs{2\pi}{\ell_{s}}} \frac{C \ k^{d-1} dk}
1593: {k^{2}}
1594: \end{equation}
1595: and the denominator is
1596: \begin{equation}
1597: \sum_{\bm{k}} \frac{|\hat{s}_{\bm{k}}|^2}{k^2 + \Pe^2 / L^{2}}
1598: \ \approx \ \left( \frac{L}{2\pi} \right)^{d} \ S_{d}
1599: \int_{\fracs{2\pi}{L}}^{\fracs{2\pi}{\ell_{s}}} \frac{C \ k^{d-1} dk}
1600: {k^{2}+\Pe^2 / L^{2}}.
1601: \end{equation}
1602: In $d=2$ spatial dimensions this implies the bound
1603: \begin{equation}
1604: {\cal E}_{1} \ \lesssim \  
1605: \sqrt{
1606: \log \left[ \frac{L^{2}}{\ell_{s}^{2}}\right]
1607: } \bigg/
1608: \sqrt{
1609: \log \left[ \frac{\Pe^{2}+4 \pi^{2} \frac{L^{2}}{\ell_{s}^{2}}}
1610: {\Pe^{2}+4 \pi^{2}}\right]
1611: } 
1612: \ \ \rightarrow \ \ 1 
1613: \quad \text{as} \ \ \frac{\ell_{s}}{L} \rightarrow 0.
1614: \end{equation}
1615: Hence there can be {\it no} reduction of the small-scale variance beyond that
1616: due to molecular diffusion for such singular source-sink distributions, no
1617: matter what flow strategy is adopted or how energetic the stirring is.
1618: 
1619: It is interesting to note as well that if $\ell_{s}$ is small but finite and
1620: $\Pe \gg \frac{L}{\ell_{s}} \gg 1$, then the 
1621: bound scales classically again.
1622: That is, as $\Pe \rightarrow \infty$,
1623: \begin{equation}
1624: {\cal E}_{1} \ \lesssim \  
1625: \sqrt{
1626: \log \left[ \frac{L^{2}}{\ell_{s}^{2}}\right]
1627: }  \bigg/
1628: \sqrt{
1629: \log \left[ \frac{\Pe^{2}+4 \pi^{2} \frac{L^{2}}{\ell_{s}^{2}}}
1630: {\Pe^{2}+4 \pi^{2}}\right]
1631: } 
1632: \ \rightarrow \ \frac{\ell_{s}}{2 \pi \sqrt{L^{2}-\ell_{s}^{2}}} \, \Pe 
1633: \ \approx \ \frac{\ell_{s}}{2 \pi L} \, \Pe.
1634: \end{equation}
1635: This can be interpreted as the statement
1636: that the largest possible value for the mixing length $\ell_{1}^{(\mathrm{max})} =
1637: {\cal O}(\ell_{s})$ in such cases.
1638: 
1639: 
1640: For such distributions in $d=3$ spatial dimensions,
1641: \begin{equation}
1642: {\cal E}_{1} \ \lesssim \  
1643: \l({
1644: 1 - \frac{\Pe}{\frac{L}{\ell_{s}}-1} \, \left[ 
1645: \arctan\l(\frac{2 \pi L}{\Pe\, \ell_{s}}\r)
1646: - \arctan\l(\frac{2 \pi}{\Pe}\r) \right]
1647: }\r)^{-1/2}
1648: \ \ \rightarrow \ \ 1 
1649: \quad \text{as} \ \ \frac{\ell_{s}}{L} \rightarrow 0.
1650: \end{equation}
1651: Again, there can be no reduction of the small-scale variance beyond that due to
1652: molecular diffusion for this kind of singular source-sink distributions, 
1653: no matter what stirring strategy is adopted.
1654: Furthermore, when $\ell_{s}$ is finite and $\Pe \gg \frac{L}{\ell_{s}} \gg 1$, 
1655: the bound scales classically again: as $\Pe \rightarrow \infty$,
1656: \begin{equation}
1657: {\cal E}_{1} \ \lesssim \  
1658: \l(
1659: {
1660: 1 - \frac{\Pe}{\frac{L}{\ell_{s}}-1} \, \left[ 
1661: \arctan\l(\frac{2 \pi L}{\Pe \ell_{s}}\r)
1662: - \arctan\l(\frac{2 \pi}{\Pe}\r) \right]
1663: }\r)^{-1/2}
1664: \ \rightarrow \ 
1665: \sqrt{
1666: \frac{3}{4\pi^{2}} \, 
1667: \frac{\frac{L}{\ell_{s}}-1}{\frac{L^{3}}{\ell_{s}^{3}}-1}
1668: }
1669:  \, \Pe 
1670: \ \approx \ \frac{\sqrt{3} \, \ell_{s}}{2 \pi L} \, \Pe.
1671: \end{equation}
1672: Not unexpectedly, the largest possible
1673: value for a mixing length for this kind of mollified distribution is
1674: $\ell_{1}^{(\mathrm{max})} = {\cal O}(\ell_{s})$.
1675: 
1676: Classical scaling for ${\cal E}_{0}$ is also eventually recovered for these
1677: regularlized sources and sinks when $\Pe \gg L/\ell_{s}$, albeit with very
1678: different relationships between the smallest source-sink length scale
1679: $\ell_{s}$ and the maximal mixing length $\ell_{0}^{(\mathrm{max})}$.  When
1680: $|\hat{s}(\bm{k})| = C > 0$ for $2\pi/L \le |\bm{k}| < 2\pi/\ell_{s}$ and
1681: $|\hat{s}(\bm{k})| = 0$ for $|\bm{k}| > 2\pi/\ell_{s}$ the numerator of the
1682: bound in (\ref{eq:varlubounda}) is
1683: \begin{equation}
1684: \sum_{\bm{k}} \frac{|\hat{s}_{\bm{k}}|^2}{k^4}
1685: \ \approx \ \left( \frac{L}{2\pi} \right)^{d} \ S_{d}
1686: \int_{\fracs{2\pi}{L}}^{\fracs{2\pi}{\ell_{s}}} \frac{C \ k^{d-1} dk}
1687: {k^{4}}
1688: \end{equation}
1689: and the denominator is
1690: \begin{equation}
1691: \sum_{\bm{k}}\frac {|\hat{s}_{\bm{k}}|^2}{k^4 +k^2 \Pe^2 / L^{2}d}
1692: \ \approx \ \left( \frac{L}{2\pi} \right)^{d} \ S_{d}
1693: \int_{\fracs{2\pi}{L}}^{\fracs{2\pi}{\ell_{s}}} \frac{C \ k^{d-1} dk}
1694: {k^{4}+k^{2}\Pe^2 / L^{2}d}.
1695: \end{equation}
1696: For $d=2$, as $\Pe \rightarrow \infty$,
1697: \begin{equation}
1698: {\cal E}_{0} \ \lesssim \
1699: \sqrt{
1700: \frac{\Pe^{2}}{8 \pi^{2}}
1701: \left( 1 - \frac{\ell_{s}^{2}}{L^{2}} \right)
1702: \frac{1}{\log \left[
1703: \frac{8\pi^{2}+\Pe^{2}}{8\pi^{2}+\ell_{s}^{2}\Pe^{2}/L^{2}}
1704: \right]}
1705: }
1706: \ \ \sim \ \ 
1707: \frac{\Pe}{2 \pi \sqrt{2}} \sqrt{
1708: \frac{\left( 1 - \frac{\ell_{s}^{2}}{L^{2}} \right)}
1709: {\log \frac{L^{2}}{\ell_{s}^{2}}}
1710: }
1711: \ \ \approx \ \  
1712: \frac{\Pe}{4 \pi}
1713: \sqrt{\frac{1}{\log \frac{L}{\ell_{s}}}} \ .
1714: \end{equation}
1715: So in spatial dimension $d=2$ the largest possible mixing length
1716: $\ell_{0}^{(\mathrm{max})} \sim L/\sqrt{\log[L/\ell_{s}]}$.  This length scale
1717: vanishes as $\ell_{s} \rightarrow 0$, but much slower than the actual smallest
1718: scale ${\cal O}(\ell_{s})$.
1719: 
1720: In $d=3$,
1721: \begin{equation}
1722: {\cal E}_{0} \ \lesssim \
1723: \frac{\Pe}{2 \pi \sqrt{2}}\sqrt{
1724: \frac{\ell_{s}}{L}
1725: }
1726: \quad \text{as} \ \Pe \rightarrow \infty
1727: \end{equation}
1728: so $\ell_{0}^{(\mathrm{max})} = {\cal O}(\ell_{s}^{1/2})$.
1729: This vanishes faster than in $2$-$d$ as $\ell_{s} \rightarrow 0$, but
1730: still much slower than $\ell_{s}$.
1731:  
1732: A log-log plot of the bound on ${\cal E}_{0}$ vs $\Pe$ for these cutoff
1733: source-sink distributions is shown in Figure~\ref{dumfig1} for the
1734: $3$-$d$ case with $\ell_{s}/L = 10^{-8}$.  The anomalous $\Pe^{1/2}$ scaling
1735: persists over the range $1 \lesssim \Pe \lesssim L/\ell_{s}$ and classical
1736: scaling $\sim \Pe^{1}$ takes over at higher P\'eclet numbers with a ``small''
1737: prefactor $\sim \sqrt{\ell_{s}/L}$.  This small prefactor is a quantitative
1738: indication of the difficulty of any statistically homogeneous and isotropic
1739: flow field to efficiently suppress the scalar variance---beyond the suppression
1740: achieved by molecular diffusion alone---in the presence of small-scale scalar
1741: sources and sinks.
1742: 
1743: %
1744: \begin{figure}
1745: \centerline{\includegraphics[width=.5\textwidth]{figure1.eps}}
1746: \setlength{\abovecaptionskip}{-5pt}
1747: \caption{${\cal E}_0$ as a function of $\Pe$ for cutoff source-sink
1748: distributions, after~\cite{DoeringThiffeault2006}.}
1749: \label{dumfig1}
1750: \end{figure}
1751: %
1752: 
1753: The scalings for the mixing efficiencies with a $\delta$-function source are 
1754: apparently not all realized by
1755: a simple uniform wind as was the case for smooth monochromatic sources.
1756: We have performed the adiabatic approximation for the bulk variance 
1757: in the case of a uniform wind blowing past a $\delta$-function source
1758: to estimate the mixing efficiencies (see Appendix B).
1759: While ${\cal E}_1$ is necessarily equal to 1 in both $2$-$d$ and $3$-$d$,
1760: we find that the uniform wind gives ${\cal E}_0 \sim \sqrt{\Pe}$ (rather than
1761: the bound $\Pe/\sqrt{\log{\Pe}}$) in $2$-$d$, and 
1762: ${\cal E}_0 \sim \sqrt{\Pe}/\sqrt{\log{\Pe}}$ 
1763: (significantly closer to the bound $\sqrt{\Pe}$) in $3$-$d$.
1764: For the large-scale efficiency, however, the uniform wind produces 
1765: ${\cal E}_{-1} \sim \sqrt{\Pe}$ in both $2$-$d$ and $3$-$d$,
1766: far below its classically scaling upper bound.
1767: 
1768: A similar integral analysis can be carried out for ``fractal'' source-sink
1769: distributions where $|\hat{s}_{\bm{k}}| \sim |\bm{k}|^{-\gamma}$ with $\gamma
1770: \ge 0$.  The scalings of the mixing efficiency bounds are summarized in
1771: Table~\ref{tab:scalingsall}.  For $\gamma > d/2$ all the bounds scale
1772: classically since then $s(\bm{x}) \in L^{2}(\mathbb{T}^d)$.  The
1773: case~$\gamma=0$ is the $\delta$-function studied above.  We observe that the
1774: bounds for ${\cal E}_{1}$ and ${\cal E}_{0}$ can exhibit anomalous scaling
1775: with exponents depending on the fractal nature of the source-sink distribution
1776: as characterized by $\gamma$~\cite{DoeringThiffeault2006}.  Of course if the
1777: fractal scaling of the source-sink distribution persists over a broad but
1778: finite range of wavenumbers, say for $2\pi/L < k < 2\pi/\ell_{s}$, then
1779: classical scaling for the efficiency estimates will again appear for $\Pe \gg
1780: L/\ell_{s}$.
1781: 
1782: \begin{table}
1783: \vspace{.5em}
1784: \caption{Scalings of the bound on the mixing efficiency~${\cal E}_{p}$ as
1785: functions of the source roughness exponent~$\gamma$ of the source in two and
1786: three dimensions.}
1787: \begin{flushleft}
1788: \begin{ruledtabular}
1789: \begin{tabular}{lccc}
1790: $d = 2$ & ${p}=1$ & ${p}=0$ & ${p}=-1$ \\
1791: \hline
1792: $\gamma = 0$       & 1 & $\Perm/(\log\Perm)^{1/2}$ & \Perm \\
1793: $0 < \gamma < 1$ & $\Perm^\gamma$ & \Perm & \Perm \\
1794: %$\gamma = 1/2$     & $\Pe^{1/2}$ & \Pe  & \Pe \\
1795: %$1/2 < \gamma < 1$ & $\Pe^\gamma$ & \Pe & \Pe \\
1796: $\gamma = 1$       & \ \ \ $\Perm/(\log\Perm)^{1/2}$ & \Perm & \Perm \\
1797: $\gamma > 1$       & \Perm & \Perm & \Perm \\[2pt]
1798: %\end{tabular}
1799: %\end{ruledtabular}
1800: %\vspace{1.0em}
1801: %\begin{ruledtabular}
1802: %\begin{tabular}{lccc}
1803: \hline
1804: $d = 3$\\ %& ${p}=1$ & ${p}=0$ & ${p}=-1$ \\
1805: \hline
1806: $\gamma = 0$         & 1 & $\Perm^{1/2}$ & \Perm \\
1807: $0 < \gamma < 1/2$   & 1 & $\Perm^{\gamma+1/2}$ & \Perm \\
1808: $\gamma = 1/2$       & 1 & \quad $\Perm/(\log{\Perm})^{1/2}$ & \Perm \\
1809: $1/2 < \gamma < 3/2$ & $\Perm^{\gamma - 1/2}$ & \Perm & \Perm \\
1810: $\gamma = 3/2$       & \ $\Perm/(\log{\Perm})^{1/2}$ & \Perm & \Perm \\
1811: $\gamma > 3/2$       & \Perm & \Perm & \Perm \\
1812: \end{tabular}
1813: \end{ruledtabular}
1814: \end{flushleft}
1815: \label{tab:scalingsall}
1816: \end{table}
1817: 
1818: 
1819: %===================================================
1820: 
1821: \section{A Single-scale Source Stirred by a Single-scale Flow}
1822: \label{sec:SSSSSF}
1823: 
1824: In the previous sections we have seen that the classical high-P\'eclet number
1825: scalings for the multi-scale mixing efficiencies, ${\cal E}_{p} \sim \Pe$, are
1826: generally upper bounds that may in fact be saturated for particular
1827: source-sink distributions stirred by certain statistically homogeneous and
1828: isotropic flows.  It was shown in Section~\ref{sec:SARSAS} that for some
1829: distribution-valued sources and sinks, some of the mixing efficiencies
1830: necessarily scale anomalously, i.e., ${\cal E}_{p} \lesssim \Pe^{\alpha_{p}}$
1831: with some $\alpha_{p} < 1$.  For those examples the bounds on the mixing
1832: efficiencies at the different scales scale differently: $0 \le \alpha_{1} <
1833: \alpha_{0} < \alpha_{-1} = 1$.  It is not presently known if the anomalously
1834: scaling bounds for distribution-valued sources and sinks are sharp or, if they
1835: are, what kinds of statistically homogeneous and isotropic flows might be
1836: required to realize them.  Those questions remain open, but in this section we
1837: settle the issue of the possibility of realizing distinct scaling exponents
1838: for the mixing efficiencies on different length scales for a smooth
1839: source-sink distribution.
1840: 
1841: The random sine flow, a.k.a.\ the renewing wave flow characterized by a single
1842: length scale, is a popular and convenient test bed for studying---both
1843: analytically and via simulation---a wide variety of stirring and mixing
1844: phenomena~\cite{Pierrehumbert1994,Antonsen1996,Thiffeault2004,Tsang2005}.
1845: Here we consider the simplest single-scale source
1846: \begin{equation}
1847: s(\bm{x}) = \sqrt{2} \, S \sin[2 \pi (x+y)/L]
1848: \end{equation}
1849: stirred by a random sine flow that switches between
1850: \begin{equation}
1851: \bm{u}(\bm{x}) = \hat{\bm{i}} \sqrt{2} \, U \sin[2 \pi y/L + \phi]
1852: \label{Z1}
1853: \end{equation}
1854: and
1855: \begin{equation}
1856: \bm{u}(\bm{x}) = \hat{\bm{j}} \sqrt{2} \, U \sin[2 \pi x/L + \phi]
1857: \label{Z2}
1858: \end{equation}
1859: at time intervals of length $\tau/2$, where the phase $\phi$ is chosen
1860: randomly and uniformly from $[0, 2\pi)$ at each switch.  This is a
1861: statistically homogeneous and isotropic flow, but it is not the maximally
1862: efficient flow for this source-sink distribution; the optimal flow is the
1863: spatially uniform flow constructed in Section~\ref{sec:saturation}.  We have
1864: performed high resolution direct numerical simulations of this system in $d=2$
1865: spatial dimensions that reveal that the multi-scale mixing efficiencies may
1866: have distinct scaling exponents.  Additionally, for this example we are able
1867: to compute the exponents theoretically in the context of the adiabatic
1868: averaging approximation utilized in Section~\ref{sec:saturation}.
1869: 
1870: Figure~\ref{dumfig2} shows the results of the direct numerical simulation for
1871: $p=1,0,-1$, each plotted along with the classically-scaling upper bounds from
1872: Section~\ref{sec:BOUNDS}.  The P\'eclet number in these simulations was varied
1873: by holding $U=1$ fixed in a box of side length $L=1$ 
1874: and changing the molecular diffusivity $\kappa$.  
1875: The computations were carried out with switching time $\tau=1$.  
1876: The details of the numerical simulations can be found
1877: in~\cite{Thiffeault2004}.
1878: %
1879: \begin{figure}
1880: \centerline{\includegraphics[width=\textwidth]{figure2.eps}}
1881: \setlength{\abovecaptionskip}{-30pt}
1882: \caption{ Mixing efficiencies ${\cal E}_p$ as a function of $\Pe$ for (a)
1883: $p=1$, (b) $p=0$, and (c) $p=-1$ for the random sine flow with source $\sim
1884: \sin k_s(x+y)$.  The solid lines are the upper bounds for this source from
1885: Section~\ref{sec:BOUNDS} and the dashed lines are the data from the direct
1886: numerical simulations with $U$, $L$ and $\tau$ fixed.}
1887: \label{dumfig2}
1888: \end{figure}
1889: %
1890: Writing ${\cal E}_{p} \sim \Pe^{\alpha_{p}}$ as $\Pe \rightarrow \infty$,
1891: it is clear that the data are 
1892: consistent with $\alpha_{0} < \alpha_{1} < 1$ and $\alpha_{-1} \approx 1$.
1893: 
1894: These simulations establish two important facts:
1895: (1) that the mixing efficiencies at different length scales generally scale differently at
1896: high P\'eclet numbers, and (2) that anomalous subclassical scaling can easily be 
1897: realized by simple and ``reasonable" flows.
1898: 
1899: We can evaluate the observed scaling exponents $\alpha_{p}$ theoretically for
1900: this particular system by appealing to a quasi-static adiabatic approximation
1901: introduced in Section~\ref{sec:saturation} for the optimally mixing uniform
1902: flow.  That is, we solve the time-independent problem for the scalar
1903: distribution under the influence of steady flows of the form (\ref{Z1}) and
1904: (\ref{Z2}), compute the variances, and average over the flow configurations.
1905: The physical justification for this approximation comes from examining
1906: the results of the direct numerical simulations: it is observed that the
1907: relevant steady-state configuration is quickly approached in the time
1908: intervals between the switches of the stirring field.  If we make the
1909: switches at longer and longer intervals, i.e., if we increase $\tau$,
1910: the scalar field is distributed (nearly) according to the static
1911: configurations for most of the time, and the steady-state variances
1912: dominates the time averages.\footnote{This approach is similar in spirit
1913: to ``rapid distortion theory" in turbulence~\cite{Savill1987,Hunt1990}.}
1914: 
1915: The simplest problem in this category is when the velocity field is oriented
1916: parallel to the gradient of the source, so we consider the steady
1917: advection--diffusion equation
1918: \begin{equation}
1919:   \sqrt{2} U \sin \ku y \  \partial_{x} \theta =
1920:   \kappa (\partial_{x}^{2}+\partial_{y}^{2}) \theta + 
1921:   \sqrt{2} S \sin \ks x
1922:   \label{eq:ADsteady}
1923: \end{equation}
1924: where we allow for different length scales in the stirring and the source with
1925: the non-dimensional number $r=\ku/\ks$ gauging the relative amount of shear in
1926: the flow.  This is not exactly the static problem corresponding to the
1927: dynamic simulations where the flow is always at a $45\degree$ angle from the
1928: source-sink alignment, but it turns out that the scaling exponents
1929: $\alpha_{p}$ are the same for that case and for the $90\degree$ alignment in
1930: (\ref{eq:ADsteady}).  For illustrative purposes it is convenient to explain
1931: the most elementary case (\ref{eq:ADsteady}) in detail here
1932: and relegate details of the $45\degree$ problem to Appendix C.
1933: 
1934: The solution to~\eqref{eq:ADsteady} takes the form
1935: \begin{equation}
1936:   \theta(\bm{x}) = f(y)\sin(\ks x) + g(y) \cos(\ks x)
1937: \end{equation}
1938: where the functions $f$ and $g$ are periodic on $y\in[-\pi/\ku,\pi/\ku]$ 
1939: and satisfy the system of ODEs
1940: \begin{subequations}
1941: \begin{align}
1942:   -\sqrt{2}U\ks \sin(\ku y)\, g(y)
1943:   &= \kappa \left [ -\ks^2 + \frac{d^2}{dy^2}
1944:     \right] f(y) + \sqrt{2} S\,,\label{54a}\\
1945:   \sqrt{2}U\ks \sin(\ku y) f(y) &= \kappa \left [ -\ks^2 + \frac{d^2}{dy^2}
1946:     \right] g(y)\,.\label{54b}
1947: \end{align}
1948: \label{eq:fgeqns}%
1949: \end{subequations}%
1950: From~\eqref{54a} we deduce that $g(y)$ is an odd function of~$y$ and~$f(y)$ is
1951: an even function of~$y$, compatible with~\eqref{54b} which says
1952: that~$f$ and~$g$ have opposite parity.
1953: We can thus infer boundary conditions on the reduced domain $[0,\pi/\ku]$:
1954: \begin{equation}
1955:   g(0)=0=g(\pi/\ku),\quad f'(0)=0=f'(\pi/\ku).
1956: \end{equation}
1957: 
1958: We are interested in the high-$\Pe$ behavior of the solution
1959: to~\eqref{eq:ADsteady}.  In Figure~\ref{dumfig3} we show a grayscale plot of
1960: the scalar field at several values of the P\'eclet number.  As is evident,
1961: internal layers develop along the lines of maximum shear around $y = n
1962: \pi/\ku$ for $n=0, \pm1, \pm2, \dots$.  Away from these lines the flow
1963: directly blows source regions onto sink regions and vice versa which, as we
1964: have seen in Section~\ref{sec:saturation}, is the most efficient stirring
1965: to reduce variance at all scales.  Even though the regions of high shear
1966: effectively stretch material lines, the mixing process is apparently
1967: frustrated by the constant replishment of the scalar variation by the steady
1968: sources and sinks and the variance is dominated by fluctuations concentrated
1969: in shear layers.
1970: 
1971: %
1972: \begin{figure}
1973: \centerline{\includegraphics[width=20.0cm]{figure3.eps}}
1974: \setlength{\abovecaptionskip}{-25pt}
1975: \caption{The scalar field stirred by a steady sine flow for (a) $\Pe$ = 0, (b) $\Pe$ = 100,
1976:  (c) $\Pe$ = 1000.}
1977: \label{dumfig3}
1978: \end{figure}
1979: %
1980: 
1981: We develop an asymptotic singular perturbation internal layer approach for the
1982: limit $\Pe \gg 1$ with~$r = {\cal O}(1)$ fixed.  Upon rescaling $\tilde{y}=\ks
1983: y$, $\hat{f}=fU\ks/S$, $\hat{g}=gU\ks/S$, $r=\ku/\ks$, and introducing the
1984: slightly re-scaled P\'eclet number $\Pec=\sqrt{2}U/\kappa \ks = \sqrt{2}
1985: \Pe/\ks L$ in Eqs.~\eqref{eq:fgeqns}, we obtain the non-dimensional ODEs
1986: \begin{subequations}
1987: \begin{align}
1988:   \frac{1}{\Pec}  \left[ -1 + \frac{d^2}{d \tilde{y}^2} \right]
1989:   \hat{f}(\tilde{y})+1 &= -\sin(r\tilde y) \hat{g}(\tilde{y}),\label{55a}\\
1990:   \frac{1}{\Pec}  \left[ -1 + \frac{d^2}{d \tilde{y}^2} \right]
1991:   \hat{g}(\tilde{y}) &= \sin(r\tilde y) \hat{f}(\tilde{y}).\label{55b}
1992: \end{align}
1993: \label{eq:fgeqnsscaled}%
1994: \end{subequations}%
1995: Proceeding as usual we construct inner and outer solutions.  The outer
1996: solution, valid away from the internal layers, is obtained by expanding in
1997: powers of $\Pec^{-1}$:
1998: \begin{equation}
1999:   \hat{f}_{\mathrm{outer}}=\sum_{n=0}^{\infty}  \Pec^{-n}
2000:   \hat{f}_{n},\qquad\hat{g}_{\mathrm{outer}}=\sum_{n=0}^{\infty} \Pec^{-n}
2001:   \hat{g}_{n}.
2002: \end{equation}
2003: To leading order the solution to~\eqref{eq:fgeqnsscaled} in the outer region is
2004: \begin{equation}
2005:   \hat{f}_{\mathrm{outer}} \sim 0,\qquad 
2006:   \hat{g}_{\mathrm{outer}} \sim -\frac{1}{ \sin(r \tilde{y})}\,.
2007:   \label{eq:outersol}
2008: \end{equation}
2009: 
2010: For the internal layer we expand in a small parameter~$\epsilon$ as
2011: \begin{equation}
2012:   \hat{f}_{\mathrm{inner}}= \sum_{n=-1}^{\infty}  \epsilon^n
2013:   \hat{f}_{n},\qquad \hat{g}_{\mathrm{inner}}=
2014:   \sum_{n=-1}^{\infty}  \epsilon^n \hat{g}_{n}\,,
2015:   \label{eq:innersol}
2016: \end{equation}
2017: so that both~$\hat{f}_{\mathrm{inner}}$ and~$\hat{g}_{\mathrm{inner}}$ are
2018: ${\cal O}(\epsilon^{-1})$ as~\hbox{$\epsilon\rightarrow 0$}.  The internal
2019: layer scaling is determined by a dominant balance argument: we choose
2020: $\epsilon= \Pec^{-1/3}$ and rescale~$\tilde{y} = \epsilon\, \eta$ to achieve a
2021: self-consistent scaling of the leading order terms.  When these scalings
2022: and~\eqref{eq:innersol} are inserted into~\eqref{eq:fgeqnsscaled}, the problem
2023: to solve at leading order ${\cal{O}}(1/\epsilon)$ is
2024: \begin{equation}
2025:   \frac{d^2 \hat{f}_{-1}}{d \eta^2} + r \eta \hat{g}_{-1} +1=0,
2026:   \qquad\frac{d^2 \hat{g}_{-1}}{d \eta^2} - r \eta \hat{f}_{-1} =0.
2027:   \label{eq:fginner}
2028: \end{equation}
2029: Then letting $\xi=r^{1/3} \eta$, $F=r^{2/3} \hat{f}_{-1}$, and
2030: $G=r^{2/3}\hat{g}_{-1}$, \eqref{eq:fginner} simplifies to
2031: the inner layer equations
2032: \begin{equation}
2033:   F^{''} + \xi G +1 =0,\qquad G^{''} - \xi F = 0,
2034:   \label{510}
2035: \end{equation}
2036: with boundary conditions
2037: %\begin{equation}
2038:   $F'(0)=0$ and $G(0)=0$.
2039: %\end{equation}
2040: The other boundary conditions come from the requirement of matching to the
2041: outer solution~\eqref{eq:outersol}: $F(\xi) \rightarrow 0$ and $ G(\xi)
2042: \rightarrow -1/\xi$ as $\xi \rightarrow \infty$.
2043: The system (\ref{510}) can be cast as a complex Airy equation for $F+iG$, 
2044: but we resort instead to a numerical shooting method to obtain the solution.  
2045: 
2046: The solution of the internal layer equations is obtained by shooting backward 
2047: (which is the more stable evolution direction) from $\xi \rightarrow \infty$.
2048: The large-$\xi$ asymptotic behavior deduced from \eqref{510} is
2049: \begin{equation}
2050:   F \approx -\frac{2}{\xi^4} + \frac{a}{\xi^{10}}\,,\qquad
2051:   G \approx -\frac{1}{\xi} + \frac{b}{\xi^7}\,,
2052: \end{equation}
2053: where in practice $a$ and $b$ are tuned numerically to produce a solution that
2054: satisfies the boundary conditions at $\xi=0$.  This is easily accomplished,
2055: and in Figure~\ref{dumfig4} we compare the internal layer solution against the
2056: exact (numerical) solution of the full advection--diffusion equation
2057: (\ref{eq:ADsteady}) at $\Pe=1000$ where the small parameter $\epsilon \approx
2058: .2$.  This agreement confirms that the leading terms
2059: \begin{equation}
2060:   \hat{f}_{\mathrm{inner}} = \frac{r^{-2/3}}{\epsilon}\,F(\xi),
2061:   \qquad
2062:   \hat{g}_{\mathrm{inner}} = \frac{r^{-2/3}}{\epsilon}\,G(\xi).
2063: \end{equation}
2064: do indeed accurately capture the asymptotic behavior.
2065: %
2066: \begin{figure}
2067: \centerline{\includegraphics[width=15cm]{figure4.eps}}
2068: \setlength{\abovecaptionskip}{-35pt}
2069: \caption{Comparison of the direct numerical solution (solid) and the internal
2070: layer solution (dashed) for \hbox{$\Pe$ = 1000}.}
2071: \label{dumfig4}
2072: \end{figure}
2073: %
2074: 
2075: The uniform asymptotic solution to the coupled ODEs are, to leading order, the
2076: composites of the inner and outer solutions.  Recovering all the scalings and
2077: letting $\delta=\epsilon/r^{1/3} \ks$, we define
2078: \begin{subequations}
2079: \begin{align}
2080:   f_{\mathrm{comp}}(y) \ &= \ \frac{S}{U\ks} \,\hat{f}_{\mathrm{inner}}(y)
2081:   %= \frac{S}{U\ks}
2082:   %\frac{r^{-2/3}}{\epsilon}F \left( \frac{r^{1/3} \ks}{\epsilon} y \right)
2083:   \ = \frac{S}{U\ks}\, \frac{1}{\ku \delta} F \left(\frac{y}{\delta}\right),\\
2084:   g_{\mathrm{comp}}(y) \ &= \ \frac{S}{U\ks} \,
2085:   \ku y \times \hat{g}_{\mathrm{inner}}(y) \times \hat{g}_{\mathrm{outer}}(y)
2086:   %=\frac{S}{U\ks} \frac{r^{-2/3}}{\epsilon}G
2087:   %\left( \frac{r^{1/3} \ks}{\epsilon} y \right) \frac{\ku y}{\sin(\ku y)}
2088:   %\\
2089:   \ = \frac{S}{U\ks}\, \frac{1}{\ku \delta}\,
2090:   \frac{\ku y}{\sin(\ku y)}\,G\left(\frac{y}{\delta}\right).
2091: \end{align}
2092: \label{eq:fginternalsol}
2093: \end{subequations}
2094: Armed with the approximate solutions~\eqref{eq:fginternalsol} we can compute the multi-scale
2095: mixing measures $\langle |\nabla^{p} \theta |^2 \rangle$ for $p=0,1,-1$.
2096: 
2097: The variance is
2098: \begin{equation}
2099:   \langle \theta^2 \rangle = \frac{1}{2} \left( \langle f^2 \rangle + \langle
2100:   g^2 \rangle \right) =\frac{1}{2} \frac{4\ku}{\pi}
2101:   \left( \int_0^{\pi/4\ku} f(y)^2 dy +
2102:   \int_0^{\pi/4\ku} g(y)^2 dy \right)
2103: \end{equation}
2104: where we use the symmetry of the solution to carry out the integral over only
2105: a quarter-period.  
2106: Letting $\eta=y/\delta$ and replacing $f$ and $g$ by $f_{\mathrm{comp}}$ and $g_{\mathrm{comp}}$,
2107: \begin{equation}
2108:   \langle \theta^2 \rangle \ \sim \ \frac{1}{\pi} 
2109:   \frac{S^2}{U^2\ks^2}\frac{1}{\ku \delta}
2110:   \left( \int_0^{\fracs{\pi}{2\ku \delta}} F(\eta)^2
2111:   d\eta+\int_0^{\fracs{\pi}{2\ku \delta}} G(\eta)^2 \frac{\ku^2 \eta^2
2112:   \delta^2}{\sin^2(\ku \eta \delta)} d\eta \right).
2113: \end{equation}
2114: Here we are interested in the scaling as $\delta\rightarrow 0$ so we are
2115: justified in replacing the upper limit of the integral of $F(\eta)^2$ by
2116: infinity.  Slightly more care must be taken with the integral involving
2117: $G(\eta)^2$: because $\ku^2 \eta^2 \delta^2/\sin^2(\ku \eta \delta)$ is
2118: uniformly bounded by $0$ and $\pi/2$ we can appeal to Lebesgue's dominated
2119: convergence theorem to deduce
2120: \begin{equation}
2121:   \langle \theta^2 \rangle \ \sim \ \frac{1}{\pi}  \frac{S^2}{U^2\ks^2}
2122:   \frac{1}{\ku\delta}
2123: \left( \int_0^\infty F(\eta)^2 d \eta + \int_0^\infty G(\eta)^2 d \eta\right).
2124: \end{equation}
2125: The unstirred variance is $\langle \theta_0^2 \rangle=S^2/\kappa^2 \ks^4$, 
2126: so the mixing efficiency is
2127: \begin{equation}
2128:   {\cal E}_{0} \ \sim \ \sqrt{\frac{\pi}{2}} \,
2129:   \left(\int_0^\infty F(\eta)^2 d \eta + 
2130:   \int_0^\infty G(\eta)^2 d \eta \right)^{-1/2} \, 
2131:    r^{1/3} \, \Pec^{5/6}.
2132:    \label{A0}
2133: \end{equation}
2134: 
2135: Figure~\ref{dumfig5} shows ${\cal E}_{0}$ as a function of $\Pe$ from the
2136: numerical solution of the steady advection--diffusion equation
2137: (\ref{eq:ADsteady}).  The data accurately match the $\Pe^{5/6}$ scaling and
2138: using the prefactor calculated from (\ref{A0}) there is only a 5\% discrepancy
2139: between the internal layer solution and that calculated from the direct
2140: numerical simulation at $\Pe=1000$ with $r=1$.  Remarkably, this $\Pe^{5/6}$
2141: scaling also fits the data for the time-dependent stirring of the tilted
2142: source shown in Figure~\ref{dumfig2}.
2143: %
2144: \begin{figure}
2145: \centerline{\includegraphics[width=.5\textwidth]{figure5.eps}}
2146: \setlength{\abovecaptionskip}{-3pt}
2147: \caption{ Mixing efficiencies ${\cal E}_{p}$ for $p=1$ (x), $p=0$ (*) and
2148: $p=-1$ (+) for the steady sine flow.  The solid lines are the scaling
2149: predictions of the asymptotic analysis and the discrete data are from the
2150: numerical solution of the steady advection--diffusion equation
2151: (\ref{eq:ADsteady}).  }
2152: \label{dumfig5}%
2153: \end{figure}%
2154: %
2155: 
2156: The gradient variance in the steady solution is given by
2157: \begin{equation}
2158:   \langle | \nabla \theta|^2 \big\rangle = \frac{\ks^2}{2} \left[ \langle f^2
2159:   \rangle + \langle g^2 \rangle \right] +\frac{1}{2} \left[ \langle (f')^2
2160:   \rangle + \langle (g')^2 \rangle \right].
2161:   \label{eq:gradvarinternal}
2162: \end{equation}
2163: The $\langle f^2 \rangle$ and $\langle g^2 \rangle$ terms were computed above
2164: in the context of ${\cal E}_{0}$, so we focus on the derivative terms.
2165: Inserting the composite asymptotic solution~\eqref{eq:fginternalsol} we obtain
2166: \begin{multline*}
2167:   \frac{1}{2} \left[ \langle (f')^2 \rangle +  \langle (g')^2 \rangle
2168:     \right]  \sim \left \langle \left( \frac{S}{U\ks} \frac{1}{\ku \delta^2}
2169:   F'(y/\delta) \right)^2  \right \rangle + \\
2170:   \left \langle \left(\frac{S}{U\ks} \frac{1}{\ku \delta^2} G'(y/\delta)
2171:     \frac{\ku y}{\sin(py)} + \frac{S}{U\ks} \frac{1}{\ku\delta} G(y/\delta)
2172:     \left[\frac{\ku y}{\sin(\ku y)}\right]' \right)^2 \right \rangle.
2173: \end{multline*}
2174: The last term is smaller (${\cal O}(\delta^{-1})$) than the immediately
2175: preceeding term (${\cal O}(\delta^{-2})$) and may thus be neglected.  The
2176: first two terms in~\eqref{eq:gradvarinternal} are also ${\cal
2177: O}(\delta^{-1})$, so they may be neglected as well.  Hence
2178: \begin{equation*}
2179:   \frac{1}{2} \left[ \langle (f')^2 \rangle + \langle (g')^2 \rangle \right]
2180:   \sim \frac{1}{\pi} \frac{S^2}{U^2 \ks^2} \frac{1}{\ku\delta^3}  \left[
2181:   \int^{\fracs{\pi}{2\ku\delta}}_0 F'(\eta)^2 d\eta +
2182:   \int^{\fracs{\pi}{2\ku\delta}}_0 G'(\eta)^2 \frac{\ku^2 \eta^2
2183:   \delta^2}{\sin^2(\ku\eta\delta)}d\eta \right]
2184: \end{equation*}
2185: and by the same arguments as for the variance above we deduce
2186: \begin{equation}
2187:   \langle | \nabla \theta|^2 \big\rangle \sim \frac{1}{\pi} \left(
2188:   \int_0^\infty F'(\eta)^{2} d \eta + \int_0^\infty G'(\eta)^{2} d \eta\right)
2189:   \frac{S^2}{U^2\ks^2} \frac{1}{\ku \delta^3}\,.
2190: \end{equation}
2191: Using $\langle| \nabla \theta_0|^2 \rangle=S^2/\kappa^2 \ks^2$,
2192: we deduce that the mixing efficiency is
2193: \begin{equation}
2194:   {\cal E}_{1} \ \sim \  \sqrt{\frac{\pi}{2}} \,
2195:   \left( \int_0^\infty F'(\eta)^2 d \eta + \int_0^\infty
2196:   G'(\eta)^2 d \eta \right)^{-1/2} \,
2197:   \Pec^{1/2}.
2198:   \label{A1}
2199: \end{equation}
2200: 
2201: Figure~\ref{dumfig5} also shows the scaling of ${\cal E}_{1}$ from the 
2202: numerical solution of (\ref{eq:ADsteady}). 
2203: The internal layer asymptotic approximation in (\ref{A1}) and the numerical
2204: solutions differ by approximately 1\% at $\Pe$ = 1000.  
2205: The (lack of) scaling in $r$ was also confirmed numerically. 
2206: Interestingly, for this problem stirring at ever smaller scales does {\em not} 
2207: further enhance the mixing efficiency on small scales. 
2208: This is apparently because the decrease in gradient variance due to stirring 
2209: on small scales is compensated by the increase in 
2210: gradient variance in the internal layer.
2211: 
2212: 
2213: The inverse-gradient variance,
2214: \begin{equation}
2215:   \langle | \nabla^{-1} \theta|^2 \rangle \ = \ \langle | \nabla^{-1}(f(y)
2216:   \sin(\ks x) + g(y) \cos(\ks x))|^2 \rangle,
2217: \end{equation}
2218: requires a slightly more subtle analysis.
2219: Expanding $f(y)$ and $g(y)$ in Fourier series,
2220: \begin{equation}
2221:   f(y)=\sum _{n=0}^{\infty} f_{n} \cos(n\ku y) , \quad \quad
2222:   g(y)=\sum _{n=1}^{\infty} g_{n} \sin(n\ku y)  
2223: \end{equation}
2224: the inverse gradient variance is
2225: \begin{equation}
2226:   \langle | \nabla^{-1} \theta|^2 \rangle \ = \ 
2227:   \frac{f_{0}^{2}}{2 \ks^{2}} + 
2228:   \frac{1}{4} \sum _{n=1}^{\infty} \frac{f_{n}^{2}+g_{n}^{2}}
2229:   {\ks^{2} + n^{2} \ku^{2}}.
2230: \end{equation}
2231: We know that ${\cal E}_{-1} \lesssim \Pe$, so if we could establish a {\it lower} bound with 
2232: the same scaling then we could conclude that ${\cal E}_{-1}$ scales classically $\sim \Pe$.
2233: Hence we focus on deducing an asymptotic {\it upper} bound on 
2234: $\langle | \nabla^{-1} \theta|^2 \rangle$.
2235: As will be seen, however, the best we can do is assert the classical scaling 
2236: with a logarithmic correction.
2237: 
2238: Toward that end, using the fact that $\sum_{n=1}^{\infty} n^{-2} = \pi^{2}/6$,
2239: \begin{equation}
2240:   \langle | \nabla^{-1} \theta|^2 \rangle \ \le \ 
2241:   \frac{f_{0}^{2}}{2 \ks^{2}} \ + \ 
2242:   \frac{\pi^{2}}{24 \ku^{2}} \, \sup_{n \ge 1} (f_{n}^{2}+g_{n}^{2}).
2243: \end{equation}
2244: Then the key is to note that the Fourier coefficients $f_{n}$ are all 
2245: ${\cal O}(S/U\ks)$---without any further factors of $\delta$ appearing---while 
2246: the $g_{n}$ are ${\cal O}(S/U\ks \times \log{\frac{1}{\delta}})$.
2247: Indeed,
2248: \begin{equation}
2249:   |f_{n}| \ = \ \left| \frac{(2-\delta_{n,0})\ku}{\pi}
2250:   \int_{-\pi/2\ku}^{\pi/2\ku} f(y) \cos(n \ku y) dy \, \right| 
2251:     \ \lesssim \ \frac{2(2-\delta_{n,0})S}{\pi U\ks}
2252:     \int_{0}^{\infty} |F(\eta)|  \, d\eta
2253: \end{equation}
2254: while %**
2255: \begin{eqnarray}
2256:   |g_{n}| \ &=& \left| \frac{2\ku}{\pi}
2257:   \int_{-\pi/2\ku}^{\pi/2\ku} g(y) \sin(n \ku y) dy \, \right|
2258:     \nonumber \\
2259:       &\sim& \frac{2S}{\pi U \ks} \left| \int_{-\pi/2\ku\delta}^{\pi/2\ku\delta} 
2260:     \frac{\ku \delta \, \eta}{\sin(\ku \delta \, \eta)}
2261:     G(\eta) \sin(n \ku \delta \, \eta) d\eta \, \right|
2262:     \nonumber \\
2263:       &\le& \frac{2 S}{U\ks} \int_{0}^{\pi/2\ku\delta} 
2264:        | G(\eta)| \, d\eta \,  \ \sim \
2265:        \frac{2S}{U\ks}\log\left[ \frac{1}{\ku \delta}\right].
2266: \end{eqnarray}
2267: In the last step above we used the fact that $|G(\eta)| \sim \eta^{-1}$ as 
2268: $\eta \rightarrow \infty$.
2269: 
2270: The unstirred large-scale variance is $\langle| \nabla^{-1} \theta_0|^2
2271: \rangle=S^2/\kappa^2 \ks^6$, so we deduce the large-scale 
2272: mixing efficiency obeys
2273: \begin{equation}
2274:   {\cal E}_{-1} \ \ge \ C \, \frac{\Pe}{\ks L}
2275:   \times \frac{1}{\sqrt{{1 + C' \left(\frac{\log{\Pe}}{r}\right)^{2}}}}
2276: \end{equation}
2277: where $C$ and $C'$ are absolute constants.  The numerical solutions of the
2278: steady advection--diffusion equation shown in Figure~\ref{dumfig5} confirms
2279: this classical scaling; the logarithmic term is not visible for this range of
2280: $\Pe$.  Again, this is consistent with the dynamic data for ${\cal E}_{-1}$ in
2281: Figure~\ref{dumfig2}.
2282: %*%*%*
2283: 
2284: 
2285: 
2286: %===================================================
2287: 
2288: %\newpage
2289: 
2290: \section{Summary and Discussion}
2291: \label{sec:SandD}
2292: 
2293: \noindent
2294: 
2295: The suppression of bulk variance of a scalar field is a natural gauge of the
2296: efficiency of stirring, and this notion allows for the examination of the
2297: effect of a flow on various length scales.  We have quantified the influence
2298: of stirring on weighted bulk variances in terms of nondimensional mixing
2299: efficiencies ${\cal E}_p$ that characterize fluctuations on relatively small
2300: ($p=1$), intermediate ($p=0$) and large ($p=-1$) length scales.  We studied
2301: these mixing efficiencies for statistically stationary, homogeneous and
2302: isotropic flow fields stirring scalars sustained by a variety of steady but
2303: spatially inhomogeneous scalar sources and sinks on periodic domains.  We
2304: reach a number of conclusions:
2305: \begin{itemize}
2306: 
2307: \item{Very generally, the mixing efficiencies ${\cal E}_p$ are bounded from
2308: below $\sim \Pe^0$ at high P\'eclet numbers---indicating ineffective
2309: stirring---and from above $\sim \Pe^1$---the classical scaling anticipated by
2310: the simplest eddy diffusion theory.  Classical scaling of the efficiencies
2311: corresponds to the existence of residual supression of variance in the
2312: vanishing diffusion ($\kappa \rightarrow 0$) limit.}
2313: \item{Source, sink and statistically stationary homogeneous
2314: and isotropic flow combinations exist that realize both the upper and lower
2315: bounds' scaling with respect to $\Pe$---and in some cases the prefactors
2316: as well---showing that the analysis is sharp at the most general level.}
2317: 
2318: \item{The small and intermediate scale mixing efficiencies
2319: ${\cal E}_1$ and ${\cal E}_0$ are ultimately limited
2320: by length scales in the source-sink distributions.
2321: That is, even if the efficiencies scale classically the inferred mixing lengths
2322: $\sim \kappa \Pe/U$ are limited by length scales defined by the
2323: sources and sinks rather than by the spectrum of scales in the flow.
2324: On the other hand the upper bound analysis does not prevent small scales 
2325: in the flow field from enhancing suppression of the variance at large scales, 
2326: i.e., ${\cal E}_{-1}$, to an unlimited degree.}
2327: 
2328: \item{Sufficiently ``rough'' scalar sources and sinks, in particular
2329: non-square-integrable distributions, necessarily produce sub-classical scaling
2330: of ${\cal E}_0$ and/or ${\cal E}_1$ at high P\'eclet numbers.  In such cases
2331: there is no residual suppression of variance in the vanishing diffusion limit
2332: for any finite mean kinetic energy statistically stationary homogeneous and
2333: isotropic stirring.}
2334: 
2335: \item{The mixing efficiencies at various length scales need not scale the same
2336: as $\Pe \rightarrow \infty$.  This was illustrated explicitly by the example of a monochromatic
2337: source distribution stirred by the random sine flow.  We found by direct
2338: numerical simulation and an adiabatic asymptotic internal layer analysis that
2339: for small scales ${\cal E}_1 \sim \Pe^{1/2}$, for intermediate scales ${\cal
2340: E}_0 \sim \Pe^{5/6}$, and the large length scale efficiency ${\cal E}_{-1}$
2341: exhibits near-classical scaling $\sim \Pe/\log{\Pe}$.}
2342: \end{itemize}
2343: 
2344: The analysis, simulations and modeling presented in this paper have addressed
2345: some fundamental questions regarding suppression of the long time averaged
2346: bulk scalar variance via stirring by incompressible and statistically
2347: homogeneous and isotropic flow fields, but compelling challenges remain for
2348: future study.  Among the unsolved problems open questions are:
2349: \begin{itemize}
2350: \item{Can the anomalous mixing efficiency bounds for measure-valued or fractal
2351: sources be achieved by any statistically stationary homogeneous and isotropic
2352: flow?  This is a difficult problem for direct numerical simulations of the
2353: advection--diffusion equation because of the small spatial scales that need to
2354: be resolved.  An alternate simulation approach is particle tracking, by
2355: following the evolution of many discrete particles moving with the flow and
2356: diffusing and approximating the continuous scalar concentration on an
2357: appropriately coarse grained level.}
2358: 
2359: \item{Can the small-flow-scale enhancement of the large-scale mixing
2360: efficiency ${\cal E}_{-1}$, as suggested by the upper bound in
2361: (\ref{eq:igvarlubound}), be realized?  It makes sense that stirring may be
2362: capable of suppressing large-scale variance, even for negligible molecular
2363: diffusion, by transferring scalar inhomogeneities from large length scales to small
2364: length scales via kinetic stretching and folding mechanisms.  The upper bound
2365: provides a quantitative estimate of this effect and it will be interesting to
2366: see to what extent it is an accurate estimate.}
2367: 
2368: \item{What is the mixing efficiency of ``real'' stationary homogeneous and
2369: isotropic turbulence in two and three dimensions?  The bounds derived here all
2370: apply to these flows, and the inevitable limitations on the mixing efficiency
2371: imposed by the source and sink structure apply. Turbulent mixing by a fluid
2372: with viscosity $\nu$ is often characterized by a Reynolds number
2373: $\text{Re}=UL/\nu$ and a Schmidt number $\text{Sc}=\nu/\kappa$.  While the
2374: mixing efficiencies ${\cal E}_p$ are studied here as functions of
2375: $\Pe=\text{Re} \times \text{Sc}$, the separate Reynolds and Schmidt number
2376: dependences are of interest, too.  These questions can be investigated
2377: theoretically and via direct numerical simulations in the periodic domain
2378: setting utilized here.}
2379: 
2380: \item{The uniform-wind optimal solution presented in
2381: Section~\ref{sec:saturation} is only appropriate for a one-dimensional source
2382: distribution in a domain with periodic boundary conditions.  
2383: It is optimal by virtue of always having a component blowing
2384: along the source gradient, so no time and little energy
2385: is wasted blowing along lines of constant source.
2386: If the source has a more complicated structure, then it is not generally possible to
2387: achieve such an efficient flow while preserving incompressibility.  Physical
2388: boundaries or a different golbal topology (e.g. a spherical domain as relevant to
2389: geophysical applications) will likewise invalidate the uniform-wind solution.
2390: The general question is then: For a given distribution of sources and sinks in
2391: a given domain, what is the optimal stirring strategy to achieve the greatest
2392: reduction of bulk variance?  This question has obvious relevance to
2393: engineering applications.  While we have answered it in the simplest setting
2394: of a single mode source-sink distribution on a periodic domain, it is clearly
2395: a complex problem in general.  It is not unlikely, however, that intuition
2396: gained from the study of idealized models may contribute to the development of
2397: valuable intuition for such systems.}
2398: 
2399: \item{Bounded domains with rigid walls are more appropriate for some
2400: engineering applications.  Then the scalar sources and/or sinks can be
2401: implemented by boundary conditions rather than as body sources and sinks
2402: \cite{Balmforth2003}.  We can also imagine situations where fluid and/or
2403: imposed scalar fluxes at boundaries contribute to the scalar variance in the
2404: bulk.  How this will affect the efficiency scalings is an open question.}
2405: 
2406: \item{We have seen that very generally ${\cal E}_1 \ge 1$, but have only
2407: established a similar lower bound on ${\cal E}_0$ and ${\cal E}_{-1}$ for the
2408: simplest case of monchromatic sources.  It is an open question whether or not
2409: there exist source, sink and flow (even steady flow) combinations where ${\cal
2410: E}_0$ and/or ${\cal E}_{-1} < 1$.  It is worthwhile noting in this context
2411: that in all cases $\langle \theta^2 \rangle^2 \le \langle |\nabla \theta|^2
2412: \rangle \langle |\nabla^{-1}\theta|^2 \rangle$.  Any stirring decreases the
2413: mean bulk gradient variance, but this suggests that if it does not decrease
2414: the variance significantly then it would have to {\it increase} the inverse
2415: gradient variance.  Hence it may not be surprising to find flows that can
2416: amplify large-scale fluctuations.}
2417: 
2418: \item{A related problem of interest is the advection--diffusion of a passive
2419: scalar $\theta$ with decay rate $\zeta$ sustained by a body
2420: source with evolution described by
2421: \begin{equation}
2422:   \frac{\partial \theta}{\partial t} + \bm{u} \cdot \nabla \theta= \kappa
2423:   \lapl \theta +s(\bm{x}) - \zeta\theta.
2424: \end{equation}
2425: In applications $\zeta$ may have an interpretation as a chemical reaction rate
2426: or a radiative relaxation rate relevant in meteorology or for other
2427: geophysical flows on the sphere.  A linear amplitude damping term in the
2428: advection--diffusion equation introduces new features such as competition
2429: between scalar decay and mixing to supress steady-state variances
2430: \cite{ShawGFD2005}.}
2431: 
2432: \item{Finally, all the questions above can be posed for time dependent 
2433: sources and sinks.}
2434: \end{itemize}
2435: 
2436: %\noindent
2437: %``It has been recognized that $L^p$ norms of the passive scalar fail to quantify
2438: %the \emph{stirring efficiency} of a mixing process because they are
2439: %insensitive to small-scale structures \cite{Danckwerts1952}.''
2440: 
2441: 
2442: 
2443: \section*{Acknowledgements}
2444: 
2445: We are grateful for stimulating discussions with Paola Cessi, Bruno Eckhardt,
2446: Paul Federbush, Matthew Finn, Francesco Paparella, J\"org Schumacher, William
2447: R. Young, and many of the participants of the 2005 GFD Program at Woods Hole
2448: Oceanographic Institution where much of this work was performed.  TAS was
2449: supported in part by the National Science and Engineering Research Council of
2450: Canada through a Canadian Graduate Scholarship.  J-LT was supported in part by
2451: the UK Engineering and Physical Sciences Research Council grant GR/S72931/01.
2452: CRD was supported in part by NSF grants PHY-0244859, PHY-0555324 and an
2453: Alexander von Humboldt Research Award.
2454: 
2455: %\newpage
2456: 
2457: \appendix
2458: 
2459: \section{Lower Bound on Variance Efficiency}
2460: \label{sec:lowbound0}
2461: 
2462: One might expect the variance efficiency ${\cal E}_0$ to have a lower bound of
2463: unity, implying that stirring \emph{always} decreases the variance.  In order
2464: to derive a uniform (in $\Pe$) lower estimate we can search for the maximum
2465: variance subject to the steady-state production-dissipation constraint:
2466: \begin{equation}
2467:   \langle \theta^2 \rangle \geq \min_{\vartheta}~ \{ \langle
2468:   \vartheta^2 \rangle~|~ \kappa \langle | \nabla \vartheta |^2
2469:   \rangle = \langle s \vartheta \rangle \} .
2470: \end{equation}
2471: The solution of this optimization problem in Fourier space is
2472: \begin{equation}
2473:   \hat{\theta}(\bm{k})= \frac{\mu}{2} \frac{\hat{s}(\bm{k})}{\mu \kappa k^2
2474:   + 1}
2475: \end{equation}
2476: where $\mu$ is the Lagrange multiplier enforcing the constraint
2477: \begin{equation}
2478:   \kappa \sum_{\bm{k}} k^2 | \hat{\theta}(\bm{k})|^2 =
2479:   \sum_{\bm{k}} \hat{\theta}(\bm{k}) \hat{s}^*(\bm{k})~~
2480:   \quad \Longrightarrow \quad \kappa \sum_{\bm{k}} \frac{\mu k^2
2481:   |\hat{s}(\bm{k})|^2}{(\mu \kappa k^2 +1)^2} = 2 \sum_{\bm{k}}
2482:   \frac{|\hat{s}(\bm{k})|^2}{\mu \kappa k^2 +1}.
2483: \label{sums}
2484: \end{equation}
2485: where $*$ denotes the complex conjugate.
2486: 
2487: In the case of a dichromatic source with wavenumber $k_1$ at amplitude $s_1$
2488: and $k_2$ at amplitude $s_2$, the constraint requires one to solve a cubic
2489: equation for $\xi=1/\mu \kappa k_1^{2}$,
2490: \begin{equation}
2491:   (1+\alpha)\xi^3 + \tfrac{1}{2} (1+\alpha \beta +4 \beta +4 \alpha) \xi^2 +
2492:   (\beta +\alpha \beta + \beta^2 +\alpha) \xi +\tfrac{1}{2} (\beta^2+\alpha
2493:   \beta)=0
2494: \end{equation}
2495: where $\alpha=s_1^2/s_2^2$ and $\beta= k_1^2/k_2^2$.
2496: Then the mixing efficiency bound is
2497: \begin{equation}
2498:   {\cal E}_0^2 \ge \frac{\sum_{k_{1}, k_{2}}|\hat{\theta}_0(\bm{k})|^2}
2499:   {\sum_{k_{1}, k_{2}}|\hat{\theta}(\bm{k})|^2} =
2500:   \frac{4 (1+ \frac{\alpha}{\beta^2})}
2501:   {\frac{1}{(1+\xi)^2}+\frac{\alpha}{(\beta+\xi)^2}},
2502: \end{equation}
2503: and a numerical evaluation of the roots reveals that the minimum value of the
2504: efficiency bound is less than 1 for $\forall~\xi$.  Thus the variance
2505: production-dissipation balance alone does {\it not} rule out the existence of
2506: flows that could \emph{increase} the scalar variance.
2507: 
2508: %\newpage
2509: 
2510: \section{Steady uniform wind on a $\delta$-function source}
2511: 
2512: Taking the Fourier transform of the steady advection--diffusion
2513: equation~\eqref{eq:AD} with a uniform velocity field along the $x_{d}$-axis,
2514: \begin{equation}
2515:   %\sum_k
2516:   \hat{\theta}(\bm{k}) =
2517:   %\sum_k
2518:   \frac{\hat{s}(\bm{k})}{\kappa k^2 + i U k_d}
2519: \end{equation}
2520: where $k_d$ is the $d$th component of the horizontal wavenumber.
2521: Now substitute~$\hat{s}(\bm{k})=\text{const}$ for a~$\delta$-function
2522: source and after approximating the sums by integrals (since we are only
2523: interested in the asymptotic behavior), we find
2524: \begin{subequations}
2525: \begin{alignat}{2}
2526:   d&=2: \qquad &\langle |\nabla^{p}\theta|^2 \rangle &= \int_0^{2\pi} d\phi
2527:   \int_{2\pi/L}^{\infty} \frac{k^{2p+1} dk}{\kappa^2 k^4 + U^2
2528:     k^2\cos^2\phi}\,,\\
2529:   d&=3: &\langle| \nabla^{p}\theta|^2 \rangle &= \int_0^{2\pi} d\phi
2530:   \int_0^{\pi} \sin \vartheta d\vartheta \int_{2\pi/L}^{\infty} \frac{k^{2(p+1)}
2531:   dk}{\kappa^2 k^4 + U^2 k^2\cos^2\vartheta}\,.
2532: \end{alignat}%
2533: \label{eq:integrals}%
2534: \end{subequations}%
2535: The variances in the absence of stirring are found by calculating the
2536: integrals~\eqref{eq:integrals} with $U=0$. Straightforward evaluation of the
2537: integrals~\eqref{eq:integrals} yields
2538: \begin{subequations}
2539: \begin{alignat}{4}
2540:   d&=2:
2541:   \qquad
2542:   &{\cal E}_{1} &= 1,
2543:   \quad &{\cal E}_{0} &\sim \sqrt{\Pe}\,,
2544:   \quad &{\cal E}_{-1} &\sim \sqrt{\Pe}\,,\\
2545:   d&=3:
2546:   &{\cal E}_{1} &= 1,
2547:   \quad &{\cal E}_{0} &\sim 
2548:   \frac{\sqrt{\Pe}}{\sqrt{\log \Pe}}\,,
2549:   \quad &{\cal E}_{-1} &\sim \sqrt{\Pe}\,.
2550: \end{alignat}
2551: \end{subequations}
2552: These anomalous scalings in $\Pe$ suggest that the uniform flow is not the
2553: optimal allowed by the bound for the $\delta$-function source in both $d=2$
2554: and~$3$. This emphasizes the source-dependent nature of the optimal stirrer.
2555: 
2556: %\newpage
2557: 
2558: \section{Steady shear at an angle}
2559: 
2560: When the source is tilted at a 45$^{\circ}$ angle, i.e. $s({\bf x}) = \sqrt{2} S \sin(k_s(x+y))$, the functions $f$
2561: and $g$ must satisfy
2562: \begin{subequations}
2563: \begin{align}
2564:   -\sqrt{2}U\ks \sin(\ku y)\, g(y)
2565:   &= \kappa \left [ -\ks^2 + \frac{d^2}{dy^2}
2566:     \right] f(y) + \sqrt{2} S\sin(k_s y)\,,\label{B1a}\\
2567:   \sqrt{2}U\ks \sin(\ku y) f(y) &= \kappa \left [ -\ks^2 + \frac{d^2}{dy^2}
2568:     \right] g(y)+ \sqrt{2}S \cos(k_s y)\,.\label{B1b}
2569: \end{align}
2570: \label{eq:fgeqnstilted}%
2571: \end{subequations}%
2572: Upon changing variables as done in section $\rm{VII}$ we obtain
2573: \begin{subequations}
2574: \begin{align}
2575:   \frac{1}{\Pec}  \left[ -1 + \frac{d^2}{d \tilde{y}^2} \right]
2576:   \hat{f}(\tilde{y})+\sin \tilde{y} &= -\sin(r\tilde y) \hat{g}(\tilde{y}),\label{B2a}\\
2577:   \frac{1}{\Pec}  \left[ -1 + \frac{d^2}{d \tilde{y}^2} \right]
2578:   \hat{g}(\tilde{y}) + \cos \tilde{y} &= \sin(r\tilde y) \hat{f}(\tilde{y}).\label{B2b}
2579: \end{align}
2580: \label{eq:fgeqnsscaledtilted}%
2581: \end{subequations}%
2582: The internal layer solution is once again obtained by expanding in powers of
2583: $\Pec^{-1}$. To leading order the solution to~\eqref{eq:fgeqnsscaledtilted} in 
2584: the outer region is
2585: \begin{equation}
2586:   \hat{f}_{\mathrm{outer}} \sim 1,\qquad 
2587:   \hat{g}_{\mathrm{outer}} \sim -\frac{1}{ \tan(r \tilde{y})}\,.
2588:   \label{eq:outersoltilted}
2589: \end{equation}
2590: The inner internal layer solution is the same as for the untilted source
2591:  problem.  In Figure~\ref{dumfig6} we compare the internal layer solution
2592:  against the exact (numerical) solution of the full advection--diffusion
2593:  equation (\ref{eq:ADsteady}) at $\Pe=1000$ where the small parameter
2594:  $\epsilon \approx .2$.
2595: %
2596: \begin{figure}
2597: \centerline{\includegraphics[width=15cm]{figure6.eps}}
2598: \setlength{\abovecaptionskip}{-30pt}
2599: \caption{Comparison of the direct numerical solution (solid) and the internal
2600: layer solution (dashed) for \hbox{$\Pe$ = 1000}.}
2601: \label{dumfig6}
2602: \end{figure}
2603: %
2604: Recovering all the scalings and
2605: letting $\delta=\epsilon/r^{1/3} \ks$, we define the composite approximations
2606: \begin{subequations}
2607: \begin{align}
2608:   f_{\mathrm{comp}}(y) \ &= \ \frac{S}{U\ks} \,[\hat{f}_{\mathrm{inner}}(y) + f_{\mathrm{outer}}(y) ]
2609:   %= \frac{S}{U\ks}
2610:   %\frac{r^{-2/3}}{\epsilon}F \left( \frac{r^{1/3} \ks}{\epsilon} y \right)
2611:   \ = \frac{S}{U\ks}\, \left[ \frac{1}{\ku \delta} F \left(\frac{y}{\delta}\right)+1\right],\\
2612:   g_{\mathrm{comp}}(y) \ &= \ \frac{S}{U\ks} \,
2613:   \ku y \times \hat{g}_{\mathrm{inner}}(y) \times \hat{g}_{\mathrm{outer}}(y)
2614:   %=\frac{S}{U\ks} \frac{r^{-2/3}}{\epsilon}G
2615:   %\left( \frac{r^{1/3} \ks}{\epsilon} y \right) \frac{\ku y}{\sin(\ku y)}
2616:   %\\
2617:   \ = \frac{S}{U\ks}\, \frac{1}{\ku \delta}\,
2618:   \frac{\ku y}{\tan(\ku y)}\,G\left(\frac{y}{\delta}\right).
2619: \end{align}
2620: \label{eq:fginternalsoltilted}%
2621: \end{subequations}%
2622: Given the internal layer solution we can compute the multi-scale mixing
2623: measures.  In fact, to leading order the dependence of the efficiencies on the
2624: P\'eclet number is the same. The difference is in the prefactors. Figure
2625: \ref{dumfig7} compares the scalings derived from the internal layer solution
2626: (including prefactors) to the discrete data from the numerical solution and
2627: the numerical solution using the random sine flow.
2628: %
2629: \begin{figure}
2630: \centerline{\includegraphics[width=\textwidth]{figure7.eps}}
2631: \setlength{\abovecaptionskip}{-20pt}
2632: \caption{Mixing efficiencies ${\cal E}_p$ as a function of $\Pe$ for (a)
2633: $p=1$, (b) $p=0$, and (c) $p=-1$. The solid lines are the bounds, the
2634: dashed-dot lines are scalings predicted from the internal layer asymptotic
2635: analysis, the open circles are the discrete data from the numerical solution
2636: of the steady advection--diffusion equation (\ref{eq:ADsteady}), and the
2637: dashed lines are the result of the direct numerical solution for the
2638: time-dependent random sine flow.}
2639: \label{dumfig7}%
2640: \end{figure}%
2641: %
2642: 
2643: 
2644: 
2645: %\newpage
2646: %\bibliography{journals_abbrev,articles}
2647: 
2648: \begin{thebibliography}{10}
2649: \expandafter\ifx\csname url\endcsname\relax
2650:   \def\url#1{\texttt{#1}}\fi
2651: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
2652: 
2653: \bibitem{Ottino1990}
2654: J.~M. Ottino, Mixing, chaotic advection, and turbulence, Annu. Rev. Fluid Mech.
2655:   22 (1990) 207--253.
2656: 
2657: \bibitem{Majda1999}
2658: A.~J. Majda, P.~R. Kramer, Simplified models for turbulent diffusion: {T}heory,
2659:   numerical modelling and physical phenomena, Physics Reports 314~(4-5) (1999)
2660:   237--574.
2661: 
2662: \bibitem{Warhaft2000}
2663: Z.~Warhaft, Passive scalars in turbulent flows, Annu. Rev. Fluid Mech. 32
2664:   (2000) 203--240.
2665: 
2666: \bibitem{Shraiman2000}
2667: B.~I. Shraiman, E.~D. Siggia, Scalar turbulence, Nature 405 (2000) 639--646.
2668: 
2669: \bibitem{Sawford2001}
2670: B.~L. Sawford, Turbulent relative dispersion, Annu. Rev. Fluid Mech. 33 (2001)
2671:   289--317.
2672: 
2673: \bibitem{Falkovich2001}
2674: G.~Falkovich, K.~Gaw\c{e}dzki, M.~Vergassola, Particles and fields in
2675:   turbulence, Rev. Mod. Phys. 73~(4) (2001) 913--975.
2676: 
2677: \bibitem{Aref2002}
2678: H.~Aref, The development of chaotic advection, Phys. Fluids 14~(4) (2002)
2679:   1315--1325.
2680: 
2681: \bibitem{Wiggins2004}
2682: S.~Wiggins, J.~M. Ottino, Foundations of chaotic mixing, Phil. Trans. R. Soc.
2683:   Lond. A 362 (2004) 937--970.
2684: 
2685: \bibitem{Batchelor1949}
2686: G.~K. Batchelor, Diffusion in a field of homogeneous turbulence. {I}.
2687:   {E}ulerian analysis, Australian Journal of Scientific Research 2 (1949)
2688:   437--450.
2689: 
2690: \bibitem{Batchelor1952a}
2691: G.~K. Batchelor, Diffusion in a field of homogeneous turbulence. {II}. {T}he
2692:   relative motion of particles, Proc. Cambridge Phil. Soc. 48 (1952) 345--362.
2693: 
2694: \bibitem{Batchelor1959}
2695: G.~K. Batchelor, Small-scale variation of convected quantities like temperature
2696:   in turbulent fluid: {P}art 1. {G}eneral discussion and the case of small
2697:   conductivity, J. Fluid Mech. 5 (1959) 134--139.
2698: 
2699: \bibitem{Kraichnan1968}
2700: R.~H. Kraichnan, Small-scale structure of a scalar field convected by
2701:   turbulence, Phys. Fluids 11~(5) (1968) 945--953.
2702: 
2703: \bibitem{Kraichnan1974}
2704: R.~H. Kraichnan, Convection of a passive scalar by a quasi-uniform random
2705:   straining field, J. Fluid Mech. 64 (1974) 737--762.
2706: 
2707: \bibitem{Constantin2005}
2708: P.~Constantin, A.~Kiselev, L.~Ryzhik, A.~Zlato\u{s}, Diffusion and mixing in
2709:   fluid flow Preprint.
2710: 
2711: \bibitem{Townsend1951}
2712: A.~A. Townsend, The diffusion behind a line source in homogeneous turbulence,
2713:   Proc. R. Soc. Lond. A 209~(1098) (1951) 418--430.
2714: 
2715: \bibitem{Townsend1954}
2716: A.~A. Townsend, The diffusion behind a line source in homogeneous turbulence,
2717:   Proc. R. Soc. Lond. A 224~(1159) (1954) 487--512.
2718: 
2719: \bibitem{Saffman1960}
2720: P.~G. Saffman, On the effect of the molecular diffusivity in turbulent
2721:   diffusion, J. Fluid Mech. 8 (1960) 273--283.
2722: 
2723: \bibitem{Durbin1980}
2724: P.~A. Durbin, A stochastic model of two-particle dispersion and concentration
2725:   fluctuations in in homogeneous turbulence, J. Fluid Mech. 100 (1980)
2726:   279--302.
2727: 
2728: \bibitem{Drummond1982}
2729: I.~T. Drummond, Path-integral methods for turbulent diffusion, J. Fluid Mech.
2730:   123 (1982) 59--68.
2731: 
2732: \bibitem{Sawford1986}
2733: B.~L. Sawford, J.~C.~R. Hunt, Effects of turbulence structure, molecular
2734:   diffusion and source size on scalar fluctuations in homogeneous turbulence,
2735:   J. Fluid Mech. 165 (1986) 373--400.
2736: 
2737: \bibitem{Thomson1990}
2738: D.~J. Thomson, A stochastic model for the motion of particle pairs in isotropic
2739:   high-{R}eynolds number turbulence, and its application to the problem of
2740:   concentration variance, J. Fluid Mech. 210 (1990) 113--153.
2741: 
2742: \bibitem{Borgas1994}
2743: M.~S. Borgas, B.~L. Sawford, A family of stochastic models for two-particle
2744:   dispersion in isotropic homogeneous stationary turbulence, J. Fluid Mech. 279
2745:   (1994) 69--99.
2746: 
2747: \bibitem{Chertkov1997}
2748: M.~Chertkov, I.~Kolokolov, M.~Vergassola, Inverse cascade and intermittency of
2749:   passive scalar in one-dimensional smooth flow, Phys. Rev. E 56~(5) (1997)
2750:   5483--5499.
2751: 
2752: \bibitem{Chertkov1998}
2753: M.~Chertkov, G.~Falkovich, I.~Kolokolov, Intermittent dissipation of a passive
2754:   scalar in turbulence, Phys. Rev. Lett. 80~(10) (1998) 2121--2124.
2755: 
2756: \bibitem{Chertkov1995}
2757: M.~Chertkov, G.~Falkovich, I.~Kolokolov, V.~Lebedev, Statistics of a passive
2758:   scalar advected by a large-scale two-dimensional velocity field: Analytic
2759:   solution, Phys. Rev. E 51~(6) (1995) 5609--5627.
2760: 
2761: \bibitem{Chertkov1995b}
2762: M.~Chertkov, G.~Falkovich, I.~Kolokolov, V.~Lebedev, Normal and anomalous
2763:   scaling of the fourth-order correlation function of a randomly advected
2764:   passive scalar, Phys. Rev. E 52~(5) (1995) 4924--4941.
2765: 
2766: \bibitem{Chertkov1997b}
2767: M.~Chertkov, Instanton for random advection, Phys. Rev. E 55~(3) (1997)
2768:   2722--2735.
2769: 
2770: \bibitem{Balkovsky1999}
2771: E.~Balkovsky, A.~Fouxon, Universal long-time properties of {L}agrangian
2772:   statistics in the {B}atchelor regime and their application to the passive
2773:   scalar problem, Phys. Rev. E 60~(4) (1999) 4164--4174.
2774: 
2775: \bibitem{Pierrehumbert1994}
2776: R.~T. Pierrehumbert, Tracer microstructure in the large-eddy dominated regime,
2777:   Chaos Solitons Fractals 4~(6) (1994) 1091--1110.
2778: 
2779: \bibitem{Antonsen1996}
2780: T.~M. {Antonsen, Jr.}, Z.~Fan, E.~Ott, E.~Garcia-Lopez, The role of chaotic
2781:   orbits in the determination of power spectra, Phys. Fluids 8~(11) (1996)
2782:   3094--3104.
2783: 
2784: \bibitem{Rothstein1999}
2785: D.~Rothstein, E.~Henry, J.~P. Gollub, Persistent patterns in transient chaotic
2786:   fluid mixing, Nature 401~(6755) (1999) 770--772.
2787: 
2788: \bibitem{Fereday2002}
2789: D.~R. Fereday, P.~H. Haynes, A.~Wonhas, J.~C. Vassilicos, Scalar variance decay
2790:   in chaotic advection and {B}atchelor-regime turbulence, Phys. Rev. E 65~(3)
2791:   (2002) 035301(R).
2792: 
2793: \bibitem{Sukhatme2002}
2794: J.~Sukhatme, R.~T. Pierrehumbert, Decay of passive scalars under the action of
2795:   single scale smooth velocity fields in bounded two-dimensional domains: From
2796:   non-self-similar probability distribution functions to self-similar
2797:   eigenmodes, Phys. Rev. E 66 (2002) 056032.
2798: 
2799: \bibitem{Wonhas2002}
2800: A.~Wonhas, J.~C. Vassilicos, Mixing in fully chaotic flows, Phys. Rev. E 66
2801:   (2002) 051205.
2802: 
2803: \bibitem{Pikovsky2003}
2804: A.~Pikovsky, O.~Popovych, Persistent patterns in deterministic mixing flows,
2805:   Europhys. Lett. 61 (2003) 625--631.
2806: 
2807: \bibitem{Thiffeault2003d}
2808: J.-L. Thiffeault, S.~Childress, Chaotic mixing in a torus map, Chaos 13~(2)
2809:   (2003) 502--507.
2810: 
2811: \bibitem{Vanneste2005}
2812: J.~Vanneste, P.~H. Haynes, What controls the decay of passive scalars in smooth
2813:   flows?, Phys. Fluids 17 (2005) 097103.
2814: 
2815: \bibitem{Schekochihin2004}
2816: A.~Schekochihin, P.~H. Haynes, S.~C. Cowley, Diffusion of passive scalar in a
2817:   finite-scale random flow, Phys. Rev. E 70 (2004) 046304.
2818: 
2819: \bibitem{Gilbert2006}
2820: A.~D. Gilbert, Advected fields in maps. {III}. {D}ecay of passive scalar in
2821:   baker's maps, Dynam. Sys. 21 (2006) 25--71.
2822: 
2823: \bibitem{Thiffeault2004b}
2824: J.-L. Thiffeault, The strange eigenmode in {L}agrangian coordinates, Chaos
2825:   14~(3) (2004) 531--538.
2826: 
2827: \bibitem{Mathew2005}
2828: G.~Mathew, I.~Mezi{\'c}, L.~Petzold, A multiscale measure for mixing, Physica D
2829:   211~(1-2) (2005) 23--46.
2830: 
2831: \bibitem{Thiffeault2004}
2832: J.-L. Thiffeault, C.~R. Doering, J.~D. Gibbon, A bound on mixing efficiency for
2833:   the advection--diffusion equation, J. Fluid Mech. 521 (2004) 105--114.
2834: 
2835: \bibitem{Doering2002}
2836: C.~R. Doering, C.~Foias, Energy dissipation in body-forced turbulence, J. Fluid
2837:   Mech. 467 (2002) 289--306.
2838: 
2839: \bibitem{Doering2003}
2840: C.~R. Doering, B.~Eckhardt, J.~Schumacher, Energy dissipation in body-forced
2841:   plane shear flow, J. Fluid Mech. 494 (2003) 275--284.
2842: 
2843: \bibitem{Tsang2005}
2844: Y.-K. Tsang, T.~M. {Antonsen, Jr.}, E.~Ott, Exponential decay of chaotically
2845:   advected passive scalars in the zero diffusivity limit, Phys. Rev. E 71
2846:   (2005) 066301.
2847: 
2848: \bibitem{Dittrich1984}
2849: P.~Dittrich, S.~A. Molchanov, D.~D. Sokoloff, A.~A. Ruzmaikin, Mean magnetic
2850:   field in renovating random flows, Astron. Nachr. 3 (1984) 119--125.
2851: 
2852: \bibitem{Zeldovich1984}
2853: Ya.~B. Zeldovich, A.~A. Ruzmaikin, S.~A. Molchanov, D.~D. Sokoloff, Kinematic
2854:   dynamo problem in a linear velocity field, J. Fluid Mech. 144 (1984) 1--11.
2855: 
2856: \bibitem{Gilbert1992}
2857: A.~D. Gilbert, B.~J. Bayly, Magnetic field intermittency and fast dynamo action
2858:   in random helical flows, J. Fluid Mech. 241 (1992) 199--214.
2859: 
2860: \bibitem{Zeldovich1982}
2861: Ya.~B. Zeldovich, Exact solution to the problem of diffusion in a periodic
2862:   velocity field and turbulent diffusion, Doklady Akademiia Nauk SSSR. 266
2863:   (1992) 821--826.  English translation in {\it Selected Works of Yakov
2864:   Borisovich Zeldovich. {V}olume {I}: {C}hemical Physics and Hydrodynamics},
2865:   Princeton University Press (1992) 86--92.
2866: 
2867: \bibitem{ShawGFD2005}
2868: T.~A. Shaw, Bounds on multiscale mixing efficiencies, in: Proceedings of the
2869:   2005 Summer Program in Geophysical Fluid Dynamics, Woods Hole Oceanographic
2870:   Institute, Woods Hole, MA, 2005, p. 291.
2871: 
2872: \bibitem{Plasting2006}
2873: S.~Plasting, W.~R. Young, A bound on scalar variance for the
2874:   advection--diffusion equation, J. Fluid Mech. 552 (2006) 289--298.
2875: 
2876: \bibitem{DoeringThiffeault2006}
2877: C.~R. Doering, J.-L. Thiffeault, Multiscale mixing efficiencies for steady
2878:   sources, Phys. Rev. E, In press.
2879: 
2880: \bibitem{Savill1987}
2881: A.~M. Savill, Recent developments in rapid-distortion theory, Annu. Rev. Fluid
2882:   Mech. 19 (1987) 531--575.
2883: 
2884: \bibitem{Hunt1990}
2885: J.~C.~R. Hunt, D.~J. Carruthers, Rapid distortion theory and the `problems' of
2886:   turbulence, J. Fluid Mech. 212 (1990) 497--532.
2887: 
2888: \bibitem{Balmforth2003}
2889: N.~J. Balmforth, W.~R. Young, Diffusion-limited scalar cascades, J. Fluid Mech.
2890:   482 (2003) 91--100.
2891: 
2892: \end{thebibliography}
2893: 
2894: \end{document}