1: \documentclass{elsart}
2: \usepackage{epsfig}
3: \usepackage{graphicx}
4: \usepackage{amssymb}
5: \usepackage{citesort}
6: \usepackage{psfrag}
7: %\usepackage[notref,notcite]{showkeys}
8: %\setlength{\textwidth}{16cm}
9: %\setlength{\textheight}{22cm}
10: %\setlength{\Omega_ddsidemargin}{1cm}
11: %\setlength{\evensidemargin}{0cm}
12: %\renewcommand{\baselinestretch}{1.3}
13: \begin{document}
14: %\bibliographystyle{prsty}
15: \def\LL {\Lambda ^{\!\!\scriptscriptstyle-1}}
16: \newcommand{\B}[1]{\mbox{\boldmath $#1$}}
17: \newcommand{\C}[1]{\mathcal #1}
18: \newcommand{\rb}{Rayleigh-B\'enard }
19: \newcommand{\Br}{\B{r}}
20: \def\Obr{\overline{\Br}}
21: \newcommand{\KO}{\hat{\mathcal{K}}}
22: \newcommand{\KM}{{\Bbb K}}
23: \newcommand{\PM}{{\Bbb P}}
24: \newcommand{\MM}{{\Bbb M}}
25: \newcommand{\UM}{{\Bbb U}}
26: \newcommand{\NM}{{\Bbb N}}
27: \newcommand{\VecII}[2]{\left( \begin{array}{c} #1 \\ #2 \end{array} \right)}
28: \newcommand{\MatII}[4]{\left( \begin{array}{cc} #1 & #2 \\
29: #3 & #4 \end{array}\right)}
30: \newcommand{\BR}{\B{R}}
31: \newcommand{\FT}{{\hat F}}
32: \newcommand{\BB}{\B{B}}
33: \newcommand{\Bx}{\B{x}}
34: \newcommand{\By}{\B{y}}
35: \newcommand{\Bk}{\B{k}}
36: \newcommand{\Bq}{\B{q}}
37: \newcommand{\Bp}{\B{p}}
38: \newcommand{\f}{\B{f}}
39: \renewcommand{\v}{\B{v}}
40: \newcommand{\Bv}{\B{v}}
41: \newcommand{\Bu}{\B{u}}
42: \newcommand{\Bt}{\B{t}}
43: \newcommand{\Bn}{\B{n}}
44: \newcommand{\Be}{\B{e}}
45: \newcommand{\be}{\begin{equation}}
46: \newcommand{\ee}{\end{equation}}
47: \newcommand{\bea}{\begin{eqnarray}}
48: \newcommand{\eea}{\end{eqnarray}}
49: \renewcommand{\k}{{\bf k}}
50: \newcommand{\x}{{\bf x}}
51: \renewcommand{\r}{{\bf r}}
52: \def\Or {\overline{r}}
53: \def\dv{{\delta_r v}}
54: \def\cP{{\mathcal P}}
55: \def\cT{{\mathcal T}}
56: \def\cS{{\mathcal S}}
57: \def\cC{{\mathcal C}}
58: \def\Btensor{ B_{n,jm}^{\alpha_1 \ldots \alpha_{n}} }
59: \def\Ctensor #1#2{ \delta^{\alpha_{#1}\alpha_{#2}}
60: B_{n-2,jm}^{ \stackrel{ \mbox{no } #1 ,#2} {
61: \overbrace{_{\alpha_1 \ldots \alpha_{n}} } }}}
62: \renewcommand{\a}{\alpha}
63: \renewcommand{\e}{\epsilon}
64: \newcommand{\eb}{\bar \epsilon}
65: \renewcommand{\b}{\beta}
66: \renewcommand{\l}{\ell}
67: \newcommand{\lp}{\left(}
68: \newcommand{\rp}{\right)}
69: \newcommand{\la}{\langle}
70: \newcommand{\ra}{\rangle}
71: \newcommand{\ut}{\underline{\theta}} \newcommand{\uu}{\underline{u}}
72: \newcommand{\ud}{\underline{\Delta}} \newcommand{\PO}{\hat{\mathcal{P}}}
73: \newcommand{\OO}{\hat{\mathcal{O}}}
74: \newcommand{\LM}{{\Bbb M}}
75: \newcommand{\OM}{{\Bbb O}}
76: \newcommand{\PRO}{\hat{\mathcal{P}}_{\mathrm{R}}}
77: \newcommand{\PLO}{\hat{\mathcal{P}}_{\mathrm{L}}}
78: \newcommand{\Eq}[1]{Eq.~(\ref{#1})}
79: \newcommand{\Fig}[1]{Fig.~\ref{#1}}
80: \newcommand{\EqDef}{\stackrel{\mathrm{def}}{=}}
81: \newcommand{\AO}{\mathcal{A}}
82: \newcommand{\BO}{\mathcal{B}}
83: \newcommand{\CO}{\mathcal{O}}
84: \newcommand{\DO}{\mathcal{D}}
85: \newcommand{\sepBR}[3]{\begin{picture}(9,1)
86: \put(#2,#3){\line(1,0){#1}\line(0,1){.25}}\end{picture}}
87: \newcommand{\sepTL}[3]{\begin{picture}(9,1)
88: \put(#2,#3){\line(0,-1){0.25}}
89: \put{#2,#3}{\line(1,0){#1}}\end{picture}}
90: \renewcommand{\S}{{\mathcal S}}
91: \newcommand{\Even}{{\mbox{\tiny Even}}}
92: \newcommand{\unitr}{{\mbox{\boldmath$\hat r$}}}
93: \newcommand{\unitR}{{\mbox{\boldmath$\hat R$}}}
94: \def\unitz{{\mbox{\boldmath$\hat z$}}}
95: \newcommand{\SO}{$SO(3)$}
96: \begin{frontmatter}
97: \title{Anisotropy in Turbulent Flows and in Turbulent Transport}
98: \author{Luca Biferale}
99: \address{Dept of Physics and INFM, University of Rome ``Tor
100: Vergata'', Via della Ricerca Scientifica 1, 00133 Roma, Italy}
101: \author{Itamar Procaccia}
102: \address{Dept. of Chemical Physics, The Weizmann Institute of Science,
103: Rehovot 76100, Israel}
104:
105: \begin{abstract}
106: The problem of anisotropy and its effects on the statistical theory of
107: high Reynolds-number (Re) turbulence (and turbulent transport) is
108: intimately related and intermingled with the problem of the
109: universality of the (anomalous) scaling exponents of structure
110: functions. Both problems had seen tremendous progress in the last five
111: years. In this review we present a detailed description of the new
112: tools that allow effective data analysis and systematic theoretical
113: studies such as to separate isotropic from anisotropic aspects of
114: turbulent statistical fluctuations. Employing the invariance of the
115: equations of fluid mechanics to all rotations, we show how to
116: decompose the (tensorial) statistical objects in terms of the
117: irreducible representation of the SO($d$) symmetry group (with $d$
118: being the dimension, $d=2$ or 3). This device allows a discussion of
119: the scaling properties of the statistical objects in well defined
120: sectors of the symmetry group, each of which is determined by the
121: ``angular momenta" sector numbers $(j,m)$. For the case of turbulent
122: advection of passive scalar or vector fields, this decomposition
123: allows rigorous statements to be made: (i) the scaling exponents are
124: universal, (ii) the isotropic scaling exponents are always leading,
125: (iii) the anisotropic scaling exponents form a discrete spectrum which
126: is strictly increasing as a function of $j$. This emerging picture
127: offers a complete understanding of the decay of anisotropy upon going
128: to smaller and smaller scales. Next we explain how to apply the SO(3)
129: decomposition to the statistical Navier-Stokes theory. We show how to
130: extract information about the scaling behavior in the isotropic
131: sector. Doing so furnishes a systematic way to assess the
132: universality of the scaling exponents in this sector, clarifying the
133: anisotropic origin of the many measurements that claimed the
134: opposite. A systematic analysis of Direct Numerical Simulations (DNS) of the
135: Navier-Stokes equations and of experiments provides a strong support
136: to the proposition that also for the non-linear problem there
137: exists foliation of the statistical theory into sectors of the
138: symmetry group. The exponents appear universal in each sector, and
139: again strictly increasing as a function of $j$. An approximate
140: calculation of the anisotropic exponents based on a closure theory is
141: reviewed. The conflicting experimental measurements on the decay of
142: anisotropy are explained and systematized, showing agreement with the
143: theory presented here.
144: \end{abstract}
145: \begin{keyword} Fully Developed Turbulence, Anisotropic Turbulence,
146: Turbulent Transport, SO(3) Decomposition, Intermittency, Universality of
147: Anomalous Exponents, Analysis of Turbulent data.
148: \PACS 47.27.Ak, 47.10+g, 47.27.Eq, 05.40-a, 47.27-i, 47.27.Nz, 47.27.Jv,
149: 47.27.Gs
150: \end{keyword}
151: \end{frontmatter}
152: \tableofcontents
153: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
154: %%
155: %
156: % Chapter 1 - Introduction
157: %
158: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
159: %%
160: \section{Introduction}
161: \label{chap:introduction}
162: The statistical theory of fluid turbulence is concerned with
163: correlation functions of the turbulent velocity vector field $\B
164: u(\B x,t)$ where $\B x$ is the spatial position and $t$ the time
165: \cite{my75}. Since the velocity field is a vector, multi-point and
166: multi-time correlation functions are in general tensor functions
167: of the vector positions and the scalar times. Naturally such
168: functions have rather complicated forms which are difficult to
169: measure and to compute. Consequently, almost from its very
170: beginning, the statistical theory of turbulence was
171: discussed in the context of an isotropic and homogeneous model.
172: The notion of isotropic turbulence was first introduced by
173: G.~I.~Taylor in 1935 \cite{tay35}. It refers to a turbulent
174: flow, in which the statistical averages of every function of the
175: velocity field and its derivatives with respect to a particular
176: frame of axes is invariant to any rotation in the axes. This is a
177: very effective mathematical simplification which, if properly
178: used, can drastically reduce the mathematical complexity of the
179: theory. For this reason, it was very soon adopted by others, such
180: as T.~D.~K\'arm\'an and L.~Howarth \cite{kar38} who derived the
181: K\'arm\'an-Howarth equation (see below), and A.~N.~Kolmogorov
182: \cite{kol41,fri95} who derived the 4/5 law (re-derived
183: below). In fact, most of the theoretical work in turbulence in the
184: past sixty years was limited to the isotropic model.
185:
186: Experimentally, however, we know that isotropy holds only as an
187: approximation with a varying degree of justification. In all realistic
188: flows there always exists some anisotropy at all scales; the
189: statistical properties of the velocity field are effected by the
190: geometry of the boundaries or the driving mechanism, which are never
191: rotationally invariant
192: \cite{hin75,she00,lum67,sad94,kur00,lum65,tav81,kur00a,fer00}. Therefore, a
193: realistic description of
194: turbulence cannot be purely isotropic and must contain some
195: anisotropic elements. Yet the problem is that once we take anisotropy
196: into account, we face a drastic increase in the complexity of the
197: theory. The number of variables that is needed to describe the common
198: statistical quantities, such as correlation functions and structure
199: functions of the velocity field, increases a lot. For example, under
200: shear there is a characteristic length scale, which can be constructed from
201: the
202: typical velocity and the typical shear \cite{lum67,bif02}.
203: This length has to be
204: considered in order to distinguish those scales where the turbulent
205: evolution is mainly dominated by the inertial effects of fluid
206: mechanics or by the direct input of energy due to the anisotropic
207: shear \cite{ben02,cas00,tos99,tos00}. Similarly,
208: all dimensional estimates acquire a significant
209: degree of ambiguity because of the proliferation of different
210: dimensional quantities related to the parameters of anisotropy. As a
211: consequence of these inherent difficulties the existing anisotropic
212: effects were simply ignored in many of the experimental and
213: simulational studies of statistical turbulence. This attitude gave
214: rise to ambiguous assessments of important fundamental issues like the
215: universality of the scaling exponents in turbulence.
216:
217: The standard justification for ignoring anisotropic effects is
218: that the basic phenomenology, since the pioneering works of
219: Kolmogorov \cite{kol41,fri95}, predicts a {\it
220: recovery of isotropy} at sufficiently small scales of the
221: turbulent flows. Nevertheless, both recent experimental works and
222: theoretical analysis suggested that the actual rate of recovery
223: is much slower than predicted by simple dimensional analysis,
224: pointing out even the possibility that some anisotropic
225: correlation function, based on velocity gradients, stays $O(1)$
226: for any Re \cite{pum95,pum96,she00,sch02}. In order
227: to settle this kind of problems, theoretically or experimentally,
228: it is crucial to possess systematic tools to disentangle
229: isotropic from anisotropic fluctuations and to distinguish among
230: different kinds of anisotropic fluctuations. Thus a central
231: challenge in the theory of anisotropy in turbulence is
232: the construction of an efficient mathematical language to
233: describe it. Without a proper description, the complexities of
234: the formalism can soon obscure the physical content of the
235: processes that we wish to study.
236:
237: The problem of anisotropy is not disconnected from the other
238: fundamental problem which has to do with the nature of
239: universality in turbulence. By universality, we mean the tendency
240: of different turbulent systems to show the same small-scales
241: statistical behavior when the measurements are done far away from
242: the boundaries. Consider, for example, the longitudinal two-point
243: structure function
244: \begin{equation}
245: S^{(2)}(\Br) \equiv \left< \delta u_{\l}^2(\Bx, \Br,t) \right> \ , \quad
246: \delta u_{\l}(\Bx,\Br,t) \equiv
247: \hat{\Br}\cdot\big[\B{u}(\Bx+\Br,t)-\B{u}(\Bx,t)\big] \ ,
248: \label{defS2long}
249: \end{equation}
250: with $\hat{\Br}$ being the unit vector in the direction of $\Br$, and
251: $\left< \cdot \right>$ stands for an appropriate ensemble
252: average. This function shows essentially the same dependence on the
253: separation vector $\Br$, whether it is measured in the atmospheric
254: boundary layer, in a wind tunnel or in a DNS, provided
255: it is measured for sufficiently small separations and far from the
256: boundaries. This high-degree of universality cannot be expected if
257: anisotropic fluctuations were the dominant contributions to the
258: two-point structure functions. Different boundary conditions and
259: different forcing mechanism necessarily introduce different
260: large-scale anisotropies in the flow, which would translate to
261: different small-scale anisotropic fluctuations. Small-scale
262: universality can be achieved only if anisotropic fluctuations are
263: sub-leading with
264: respect to the isotropic fluctuations. In the following, we also
265: discuss which aspects of the anisotropic fluctuations are {\it universal}
266: and
267: which are not. We will see that some aspects of the anisotropic
268: fluctuations
269: depend on the boundary conditions while other aspects do
270: not. In fact we will show that scaling exponents are expected to be
271: universal whereas amplitudes depend on the boundary conditions.
272:
273: In the last 5 years, a tremendous progress in the understanding of
274: the two aforementioned problems, i.e., finding a mathematical
275: language that properly describes anisotropic turbulence and its universal
276: properties has been achieved. Not surprisingly, the two
277: problems are closely related, as it often happens in physics - a
278: problem becomes considerably simpler if described in the proper
279: mathematical language.
280: The technical core of these recent achievements is the SO($3$)
281: decomposition \cite{ara99b}. This tool enjoys the advantages of being
282: mathematically simple, yet very powerful and systematic. By using
283: it, many of the mathematical complexities of dealing with
284: anisotropy in turbulence and in other hydrodynamic problems are
285: greatly simplified. The principal idea is to represent the main
286: statistical observables, such as structure functions and
287: correlation functions, in terms of their projections on the
288: different $(j, m)$ sectors of the group of rotations.
289: It can be applied to all the
290: statistical quantities in turbulence, creating a detailed profile
291: of the effects of anisotropy. Additionally, and perhaps more importantly,
292: the
293: SO($3$) decomposition reveals some new universal properties of
294: fully developed turbulence. It is expected that each sector of the
295: SO($3$) group has its own universal exponents. In particular, it is shown
296: that the exponents associated with the anisotropic sectors are
297: larger than the isotropic exponents, in accordance with
298: the isotropization of the statistics as smaller and smaller
299: scales are observed.
300:
301: As already mentioned, the SO($3$) decomposition is useful also to
302: investigate isotropic and anisotropic fluctuations in other
303: hydrodynamic problems. In particular we will focus on the case
304: of scalar and vector quantities passively advected by a turbulent velocity
305: field. In
306: these cases, one may often elevate the phenomenological assumptions
307: made for turbulent anisotropic fluctuations to the status of rigorous
308: statements \cite{fal01}. By using a systematic decomposition in
309: different sectors of the SO($3$) group one may show that passive
310: scalars, advected by stochastic self-similar Gaussian velocity
311: fields, always possess isotropic {\it leading} small-scale
312: fluctuations. Moreover, one may quantitatively distinguish among
313: different kinds of anisotropies, assessing their rate of decay by
314: going to smaller and smaller scales. It turns out that the rate
315: of {\it recovery of isotropy} is typically much slower than
316: expected on the basis of dimensional analysis. Moreover, all
317: different anisotropic fluctuations decay in a self-similar way
318: but with different rates; the scaling exponents being universal,
319: while prefactors are non-universal \cite{ara00}. The very same
320: can be rigorously proved for the passive advection of vector-like
321: quantities, as for the case of magnetic fields when the feedback
322: on the velocity evolution due to the Lorentz force is neglected.
323: There, the vector nature of the transported quantity leads to an
324: even richer, and more complex, list of possible anisotropic
325: fluctuations \cite{ara00a,lan99}. Another important problem which we address
326: in
327: detail is the case of the passive advection of a vector-like
328: incompressible quantity, i.e. a {\it passive vector with
329: pressure} \cite{ara01}. Although without any counterpart in nature, such a
330: system is particularly interesting because it can be seen, for
331: some aspects, as the closest {\it linear} approximation to the
332: non-linear Navier-Stokes evolution. For example, it allows to
333: study in a systematic way some problems
334: connected to the convergence of integrals involving the
335: pressure term. Similar technical problems arise
336: also in the analysis of both isotropic and anisotropic multi-point
337: velocity correlations in Navier-Stokes equations.
338:
339: Of course, a significant part of this review will be devoted to
340: applications of the theoretical and technical tools to physical
341: experimental data \cite{ara98,kur00,kur00a,she02,she02a,wander,iacob04}
342: and numerical data sets \cite{ara99,bif01,bif01a,bif02,ish02,sch00}.
343: In order to exploit the
344: entire potentiality of the SO($3$) decomposition one needs to measure
345: the whole velocity field, $\B u(\B x)$, in a 3 dimensional volume. This is
346: because in order to disentangle different projections on different
347: sectors one needs to integrate the given correlation function against
348: the proper eigenfunction of the rotation group on the 3$d$ sphere of
349: radius $r$. By doing that, the exact projection on each different
350: sector of the SO($3$) decomposition is under control, with the only
351: practical limitations for reaching highly anisotropic sectors being
352: the lack of resolution of highly fluctuating angular properties. At
353: the present stage of experimental capabilities the exact decomposition
354: can be carried out explicitly only in data sets coming from DNS.
355: Here the velocity field in the
356: whole testing volume is available. For experimental data, the best way
357: to exploit the SO($3$) decomposition is to either select observables
358: with vanishing isotropic components, in order to focus directly on
359: anisotropic sectors, or to perform a multi-sector analysis, i.e. to
360: fit simultaneously the isotropic and anisotropic components.
361:
362: The review is organized as follows. Sect. \ref{historyiso} offers a
363: historical review of isotropic turbulence. We present a modern derivation
364: of the exact results pertaining to the 3'rd order structure function
365: and the celebrated 4/5 law. We review the standard theory for all
366: correlation functions, and discuss the experimental difficulties with
367: the isotropic theory. These difficulties included apparent persistence
368: of anisotropies into the small scales for high Re,
369: apparent location dependent scaling exponents etc. In Sect.
370: \ref{history-aniso}
371: we review the history of attempts to deal with anisotropy. In Section
372: \ref{chap:modern}, the technical basis of the SO($3$) decomposition is
373: introduced focusing on the particular statistical problems of
374: anisotropic fluctuations discussed in the previous section. Then, in
375: Sect. \ref{chap:analytical} we switch to study {\it exactly solvable}
376: hydrodynamic problems with emphasis on either those aspects
377: peculiar to each different model and to those features in common
378: with the non-linear Navier-Stokes case. Among the common aspects
379: we cite the possibility to study in these models in full details the
380: {\it foliation} of the equations of correlation functions in different
381: anisotropic sectors; the universality of isotropic and anisotropic
382: exponents; the hierarchical organization of exponents --leading to
383: {\it recovery of small-scales isotropy}. At the end of this section
384: we present closure results for two-point turbulent structure function
385: in the anisotropic sectors $j=2,4,6$. In Sect.
386: \ref{chap:experiment}, the utility of this language is demonstrated by
387: discussing experimental data in atmospheric boundary layer and on
388: homogeneous-shear flows. In Sect.
389: \ref{chap:numerics},
390: we present the analysis of anisotropy in DNS of
391: typical strongly anisotropic flows. Two cases are discussed in depth:
392: channel flows and random Kolmogorov flows, the latter being homogeneous
393: flows
394: stirred at the large scales. Sect.
395: \ref{chap:conclusions} presents a summary and conclusions.
396: Technical details are collected in the appendices.
397: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
398: %%
399: %
400: % Section 1 - Historical review and state of the art
401: %
402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
403: %%
404: \section{Historical Review: Isotropic Turbulence}
405: \label{historyiso}
406: In the first two sections we present a historical review. We start
407: with the model of homogeneous isotropic turbulence, and then turn to
408: previous attempts to treat theoretically anisotropy in
409: turbulence.
410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
411: \subsection{Homogeneous and Isotropic Turbulence}
412: The Navier-Stokes equations for the velocity field are invariant to
413: all rotations:
414: \begin{eqnarray}
415: \frac{\partial \Bu(\Bx,t)}{\partial t} +\left[\Bu(\Bx,t)\cdot
416: \B\nabla\right]\Bu(\Bx,t)&=&
417: -\B\nabla p(\Bx,t) +\nu \Delta^2 \Bu(\Bx,t) \ , \label{NS}\\
418: \B \nabla \cdot \Bu(\Bx,t) &=&0,\nonumber
419: \end{eqnarray}
420: with $p(\Bx,t)$ and $\nu$ being the pressure and kinematic viscosity
421: respectively. Since the gradient and Laplacian operators are both
422: rotationally invariant, the rotation symmetry of the equation can be
423: broken only by anisotropic forcing terms or anisotropic boundary
424: conditions. Rather naturally then the statistical theory of turbulence
425: was mostly developed in the framework of \emph{isotropic turbulence}
426: \cite{tay35}.
427: The central idea of this approximation is that
428: the statistical average of any function of the velocity
429: components in any coordinate system is unaltered
430: if this coordinate system is rotated or reflected in any manner.
431: The assumption of isotropy was widely adopted. In 1938,
432: K\'arm\'an and Howarth \cite{kar38} used it to explore the
433: second- and third-order correlation-functions of the velocity field. Their
434: use of tensor notation was more elegant and compact than that used by
435: Taylor. It enabled them to derive some constraints on these correlation
436: functions and express them in terms of a few scalar functions. For example,
437: for the second-order correlation-function in homogeneous turbulence
438: \begin{equation}
439: \label{def:KA-C2}
440: C^{\alpha\beta}(\Br,t)
441: \equiv \left< u^\alpha(\Bx+\Br,t) u^\beta(\Bx,t) \right> \ ,
442: \end{equation}
443: they used the representation
444: \begin{equation}
445: \label{eq:KA-C2}
446: C^{\alpha\beta}(\Br,t) = [f(r,t)-g(r,t)]\hat{r}^\alpha\hat{r}^\beta
447: + g(r,t)\delta^{\alpha\beta} \ ,
448: \end{equation}
449: and then derived a linear differential relation between $f(r,t)$ and
450: $g(r,t)$ using the solenoidal condition of $\partial_\alpha
451: C^{\alpha\beta}(\Br,t)=0$,
452: $$
453: 2f(r,t) - 2g(r,t) = -r\frac{\partial f(r,t)}{\partial r} \ .
454: $$
455: This means that under the assumption of isotropy, and using the
456: solenoidal condition, the second-order correlation-function can be
457: written in terms of one scalar function instead of nine.
458: Similarly, K\'arm\'an and Howarth analyzed the third-order correlation
459: function by representing it as an isotropic tensor and then reducing the
460: number of scalar functions using the solenoidal condition. They were
461: also able to connect it to the second-order correlation function in
462: decaying turbulence using the Navier-Stokes equations. These computations
463: have
464: since found their way into every standard text-book on the
465: statistical theory of turbulence.
466:
467: The mathematical representation of isotropic turbulence has
468: reached its most elegant and powerful form in a paper by
469: H.~P.~Robertson from 1940 \cite{rob40}.
470: Robertson provided a systematic way to represent isotropic
471: tensors using the theory of invariants. For example, to derive the general
472: representation (\ref{eq:KA-C2}) in the stationary case using Robertson's
473: method, we
474: consider the \emph{scalar} function
475: $$
476: C(\B{a}, \B{b}, \Br) \equiv C^{\alpha\beta}(\Br)a_\alpha b_\beta \ ,
477: $$
478: with $\B{a}$ and $\B{b}$ being two arbitrary vectors. If
479: $C^{\alpha\beta}(\Br)$ were an isotropic tensor, $C(\B{a}, \B{b},
480: \Br)$ would preserve its functional form upon an arbitrary
481: (simultaneous) rotation of the three vectors $\Br, \B{a}, \B{b}$.
482: Using invariant theory, Robertson deduced that $C(\B{a}, \B{b},
483: \Br)$ must be a function of the six possible scalar products
484: $(\Br\cdot\Br)$, $(\Br\cdot\B{a})$,... and of the determinant
485: $[\Br\B{a}\B{b}]\equiv\epsilon_{\mu\alpha\beta}r^\mu a^\alpha
486: b^\beta$. Additionally, by definition, it must be a bilinear
487: function of $\B{a}$ and $\B{b}$ and therefore must have the
488: following form:
489: $$
490: C(\B{a}, \B{b}, \Br) = A(r)(\Br\cdot\B{a})(\Br\cdot\B{b})
491: + B(r)(\B{a}\cdot\B{b}) + C(r)[\Br\B{a}\B{b}] \ ,
492: $$
493: where $A(r)$, $B(r)$ and $C(r)$ are arbitrary functions. Finally,
494: recalling that $C(\B{a}, \B{b}, \Br)$ is the contraction of
495: $C^{\alpha\beta}(\Br)$ with $\B{a}$ and $\B{b}$, we find that
496: $$
497: C^{\alpha\beta}(\Br) = A(r)r^\alpha r^\beta + B(r)\delta^{\alpha\beta}
498: + C(r)\epsilon^{\mu\alpha\beta}r_\mu \ .
499: $$
500: If we further demand $C^{\alpha\beta}(\Br)$ to be invariant to
501: improper rotations as well (i.e., rotations plus reflections), we
502: can drop the skew-symmetric part $\epsilon^{\mu\alpha\beta}$,
503: thus retaining a representation which is equivalent to
504: \Eq{eq:KA-C2}.
505: %%%%%%%%%%%%%%%%%%%%%
506: \subsection{The 4/5 law in Isotropic Turbulence and its Generalization}
507: By using the isotropic representation of the third-order
508: correlation function, in 1941 Kolmogorov proved the ``four-fifth law''
509: well inside the inertial range of a fully developed turbulence. This law
510: pertains
511: to the third order moment of longitudinal velocity differences,
512: stating that in homogeneous, isotropic and stationary turbulence,
513: in the limit of vanishing kinematic viscosity $\nu\to 0$
514: $$
515: \left<[\delta u_l(\B x,\Br,t)]^3\right> = -\frac{4}{5}
516: \eb r \ , $$
517: where $\eb$ is the mean
518: energy flux per unit time and mass, $\eb\equiv
519: \nu\left<|\nabla_\alpha u_\beta|^2\right>$. The fundamental assumption
520: needed to derive this law is the so-called ``dissipation anomaly''
521: which means that the dissipation is finite in
522: the limit $\nu\to 0$. As noted in \cite{fri95}, ``this is one of the
523: most important results in fully developed turbulence because it
524: is both exact and nontrivial. It thus constitutes a kind of
525: `boundary condition' on theories of turbulence: such theories, to
526: be acceptable, must either satisfy the four-fifth law, or
527: explicitly violate the assumptions made in deriving it".
528:
529: To demonstrate how isotropy helps in deriving this result, we
530: present a re-derivation in which we will obtain an additional
531: exact relation that appears to have the same status as the
532: four-fifth law, pertaining to homogeneous, stationary and
533: isotropic turbulence with helicity \cite{chk96,lvo97a}. Defining the
534: velocity $\Bv(\Bx,t)$ as
535: $\B v(\Bx,t)\equiv \B u(\Bx,t)-\left<\B u\right>$
536: we consider the simultaneous 3rd order tensor
537: correlation function which depends on two space points:
538: \begin{equation}
539: J^{\alpha,\beta\gamma}(\Br)\equiv
540: \left<v^\alpha(\Bx+\Br,t)v^\beta(\B x,t) v^\gamma(\B x,t)\right> \ .
541: \label{defJ}
542: \end{equation}
543: We show that in the limit $\nu\to 0$, under the same assumption
544: leading to the fourth-fifth law, this correlation function reads
545: \cite{lvo97a}
546: \begin{equation}
547: J^{\alpha,\beta\gamma}(\Br)=-{\eb \over
548: 10}(r^\gamma\delta_{\alpha\beta}
549: +r^\beta\delta_{\alpha\gamma}-{2\over
550: 3}r^\alpha\delta_{\beta\gamma}) - {h\over 30}
551: (\epsilon_{\alpha\beta\delta}r^\gamma+\epsilon_{\alpha\gamma\delta}r^\beta)
552: r^\delta \ , \label{result}
553: \end{equation}
554: where $\delta_{\alpha\beta}$ is the Kronecker delta and
555: $\epsilon_{\alpha\beta\gamma}$ is the fully antisymmetric tensor.
556: The quantity $h$ is the mean dissipation of helicity per
557: unit mass and time,
558: $$
559: h \equiv \nu \left<(\nabla^\alpha u^\beta)(\nabla^\alpha[\B
560: \nabla\times \B u]^\beta)\right> \ , $$
561: where repeated indices are summed upon. In the derivation below it assumed
562: that $h$ remains constant when $\nu \rightarrow 0$ in the same spirit of the
563: dissipation anomaly \cite{eyink1,eyink2,bif98c}. The first term in Eq.
564: (\ref{result}) is just the 4/5 law. The new part of result
565: (\ref{result}) can be also displayed in a form that depends on
566: $h$ alone by introducing the longitudinal and transverse
567: parts of $\B u$: the longitudinal part is $\B u_l\equiv \B r(\B
568: u\cdot \B r)/r^2$ and the transverse part is $\B u_t \equiv \B
569: u-\B u_l$. In terms of these quantities we
570: can present a ``two fifteenth law"
571: \begin{equation}
572: \left<[\delta\B u_l(\B x,\B r,t)]\cdot [\B u_t(\B r+\Bx,t)
573: \times \B u_t(\B x,t)]\right> =\frac{2}{15}h r^2 \ .
574: \label{short}
575: \end{equation}
576: We note that this result holds also when we replace $\B u$ by $\B
577: v$ everywhere.
578:
579: To derive the result (\ref{result}) we start from the correlation
580: function $J^{\alpha,\beta\gamma}(\B r)$ which is symmetric with
581: respect to exchange of the indices $\beta$ and $\gamma$ as is
582: clear from the definition. In an isotropic homogeneous medium with
583: helicity (no inversion symmetry), the most general form of this
584: object is \cite{lvo97a}:
585: \begin{eqnarray}
586: &&J^{\alpha,\beta\gamma}(\B
587: r)=a_1(r)[\delta_{\alpha\beta}r^\gamma+\delta_{\alpha\gamma}
588: r^\beta+\delta_{\beta\gamma}r^\alpha] \label{general}
589: +\tilde
590: a_1(r)[\delta_{\alpha\beta}r^\gamma+\delta_l{\alpha\gamma}
591: r^\beta-2\delta_{\beta\gamma}r^\alpha]\nonumber \\
592: &&+b_2(r)[\epsilon_{\alpha\beta\delta}
593: r^\gamma+\epsilon_{\alpha\gamma\delta} r^\beta]r^\delta
594: +a_3(r)[\delta_{\alpha\beta}r^\gamma+\delta_{\alpha\gamma}
595: r^\beta+\delta_{\beta\gamma}r^\alpha-5r^\alpha r^\beta
596: r^\gamma/r^2] \ . \nonumber
597: \end{eqnarray}
598: This general representation is invariant to the choice of
599: orientation of the coordinates.
600: Not all the coefficients are independent for incompressible
601: flows. Requiring $\partial J^{\alpha,\beta\gamma}(\B r)/\partial
602: r^\alpha=0$ leads to two relations among the coefficients:
603: $$
604: \Big({d\over dr}+{5\over r}\Big)a_3(r)={2\over 3}{d\over
605: dr}\big[a_1(r)
606: +\tilde a_1(r)\big] \ , \qquad
607: \Big({d\over dr}+{3\over r}\Big)\big[5a_1(r)-4\tilde
608: a_1(r)\big]=0 \ . $$
609: As we have two conditions relating the three coefficients
610: $a_1,\tilde a_1$ and $a_3$ only one of them is independent.
611: Kolmogorov's derivation related the rate of energy dissipation to
612: the value of the remaining unknown. Here the coefficient $b_2$
613: remains undetermined by the incompressibility constraint; it will
614: be determined by the rate of helicity dissipation.
615:
616: Kolmogorov's derivation can be paraphrased in a simple manner.
617: Begin with the second order structure function $\tilde S^{(2)}(r)\equiv
618: \la |\Bu (\B x +\B r)- \Bu(\B x)|^2 \ra$.
619: Computing the rate of change of this (time-independent) function
620: from the Navier-Stokes equations (\ref{NS}) we find
621: \begin{equation}
622: 0={\partial \tilde S^{(2)}(r)\over 2\partial t}=-{\mathcal
623: D}^{(2)}(r)-2\eb+\nu\nabla^2 \tilde S^{(2)}(r) \ , \label{bal}
624: \end{equation}
625: where ${\mathcal D}^{(2)}(r)$ stems from the nonlinear term $(\B u\cdot\B
626: \nabla)\B u$ and as a result it consists of a correlation
627: function including a velocity derivative. The conservation of
628: energy allows the derivative to be taken outside the correlation
629: function:
630: \begin{equation}
631: {\mathcal D}^{(2)}(r) \equiv {\partial\over \partial r^\beta} \langle
632: u^\alpha(\B x,t)u^\alpha(\B x+\B r,t)
633: \big[u^\beta(\Bx,t)-\!u^\beta(\B x+\B r,t)\big]\rangle \ . \label{D2}
634: \end{equation}
635: In terms of the function of Eq. (\ref{defJ}) we can write
636: \begin{equation}
637: {\mathcal D}^{(2)}(r) = {\partial\over \partial
638: r^\beta}\Big[J^{\alpha,\beta\alpha}(\B r,t)-
639: J^{\alpha,\beta\alpha}(-\B r,t)\Big] \ . \label{relate}
640: \end{equation}
641: Note that Eq. (\ref{defJ}) is written in terms of $\B v$ rather
642: than $\B u$, but using the incompressibility constraint we can
643: easily prove that Eq. (\ref{D2}) can also be identically written
644: in terms of $\B v$ rather than $\B u$. We proceed using Eq.
645: (\ref{general}) in Eq. (\ref{relate}), and find
646: \begin{equation}
647: {\mathcal D}^{(2)}(r) = 2{\partial\over \partial
648: r^\beta}r^\beta\big[5a_1(r)+2\tilde a_1(r)\big] \ . \label{D2a1}
649: \end{equation}
650: For $r$ in the inertial interval, and for $\nu\to 0$, we can read
651: from Eq. (\ref{bal}) ${\mathcal D}^{(2)}(r)=-2\eb$ and therefore
652: have the third relation that is needed to solve all the three
653: unknown coefficients. A calculation leads to
654: $$
655: a_1(r)=-2\eb/45\ , \quad \tilde a_1=-\eb/18\ ,
656: \quad a_3=0\ . $$
657: The choice of the structure function $\tilde S_2(r)$ leaves the
658: coefficient $b_2(r)$ undetermined, and another correlation
659: function is needed in order to remedy the situation. Since the
660: helicity is $\B u\cdot [\B \nabla\times \B u]$, we seek a
661: correlation function which is related to the helicity of eddys of
662: scale of $r$:
663: $$
664: T^{(2)}(r)\equiv \langle \big[\B u(\B r+\B x,t)-\B u(\B x,t)\big]
665: \nonumber \cdot \big[\B \nabla\times \B u(\B x+\B r,t)-\B
666: \nabla\times \B u(\B x,t)\big]\rangle \ . $$
667: Using the Navier-Stokes equations to compute the rate of change
668: of this quantity we find
669: \begin{equation}
670: 0={\partial T^{(2)}(r)\over 2\partial t}=-G^{(2)}(r)-2h
671: -\nu\nabla^2T^{(2)}(r) \ , \label{bal2}
672: \end{equation}
673: which is the analog of (\ref{bal}), and where
674: \begin{eqnarray}
675: G^{(2)}(r)&=&\left\{\langle\B u(\B x,t)\cdot\left[\B \nabla_r\times\left[\B
676: u(\B x+\B r,t)\times \left[\B \nabla_r \times \B
677: u(\B x+\B r,t)\right]\right]\right]\rangle\right\} \nonumber\\&+&\{{\rm
678: term}~\B r\to -\B
679: r\} \ . \label{G2}
680: \end{eqnarray}
681: The conservation of helicity allows the extraction of two
682: derivatives outside the correlation functions. The result can be
683: expressed in terms of the definition (\ref{defJ}):
684: $$
685: G^{(2)}(r)={\partial\over \partial R^\lambda}{\partial\over \partial
686: r^\kappa}
687: \epsilon_{\alpha\lambda\mu}\epsilon_{\mu\beta\nu}\epsilon_{\nu\kappa\gamma}
688: \big[J^{\alpha,\beta\gamma}(\B r)+J^{\alpha,\beta\gamma}(-\B
689: r)\big] \ . $$
690: Substituting Eq. (\ref{general}) we find
691: $$
692: G^{(2)}(r)=2{\partial^2\over \partial r^\lambda \partial
693: r^\kappa}b_2(r)\big[r^\lambda r^\kappa
694: -\delta_{\lambda\kappa}r^2\big] \ , $$
695: which is the analog of Eq. (\ref{D2a1}). Using Eq. (\ref{bal2})
696: in the inertial interval in the limit $\nu\to 0$ we find the
697: differential equation
698: $$
699: r^2{d^2 b_2(r)\over dr^2}+9r{db_2(r)\over dr} +15 b_2(r)=-{h
700: \over 2} \ . $$
701: The general solution of this equation is
702: $b_2(r) = -h / 30+\alpha_1 r^{-5}+\alpha_2 r^{-3}$.
703: Requiring finite solutions in the limit $r\to 0$ means that
704: $\alpha_1=\alpha_2=0$. Accordingly we end up with Eq.
705: (\ref{result}).
706: The moral of this example is that {\em even in isotropic and
707: homogeneous systems there exist sub-leading terms} which can
708: become dominant for specially selected objects like (\ref{short}).
709: Once anisotropy exists, there are many more (in fact infinitely
710: many) sub-leading contributions that need to be assessed
711: carefully. Similar results for slightly different correlation functions
712: have also been found in \cite{kur03,pol03}
713: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
714: \subsection{Kolmogorov's Theory for 2'nd, 4'th, and Higher Order Structure
715: Functions}
716: \label{K41forSp}
717: Unfortunately, the exact result pertaining to the 3'rd order
718: structure function is rather unique. The moment we consider 2'nd,
719: 4'th or higher order correlation functions there is no exact
720: result for the scaling exponents. Kolmogorov's 1941 theory states
721: that, for Re large enough, small-scales turbulent
722: fluctuations should recover isotropy and homogeneity (if measured
723: far enough from boundaries) and should possess universal scaling
724: properties depending only on the mean energy flux, $\eb$.
725: For homogeneous and isotropic ensembles one defines
726: \begin{equation}
727: S^{(n)}(r) \equiv \left< \delta u_{\l}^n(\Bx, \Br,t)
728: \right>\ = (\eb r)^{n \over 3}f^{(n)}({r \over L_0},{\eta
729: \over r}), \label{eq:k41a} \label{Sp}
730: \end{equation}
731: where $L_0$ and $\eta$ are the integral length scale and the
732: viscous scale, respectively. The function $f^{(n)}$ is supposed to be
733: well behaved in the limit of infinite Reynolds numbers for fixed
734: separation, $r$:
735: $\lim_{x,y \rightarrow 0} f^{(n)}(x,y) =$ const. In this limit,
736: the celebrated K41 scaling prediction for structure
737: functions in the inertial range, $ \eta \ll r \ll L_0$, follows:
738: \begin{equation}
739: S^{(n)}(r) \sim C^{(n)} (\eb r)^{\zeta^{(n)}}, \qquad \mbox{with}\;\;
740: \zeta^{(n)} = \frac{n}{3}. \label{eq:k41}
741: \end{equation}
742: In (\ref{eq:k41}) the constants
743: $C^{(n)}$ depend only on the large scale
744: properties. Because of stationarity, the mean energy flux in
745: Eq. (\ref{eq:k41}) can be equally taken to be the mean energy input
746: or the mean energy dissipation.
747:
748: Kolmogorov's theory goes beyond the scaling prediction
749: (\ref{eq:k41}). For example, any non-vanishing $p$th order
750: structure functions, including purely transversal
751: and mixed
752: longitudinal-transversal velocity increments, must possess the
753: same scaling exponents:
754: \begin{equation}
755: S^{(n,m)}(r) \equiv \left< \delta u_{\l}^{n}(\Br) \delta
756: u_t^{m}(\Br) \right>
757: \sim C^{(n,m)}(\eb r)^{\frac{(n+m)}{3}}, \label{eq:mixed}
758: \end{equation}
759: where $p=n+m$, $ \delta \Bu_t(\Br) \equiv \delta \Bu(\Br)- \delta
760: u_{\ell}(\Br)\hat{\Br}$,
761: and $\delta u_t(\Br)$ is one of the components of the two-dimensional
762: transverse velocity difference.
763: Notice that due to the assumption of isotropy, only even combinations
764: of transversal increment in (\ref{eq:mixed}) have a non vanishing
765: average. It is also not difficult to extend the K41 reasonings to
766: describe also correlation functions at the viscous scales,
767: i.e. observables based on gradients statistics
768: \cite{fri90,ben91}.
769: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
770: \subsection{Experimental Difficulties with the Isotropic Theory}
771: \label{expcheck}
772:
773: On the whole, experimental tests of Kolmogorov's theory ran into increasing
774: difficulties
775: when the data were analyzed with greater detail. The first systematic
776: attempt to check the isotropic scaling (\ref{eq:k41}) for high
777: Re number turbulence was
778: \cite{ans84}. These authors performed
779: a high statistical test of K41 theory by going beyond
780: the usual two-point correlations. They measured structure
781: functions of higher order, reaching good evidence that there exist
782: {\it anomalous} deviations from the scaling exponents (\ref{eq:k41}). Their
783: data substantiate a power-law behavior with $\zeta(n) \ne n/3$. At
784: that time, and for many year later, the situation was very
785: controversial both theoretically and experimentally. There appeared
786: many claims that the observed deviations were due to sub-leading
787: finite-Reynolds effects. One should not underestimate the difficulties
788: of getting reliable estimates of the scaling exponents. First, one
789: must expect finite Reynolds numbers corrections which may {\it
790: strongly } reduce the inertial range where scaling laws are
791: expected or introduce anisotropic corrections to the isotropic K41
792: predictions. Both effects are usually present in all experiments and
793: numerical simulations. Nowadays, state-of-the-art experiments of
794: turbulence in controlled geometries reach a maximum Re numbers
795: measured on the gradient scales, $\lambda^{-1} = \langle | \B\nabla \Bu|
796: \rangle
797: /\langle |\Bu| \rangle$, of $R_{\lambda}
798: \sim 5000$
799: where $R_{\lambda} = \frac{\lambda U}{\nu}$ and $U$ is the typical
800: large scale velocity. In atmospheric flows $R_{\lambda}$ can be
801: as high as $R_{\lambda }\sim 20000$ but at the expense of high
802: anisotropy. More complex is the situation of DNS where the best resolution
803: ever reached up to now is
804: $4096^3$ \cite{kan03}, corresponding to a $R_{\lambda}
805: \sim 1100$. DNS allow a minimization of the anisotropic corrections,
806: by implementing periodic boundary conditions
807: and fully isotropic forcing, something which is not
808: experimentally feasible.
809: However, also in DNS the discrete symmetries
810: induced
811: by the finite lattice-spacing do not allow for perfect isotropic
812: statistics. We thus either have high-Reynolds-numbers experiments
813: which are strongly perturbed by anisotropic effects, or DNS isotropic
814: flow at moderate Reynolds numbers. Therefore, one has to face the
815: problem of how to disentangle isotropic from anisotropic fluctuations
816: and how to extract information on the asymptotic scaling with a finite
817: --often short-- inertial-range extension. Only recently, after many
818: experimental and numerical confirmations of the results of
819: \cite{ans84},
820: the situation became clearer \cite{ben93b}. We may affirm now with
821: some degrees of certitude that the isotropic scaling exponents are
822: {\it anomalous}, the K41 prediction $\zeta(n) = n/3$ is wrong, except
823: for $n=3$ which is fixed to be $\zeta(3)=1$ by the exact $4/5$ law.
824: Moreover, the possibility to show analytically the existence of
825: anomalous scaling in turbulent advection
826: \cite{fal01}, definitely eliminated those arguments
827: supporting the {\it impossibility} to have a Re-independent
828: anomalous scaling in any hydrodynamic system. From a phenomenological point
829: of view, it is easy to extend the K41
830: theory such as to include anomalous scaling.
831: Already Kolmogorov noticed, after Landau's criticism in 1962, that it is
832: unrealistic to expect the isotropic inertial range fluctuations to
833: depend only on the {\it mean} energy dissipation, $\eb$.
834: Kolmogorov proposed in 1962 \cite{kol62} to employ the
835: coarse-grained energy dissipation over a box of size $r$,
836: \be
837: \label{eq:coarse}
838: \tilde
839: \epsilon(r,\Bx) = {1\over r^3}
840: \int_{|y| < r} d\By \eb(\Bx +\By),
841: \ee
842: to
843: match the correct dimensions of structure functions in
844: (\ref{eq:k41a}), the so-called Refined Kolmogorov Hypothesis:
845: \begin{equation}
846: S^{(n)}(r) = C^{(n)} \la \tilde \epsilon^{n\over 3}(r)\ra r^{n \over
847: 3}.
848: \label{eq:multifractal}
849: \end{equation}
850: This hypothesis connects the deviation from the K41
851: prediction, $\zeta(n) - n/3 = \tau(n/3)$, to the anomalous scaling of
852: the coarse-grained energy dissipation : $\la \tilde \epsilon^{n\over
853: 3}(r)\ra
854: \sim r^{\tau(n/3)}$.
855: Anomalous scaling of isotropic structure functions is therefore
856: connected to the multifractal properties of the three dimensional
857: energy dissipation field \cite{fri95}. It should be noted however
858: that the Refined Kolmogorov Hypothesis related inertial range scaling
859: to scaling of dissipative quantities, and delicate issues connected
860: to small distance expansions and fusion rules are being disregarded
861: here \cite{lvo96,ben98}. At any rate, the relation presented by
862: Eq.(\ref{eq:multifractal}) did not advance the calculation of the
863: scaling exponents beyond crude phenomenology.
864: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
865: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% section %%%%%%%%%%%%%%%%%%%%%%%%%
866: \subsection{Persistence of Anisotropies}
867: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
868: \label{sec:persis}
869: A central issue of K41 phenomenology is the assumption of {\it
870: return-to-isotropy} for smaller and smaller scales. Recently this
871: assumption had been put to test in experiments and simulations
872: \cite{pum96,gar98,kur00,bif01}. A useful experimental
873: set-up to test the return to isotropy is a {\it homogeneous} shear
874: flow
875: \cite{hin75} where the large-scale
876: mean-velocity has a linear profile: $\B{V} = (V_0 y,0,0)$. The shear
877: is given by ${\mathcal S}_{ij} = \partial_i V_j = \delta_{iy}
878: \delta_{jx} V_0$. We thus have
879: a homogeneous but anisotropic flow, close to the ideal case for
880: studying the influence of large scale anisotropies on the small scale
881: statistics. ``Small scales" are defined here in comparison to the
882: characteristic shear length, $L_{\mathcal S} = \eb^{1/3}/{\mathcal S}$;
883: for $ r \ll L_{\mathcal S}$ we may expect that anisotropic fluctuations
884: are sub-leading with respect to the isotropic ones. The case $ r \gg
885: L_{\mathcal S}$ is of interest in situations where the shear is very
886: intense, as very close to the walls in bounded flows. In such cases we
887: expect a dramatic change from the K41 phenomenology
888: \cite{ben02,tos99,tos00}.
889: Fortunately, it is not that difficult to design experiments or DNS
890: possessing an almost perfect linear profile with homogeneous shear
891: \cite{pum96,she00,gar98,ben02}.
892: A popular way to measure small-scales anisotropies is to focus on the
893: Re dependence of isotropic and anisotropic statistical
894: observables built in terms of velocity gradients. For example, due to
895: the symmetries of the mean flow, gradients of the stream-wise
896: component in the shear direction, $\partial_y u_x$, may have a skewed
897: distribution only due to the anisotropic fluctuations; they have a
898: symmetric PDF in a perfectly isotropic flow. A natural measure of the
899: residual anisotropy at small scales as a function of Re
900: is the mixed generalized skewness based on gradients:
901: \begin{equation}
902: M^{(2n+1)}(R_{\lambda})
903: \frac{ \la (\partial_y u_x)^{2n+1} \ra}{\la (\partial_y u_x)^2
904: \ra^{\frac{2n+1}{2}}}.
905: \label{eq:der_ske}
906: \end{equation}
907: These objects vanish in isotropic ensembles. Of course, at finite
908: Reynolds numbers one expects that the large-scale anisotropy
909: introduced by the shear still remains, even on the gradient scale.
910: Therefore, the rate of decay of (\ref{eq:der_ske}) as a function of
911: Re is a quantitative indication of the rate of decay of
912: anisotropy at small scales. In the next section we review Lumley's
913: dimensional arguments \cite{lum67} for anisotropic fluctuations, which
914: predicts:
915: \begin{equation}
916: M^{(2n+1)}(R_{\lambda}) \sim R^{-{1 \over 2}}_{\lambda},\qquad \forall \;\;
917: n.
918: \label{eq:der_ske_mf}
919: \end{equation}
920: In fact, both numerical
921: \cite{pum96,pum95} (at low Reynolds numbers) and experimental
922: tests (up to $R_{\lambda} \sim 1000$) showed a clear disagreement with
923: the dimensional prediction (\ref{eq:der_ske_mf}). For example in
924: \cite{she00} the
925: authors quote a decay in agreement with the prediction for
926: $M^{(3)}(R_{\lambda})$, an almost constant
927: behavior as a function of Re for the fifth order,
928: $M^{(5)}(R_{\lambda}) \sim O(1)$ and an {\it increasing} behavior for
929: the seventh order $M^{(7)}(R_{\lambda}) \sim R^{+0.63}_{\lambda}$ !
930: These results have cast a severe doubt
931: on the fundamental assumption of the K41 theory. Similar results, with even
932: more
933: striking contradictions with the hypothesis of the return-to-isotropy, have
934: been measured in the problem of passive scalar fluctuations, $\theta = T -
935: \la T \ra $, advected by an isotropic velocity field in the presence of a
936: mean homogeneous scalar gradient, $\B \nabla \la T \ra = (g,0,0)$.
937: The equation of motions for the passive advected field in this case
938: are: $$
939: \partial_t \theta + \Bu \cdot \B \nabla \theta =
940: g u_x + \chi \partial^2 \theta.
941: $$
942: Both experimental and numerical data show a strong
943: disagreement with the prediction that generalized skewness of
944: temperature gradients becomes smaller upon increasing
945: Reynolds and Peclet numbers \cite{war00,gib77,ant78,sre91,p94,cel01}.
946:
947: We will show below how the analysis based on SO(3) decomposition and
948: its theoretical consequences settles this puzzle of strong {\it
949: persistence of anisotropies} \cite{bif01}. In fact, contrary to
950: what appears, the K41 phenomenology with its assumption of {\it
951: return-to-isotropy} and the above experimental results are not at all
952: in contradiction (see section \ref{pers_ani}).
953: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
954: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
955: \subsection{Longitudinal and Transversal Isotropic Structure Functions}
956: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
957: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
958: \label{sec:long-tran}
959: Another debated issue concerning the K41 phenomenology and its
960: multifractal generalization (\ref{eq:multifractal}) has to do with the
961: observed discrepancies between the scaling properties of longitudinal,
962: transversal, and mixed longitudinal-transversal structure functions in
963: (supposedly) isotropic fully developed turbulence
964: \cite{she02,che97,dhr97,got02,wat99}. As previously
965: stated, K41 theory, for isotropic flows, predicts the same scaling
966: behavior, in the limit of high Re, independent of
967: the Cartesian components of the velocity increments in the structure
968: functions. For a given order $p=m+n$ only prefactors
969: in (\ref{eq:mixed}) may depend on the particular choice of $n$ and $m$.
970: Let us denote with
971: $\zeta^{(n,m)}$ the scaling exponent of the mixed structure function
972: (\ref{eq:mixed}) made of $n$ longitudinal increments and of $m$
973: transversal increments in a isotropic ensemble:
974: $$
975: S^{(n,m)}(r)\sim C^{(n,m)} r^{\zeta^{(n,m)}}.
976: $$
977: For $p<4$ the issue does not exist; due to the incompressibility
978: constraint all second and third order longitudinal or transversal
979: structure functions have the same scaling in a isotropic ensemble. For
980: $p>3$ many experiments and numerical simulations found that
981: $\zeta^{(n,m)} < \zeta^{(n',m')}$ if $n<n'$ and $m>m'$ when $n+m=n'+m'$. It
982: appears that with increasing $m$ the
983: scaling exponents reduce (the signal is more intermittent). The
984: largest difference for a structure function of order $p$ is therefore
985: achieved when we compare the purely longitudinal scaling $\zeta(p,0)$
986: with the purely transversal scaling $\zeta(0,p)$. Other experimental
987: data suggests the possibility of a {\it slow} tendency of the
988: longitudinal and transversal scaling exponents to coalesce for
989: increasing Re \cite{zho99,nou97,cam96}.
990:
991: We will argue below that the experimental measurements of different
992: exponents stems from anisotropic corrections that affect differently
993: the longitudinal and transverse components. In other words, by not removing
994: the
995: anisotropic contributions, one cannot expect pure power-law behavior.
996: The situation is more
997: complex for the analysis of data from DNS. There, one may implement
998: highly isotropic forcing and boundary conditions, such that in most
999: cases any residual anisotropic effects may safely be neglected even at
1000: moderate Re numbers. On the other hand, state-of-the-art
1001: numerical simulations are still strongly limited in the maximum
1002: Re achievable. Only very recently reliable data with
1003: high-statistics became available at resolution $1024^3$
1004: \cite{got02}, while most of the previous DNS where limited to lower
1005: resolutions. At resolution of $1024^3$ one reaches a moderate
1006: $R_{\lambda} \sim 400$, far below many experiments. Because of the
1007: consequent limited extension of the inertial range, such DNS did not
1008: resolve the puzzle of longitudinal {\it vs} transversal scaling. The
1009: numerical results oscillate between evidence for different scaling
1010: properties and for its opposite
1011: \cite{che97,got02,bor97,ker01,gro98b}. The issue is complicated
1012: by the fact that longitudinal and transversal structure functions
1013: possess different finite Re effects. For example, in
1014: \cite{got02} it was shown that structure functions
1015: of different order have different dependence on the viscous cut-off;
1016: this introduces ambiguity in defining a common inertial range where
1017: power law is expected. We thus propose that until high resolution
1018: isotropic measurements became available, all evidence for different
1019: scaling exponents for longitudinal and transverse structure functions
1020: should be considered with suspicion.
1021: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1023: \subsection{Position Dependent Scaling Exponents}
1024: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1025: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1026: \label{sec:position-dep}
1027: In some inhomogeneous
1028: simulations and experiments it was claimed that the measured scaling
1029: exponents depended on the point of measurement within the flow domain
1030: \cite{gau98,sad94,tos99,tos00,ono00,str96}. If true, such finding would
1031: deal a death blow to the idea of universality of the scaling exponents
1032: in turbulence. It should be stressed that in all the examples where
1033: such findings were reported the flow contained strong anisotropic and
1034: inhomogeneous components and/or the scaling range was not sufficient
1035: to actually present direct log-log plots for the structure functions
1036: vs. $r$. In some of these cases, scaling was extracted by using the
1037: method called ``Extended Self Similarity" (ESS) \cite{ben93b}; the use
1038: of this method can be dangerous in presence of anisotropic and
1039: inhomogeneous effects. Whenever strong anisotropies are present one
1040: has to distinguish among two scaling ranges. At scales larger than the
1041: shear length, the energy cascade mechanism of the Kolmogorov theory is
1042: overwhelmed by shearing effects \cite{ben02,tos00}. Only
1043: for scales smaller than the shear length the meaning of anisotropic
1044: corrections to the isotropic K41 scaling theory is well posed. We
1045: argue below that the reported position-dependent isotropic exponents
1046: in the latter case stem from anisotropic components which appear with
1047: different amplitudes at different points in the flow. The different
1048: ``exponents" that were measured were not real exponents but the result
1049: of a crossover between the isotropic and anisotropic corrections.
1050: Once the data is projected onto the isotropic sector the leading
1051: exponents become position independent as expected.
1052: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1053: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1054: %%%%%%%%%%%%%%%%%%%%%%%%%%%%% section 3 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1055: \section{Historical Review: Attempts at Anisotropy}
1056: \label{history-aniso}
1057: %%%%%%%%%%%%%%%%%%%%%%%%%
1058: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1059: \subsection{Bachelor's Approach}
1060: The first systematic approach to anisotropy in turbulence was
1061: suggested by G.~K.~Batchelor in 1946 \cite{bat46}. Batchelor did
1062: not attempt to describe the most general form of anisotropy in
1063: turbulence, but instead confined himself to the easier case of
1064: \emph{axisymmetric turbulence}. In axisymmetric turbulence
1065: the mean value of any function of the velocity field and its
1066: derivatives is invariant to all rotations of the axes in a given
1067: direction. Therefore, the anisotropy in axisymmetric turbulence is
1068: induced by a single direction in space. We denote this symmetry
1069: axis by the unit vector $\Bn$. Being the easiest case of anisotropic
1070: turbulence, axisymmetric turbulence was the main model for
1071: studying anisotropy in subsequent years.
1072:
1073: Batchelor used the invariant theory in order to take the anisotropy
1074: vector $\Bn$ into account in the tensor representations. His method
1075: is simple: add the vector $\Bn$ to the list of vectors in Robertson's
1076: method. For example, suppose we wish to construct an axisymmetric
1077: representation of the second-order correlation function
1078: $C^{\alpha\beta}(\Br)$ defined in (\ref{def:KA-C2}). Then, just as in
1079: the isotropic case, we create a scalar function by contracting the two
1080: indices of $C^{\alpha\beta}(\Br)$ with two arbitrary vector $\B{a}$
1081: and $\B{b}$, with the difference that now we assume that the resultant
1082: scalar function depends on the unit vector $\Bn$ as well as on the
1083: other vectors $\Br$, $\B{a}$ and $\B{b}$. We therefore look for an
1084: invariant representation of the scalar function $C(\Br, \Bn, \B{a},
1085: \B{b})$, which depends only on the different scalar products
1086: $\Br\cdot\Br$, $\Br\cdot\Bn$, $\Br\cdot\B{a}\ $,... and the various
1087: determinants $[\Br\Bn\B{a}]$, $[\Br\Bn\B{b}]\ $,... For some reason,
1088: Batchelor decided to ignore the skew-symmetric parts and considered
1089: only the scalar products. Using the fact that $C(\Br, \Bn, \B{a},
1090: \B{b})$ is a bilinear function of $\B{a}$ and $\B{b}$, Batchelor found
1091: that
1092: $$
1093: C^{\alpha\beta}(\Br) = A r^\alpha r^\beta + B \delta^{\alpha\beta} +
1094: C n^\alpha n^\beta + D n^\alpha r^\beta + E r^\alpha n^\beta \ ,
1095: $$
1096: where $A,B,C,D,E$ are functions of the amplitude $r$ and of the scalar
1097: product $\hat{\Br}\cdot\Bn$. Notice that in this expansion the number
1098: of unknowns has grown from two to five with respect to the isotropic
1099: expansion. It would have been nine, had we taken the skew-symmetric
1100: parts into account. Indeed, a prominent characteristic of anisotropic
1101: representations is that they are far more complex than their isotropic
1102: counterparts.
1103:
1104: Using this sort of representations, Batchelor was able to generalize
1105: K\'arm\'an-Howarth results to the case of axisymmetric turbulence. That
1106: is, after representing the second- and third-order correlation-functions in
1107: terms of few scalar functions, Batchelor used the solenoidal condition
1108: and the Navier-Stokes equations to derive some linear differential
1109: relations among them.
1110: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1111: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
1112: \subsection{Chandrasekhar and Lindborg's Approaches}
1113: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1114: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
1115: A somewhat more elegant approach to axisymmetric turbulence was
1116: offered a few years later by S.~Chandrasekhar
1117: \cite{cha50}. Chandrasekhar's treatment is similar to Batchelor's
1118: in following Robertson's work \cite{rob40}. Chandrasekhar
1119: took advantage of the skew-symmetric tensor
1120: $\epsilon^{\mu\alpha\beta}$ for creating a representation of
1121: solenoidal axisymmetric tensors. He noticed that the curl of an
1122: axisymmetric tensor automatically satisfies the solenoidal
1123: condition. Therefore, by representing the second- and third-
1124: correlation-functions as a curl of auxiliary tensors, Chandrasekhar
1125: automatically solved the solenoidal equations, and was left with the
1126: dynamical equations (which are derived from the Navier-Stokes
1127: equation) only. Chandrasekhar's dynamical equations are considerably
1128: simpler than those of Batchelor. Nevertheless, they are still very
1129: complicated and this, perhaps, explains why there was no serious
1130: attempt to continue Chandrasekhar's work in subsequent years.
1131:
1132: In 1995 there was another attempt to formulate the kinematics of
1133: homogeneous axisymmetric turbulence in \cite{lin95}. The
1134: representation in this paper was ``experimentally oriented'', in the
1135: sense that the scalar functions that are used can be measured directly
1136: in experiment. To accomplish that, one defines two auxiliary unit
1137: vectors (that were also used in \cite{her74}): $
1138: \Be_1(\Br) \equiv \Bn \times \Br/|\Bn\times\Br| \ , \quad \Be_2(\Br)
1139: \equiv \Be_1(\Br) \times \Bn \ $.
1140: The triplet $(\Bn, \Be_1, \Be_2)$ is an orthonormal basis of
1141: $\mathbb{R}^3$. But since it is made out of $\Br$ and $\Bn$ using
1142: invariant operations (i.e., vectorial products), it is invariant to
1143: simultaneous rotations of $\Br$ and $\Bn$, and thus it is invariant to
1144: rotations of $\Br$ alone around $\Bn$ (because in these rotations
1145: $\Bn$ remains fixed anyhow). Therefore any tensor that is built from
1146: these unit vectors and the products $\Br\cdot\Br$, $\Br\cdot\Bn$ is
1147: necessarily an axisymmetric tensor. For example, in order to represent
1148: the second-order correlation-function $C^{\alpha\beta}(\Br)$, one writes
1149: \begin{eqnarray}
1150: C^{\alpha\beta}(\Br) &=& R_1 n^\alpha n^\beta +R_2 e_2^\alpha
1151: e_2^\beta + R_3 e_1^\alpha e_1^\beta +R_4 \big[n^\alpha e_2^\beta +
1152: n^\beta e_2^\alpha\big] \nonumber \\ &&+\ Q_1 \big[n^\alpha e_1^\beta +
1153: n^\beta e_1^\alpha\big] + Q_2 \big[e_2^\alpha e_1^\beta + e_2^\beta
1154: e_1^\alpha\big] \ ,\nonumber
1155: \end{eqnarray}
1156: where the six scalar functions $R_1,{\ldots}, R_4$ and $Q_1, Q_2$ are
1157: functions of $\mu \equiv \Br\cdot\Bn$ and $\rho \equiv |\Br \times
1158: \Bn|$.
1159: Notice that this representation takes into account the
1160: skew-symmetric part of $C^{\alpha\beta}(\Br)$ using the scalar
1161: functions $Q_1$ and $Q_2$.
1162:
1163: The major advantage in this representation is that the scalar
1164: functions have an immediate interpretation in terms of measurable
1165: quantities. With respect to the example above, if $(u,v,w)$ are the
1166: velocity components in the direction of $(\Bn, \Be_2, \Be_1)$
1167: respectively, then due to the orthonormality of triplet $(\Bn, \Be_2,
1168: \Be_1)$, we get
1169: \begin{eqnarray}
1170: && R_1 = \left<u(\Bx)u(\Bx+\Br)\right> \ , \quad R_2 =
1171: \left<v(\Bx)v(\Bx+\Br)\right> \ ,\nonumber \\ && R_3 =
1172: \left<w(\Bx)w(\Bx+\Br)\right> \ , \quad R_4 =
1173: \left<u(\Bx)v(\Bx+\Br)\right> \ , \nonumber \\ && Q_1 =
1174: \left<u(\Bx)w(\Bx+\Br)\right> \ , \quad Q_2 =
1175: \left<v(\Bx)w(\Bx+\Br)\right> \ .\nonumber
1176: \end{eqnarray}
1177: Next one uses the solenoidal condition
1178: to derive linear differential relations between
1179: the scalar functions. One can also consider the triple correlation
1180: function and the velocity-pressure correlation function in the very
1181: same method. This way one derives a representation for the dynamical
1182: equation of $C^{\alpha\beta}(\Br)$. Mathematically speaking, this
1183: representation is no better than Batchelor's representation, and may
1184: even be considered worse than Chandrasekhar's. This is because there
1185: is no reduction in the number of scalar functions and there is no
1186: simplification in the resulting equations. The motivation for this
1187: representation is experimental without compelling physical or
1188: mathematical contents.
1189: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1190: \subsection{Case Specific Approaches}
1191: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1192: The above mentioned works can be viewed as
1193: \emph{systematic} attempts to deal with the problem of anisotropic
1194: turbulence, where a general method to describe the anisotropic (or,
1195: more exactly, axisymmetric) quantities in turbulence is suggested. In
1196: that respect they differ from most research that followed on the
1197: subject, which was usually confined to a particular model or a
1198: specific problem related to anisotropy.
1199: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1200: \subsubsection{Temporal Return to Isotropy}
1201: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1202: One such problem is the
1203: temporal return to isotropy in which one tries to understand the
1204: mechanisms that drive a decaying turbulence which is initially
1205: anisotropic into being statistically isotropic. As in the works of
1206: Batchelor and Chandrasekhar, the statistics was usually assumed to be
1207: spatially homogeneous to simplify the problem. The theoretical
1208: attempts to explain this phenomenon can be roughly divided into two
1209: groups. The first group, which was initiated by Rotta
1210: \cite{rot51} in 1951, consists of attempts to model the decay of
1211: anisotropy using \emph{one-point} closures. In that framework, one
1212: usually considers the dynamical equation of the Reynolds stress, which
1213: is the \emph{same-point} correlation-function of the velocity field
1214: $
1215: C^{\alpha\beta}(t) \equiv \left<u^\alpha(\Bx,t) u^\beta(\Bx,t)\right> \ .
1216: $
1217: In a homogeneous, decaying turbulence, this correlation obeys the following
1218: equation
1219: $$
1220: \partial_t C^{\alpha\beta} = -\eb \phi^{\alpha\beta}
1221: - \frac{2}{3}\eb\delta^{\alpha\beta} \ ,
1222: $$
1223: with
1224: \begin{equation}
1225: \label{eq:return-to}
1226: \eb \phi^{\alpha\beta} \equiv \left<u^\alpha\partial^\beta p\right>
1227: + \left<u^\beta\partial^\alpha p\right>
1228: + 2\left[\nu\left<\partial_\mu u^\alpha \partial^\mu u^\beta\right> -
1229: \frac{1}{3}\eb\delta^{\alpha\beta}\right] \ .
1230: \end{equation}
1231: Notice that $C^{\alpha\beta}$ is a second-rank $\Br$-independent
1232: tensor that contains an isotropic part (which is its trace) and an
1233: anisotropic (traceless) part. This explains the motivation behind
1234: the definition of $\phi^{\alpha\beta}$, which is to capture the
1235: anisotropic part of the decay rate of $C^{\alpha\beta}$. To solve
1236: \Eq{eq:return-to}, one must model the $\phi^{\alpha\beta}$
1237: tensor. Usually this has been done on a phenomenological basis. A
1238: systematic treatment to this problem was offered by in
1239: \cite{lum77}. In that paper, the authors suggested that
1240: $\phi^{\alpha\beta}$ should depend on time implicitly through
1241: $C^{\alpha\beta}$, $\eb$ and $\nu$. Additionally, since
1242: $\phi^{\alpha\beta}$ is a dimensionless tensor, it must depend only
1243: on dimensionless parameters. There are six such independent
1244: dimensionless quantities. The authors chose to represent them in a
1245: way which isolates the property of anisotropy from other properties,
1246: and form the tensor
1247: $$
1248: b^{\alpha\beta} \equiv C^{\alpha\beta}/q^2 -
1249: \frac{1}{3}\delta^{\alpha\beta}
1250: \ , \quad q^2 = C^{\alpha \alpha} = \left<\Bu^2(\Bx)\right> \ .
1251: $$
1252: The tensor $b^{\alpha\beta}$ is proportional to the anisotropic, traceless
1253: part of $C^{\alpha\beta}$ and hence contains five independent components.
1254: It is often denoted as the ``Reynolds stress anisotropy'' or simply as the
1255: ``anisotropy tensor''. This tensor has become a central measure for
1256: anisotropy in turbulence and has been used extensively in experimental and
1257: numerical analysis of anisotropy in turbulence. The sixth component was
1258: defined to be proportional to the isotropic part of $C^{\alpha\beta}$ (the
1259: energy) by
1260: $ R_l \equiv \frac{q^4}{9\eb\nu}
1261: $. With these dimensionless quantities, $\phi^{\alpha\beta}$ can be written
1262: as
1263: $
1264: \phi^{\alpha\beta} = \phi^{\alpha\beta}(\B{b}, R_l) \ $.
1265: They further simplified that expression by noticing that
1266: if $\phi^{\alpha\beta}$ depends solely on $\B{b}$ and $R_l$ then it must
1267: depend on them in an \emph{isotropic} manner, since any anisotropic
1268: dependence necessarily means that $\phi^{\alpha\beta}$ also depends on
1269: the boundary conditions. To represent this isotropic dependence
1270: explicitly, they used the invariant theory \cite{lum70} and
1271: introduced the second- and third- principal-invariants of the
1272: traceless tensor $\B{b}$:
1273: $$
1274: II \equiv Tr[\B{b}^2] \ , \quad III \equiv Tr[\B{b}^3] \ .
1275: $$
1276: According to the invariant theory, $\phi^{\alpha\beta}$ can be most
1277: generally written as
1278: \begin{equation}
1279: \phi^{\alpha\beta} = \beta(II,III, R_l) b^{\alpha\beta} +
1280: \gamma(II,III, R_l) \Big[b^{\alpha\mu}b^{\mu\beta} -
1281: \frac{1}{3}II\delta^{\alpha\beta}\Big] \ ,
1282: \end{equation}
1283: and the problem is reduced to determining the functional form of
1284: $\beta(II,III,R_l)$ and $\gamma(II,II,R_l)$. Based on this formalism,
1285: there have been many attempts to model the functions $\beta(II,III,
1286: R_l)$ and $\gamma(II,III, R_l)$ to match experimental results
1287: \cite{chu95}. For example, Rotta's model is considered a linear
1288: model for the anisotropy decay because he used
1289: $$
1290: \beta(II,III,R_j) = C_1 \approx 3.0 \ , \qquad
1291: \gamma(II,III,R_j) = 0 \ .$$
1292: Consequently, the decay of the anisotropy tensor is given by
1293: $$
1294: \partial_t b^{\alpha\beta} = -(\eb/q^2)(C_1-2)b^{\alpha\beta} \,
1295: $$
1296: which is a linear equation in $b^{\alpha\beta}$, provided that the
1297: isotropic quantities $\eb, q^2$ are independent of
1298: $b^{\alpha\beta}$. This sort of equation predicts that
1299: $b^{\alpha\beta}(t)$ is proportional to $b^{\alpha\beta}(t=0)$ and
1300: that every component of the tensor decays at the same
1301: rate. Experimentally however linearity is not supported. For example,
1302: it has been observed experimentally that the return to isotropy is
1303: relatively rapid, at least at the beginning of the process, when the
1304: invariant $III$ is negative, whereas the return to isotropy is fairly
1305: slow in the case where the invariant is positive \cite{gen80}.
1306:
1307: The other line of research that was used to study the problem of the
1308: return-to-isotropy consists of attempts to model the decay with
1309: \emph{two-point closures}. In these models one considers the different
1310: correlation functions of the velocity field across a separation vector
1311: $\Br$, instead of using the same-point correlations as in the
1312: one-point closures. The mathematical structure here is usually much
1313: more complicated than that of the one-point closures, but in return,
1314: the two-points models often provide a deeper understanding of the
1315: physics involved.
1316:
1317: An important example for such a model is given by \cite{her74}. In
1318: this work the author used the direct-interaction approximation (DIA)
1319: \cite{kra59} to study the decay of an axisymmetric turbulence into
1320: isotropy. The DIA is a well-known truncation of the renormalized
1321: perturbation theory for turbulence. The perturbation is done in the
1322: interaction strength parameter (which is set to unity in the end), and
1323: is truncated at the second-order - i.e., at the direct-interaction
1324: terms. The calculations are done in Fourier space. They result in two
1325: coupled equations for the time evolution of the two-time, second-order
1326: correlation-function $C^{\alpha\beta}(\Bk,t,t') \equiv \left<
1327: u^\alpha(\Bk,t)u^\beta(-\Bk,t')\right>$ and the response
1328: function $G^{\alpha\beta}(\Bk,t,t')$. The latter is defined as the
1329: average of the change in the velocity field at time $t$ as a result of
1330: an infinitesimal change in the forcing at time $t'$.
1331: The equations for
1332: $C^{\alpha\beta}(\Bk,t,t')$ and $G^{\alpha\beta}(\Bk,t,t')$ determine
1333: their time evolution. They are nonlinear, nonlocal
1334: integro-differential equations and are therefore very hard to deal
1335: with. In the isotropic case the equations can be considerably
1336: simplified by noting that both $C^{\alpha\beta}(\Bk,t,t')$ and
1337: $G^{\alpha\beta}(\Bk,t,t')$ must satisfy the solenoidal condition,
1338: which in Fourier space means that both tensors must vanish once we
1339: contract any of their indices with the vector $\Bk$. It is easy to see
1340: that under such a condition, $C^{\alpha\beta}(\Bk,t,t')$ and
1341: $G^{\alpha\beta}(\Bk,t,t')$ can be represented in terms of one scalar
1342: function:
1343: \begin{equation}
1344: C^{\alpha\beta}(\Bk,t,t') = c(k,t,t')D^{\alpha\beta}(\hat{\Bk}) \ ,
1345: \quad
1346: G^{\alpha\beta}(\Bk,t,t') = g(k,t,t')D^{\alpha\beta}(\hat{\Bk}) \ ,
1347: \end{equation}
1348: where
1349: \begin{equation}
1350: \label{eq:projection-D}
1351: D^{\alpha\beta}(\hat{\Bk}) \equiv \delta^{\alpha\beta} -
1352: \hat{k}^\alpha\hat{k}^\beta \ .
1353: \end{equation}
1354: When turning to the axisymmetric case, the representation of the tensors
1355: become much more complex. Instead of one scalar function for each tensor,
1356: two functions must be used corresponding to the two scalar functions that
1357: where used by Batchelor \cite{bat46} and Chandrasekhar \cite{cha50} to
1358: describe the second-order correlation-function in real space. One uses
1359: the separating vector $\Bk$ and the anisotropy unit-vector $\Bn$ to create
1360: two unit vectors which are orthogonal to $\Bk$:
1361: \begin{equation}
1362: \Be_1(\Bk) \EqDef \Bk \times \Bn/|\Bk\times\Bn| \ , \quad
1363: \Be_2(\Bk) \EqDef \Bk\times(\Bk \times \Bn)/|\Bk\times(\Bk\times\Bn)| \ .
1364: \end{equation}
1365: With these vectors, $C^{\alpha\beta}(\Bk, t,t')$ was written as
1366: \begin{equation}
1367: \label{eq:Herring-rep}
1368: C^{\alpha\beta}(\Bk, t, t') = c_1(\Bk,t,t')\Be_1^\alpha\Be_1^\beta +
1369: c_2(\Bk,t,t')\Be_2^\alpha\Be_2^\beta \ ,
1370: \end{equation}
1371: and $G^{\alpha\beta}(\Bk,t,t')$ was written in a similar way using the
1372: functions $g_1(\Bk,t,t')$ and $g_2(\Bk,t,t')$. To parameterize the angular
1373: dependence of the scalar functions, one expands the four scalar
1374: functions in terms of Legendre polynomials,
1375: $$
1376: c_i(\Bk,t,t') = \sum_j c_{i,j}(k,t,t')P_j(\hat{\Bk}\cdot\Bn)
1377: \ ,\quad
1378: g_i(\Bk,t,t') = \sum_j g_{i,j}(k,t,t')P_j(\hat{\Bk}\cdot\Bn) \ ,
1379: $$
1380: obtaining an infinite set of coupled equations for the infinite set of
1381: functions $c_{i,j}(k,t,t')$ and $g_{i,j}(k,t,t')$. These
1382: equations were solved numerically after truncating all the $j>0$
1383: part of the expansion. Doing so,one finds the time evolution of
1384: $c_{i,0}(k,t,t')$ and $g_{i,0}(k,t,t')$ and connects them to the
1385: physical observables of the one-point closures, such as Rotta's
1386: constant. The conclusions (partially numerical and partially obtained
1387: after a long series of uncontrolled approximations) were that the
1388: return-to-isotropy is much stronger at small scales (large $k$) and
1389: that in some classes of initial conditions, the return-to-isotropy is
1390: indeed a linear phenomenon.
1391:
1392: This calculation was soon revised in \cite{sch76}. In this
1393: paper, the authors compared the DIA calculation to the results of a
1394: numerical simulation of a homogeneous and axisymmetric
1395: turbulence. This time, however, the Legendre polynomials expansion in
1396: the DIA calculation was extended to include also the $j=2$
1397: components. Their conclusions were that the DIA calculation that
1398: included the $j=2$ parts were in a good agreement with the numerical
1399: simulation, especially the $j=0$ parts, provided that the initial
1400: anisotropy was small. Additionally, the authors found that the
1401: previous calculation \cite{her74}, which considered the $j=0$ part
1402: only, was quite inadequate to describe the process of
1403: return-to-isotropy - even in the case of weak anisotropy.
1404:
1405: Other attempts to study the problem of return-to-isotropy in an
1406: axisymmetric turbulence used two-point closures. For example,
1407: eddy-damped quasinormal Markovian approximation (EDQNM) \cite{ors70,ors77}
1408: have been used in \cite{nak87}. In this closure scheme,
1409: one approximates the fourth order cumulants of the velocity field by a
1410: linear damping term of the third-order correlation-function of the
1411: velocity field (the eddy dumping). Additionally, a
1412: ``Markov'' assumption is used that allows one to integrate
1413: the history integral in the equations and retain an equation for the
1414: \emph{same time} second-order correlation-function
1415: $C^{\alpha\beta}(\Bk,t)$. As in the DIA model, this equation is
1416: formulated in Fourier space and is both nonlinear and nonlocal. To
1417: parameterize the axisymmetric correlation function in Fourier space,
1418: the authors used the following representation
1419: $$
1420: C^{\alpha\beta}(\Bk, t)
1421: =
1422: \frac{1}{4\pi k^2}\big[D^{\alpha\beta}(\hat{\Bk})E(k,\mu,t) + n^\tau
1423: n^\sigma D^{\alpha\tau}(\hat{\Bk})D^{\beta\sigma}(\hat{\Bk})
1424: F(k,\mu,t)\big] \ ,
1425: $$
1426: where here $ \mu \equiv \Bk\cdot\Bn $ and $D^{\alpha\beta}(\hat{\Bk})$
1427: is defined in \Eq{eq:projection-D}. This representation incorporates
1428: the solenoidality condition and is more elegant than
1429: (\ref{eq:Herring-rep}) in the sense that the isotropic case is very
1430: easily recovered once we set $F(k,\mu,t)=0$ and let $E(k,\mu,t)$
1431: become $\mu$-independent. Plugging this expansion to the dynamical
1432: equation of $C^{\alpha\beta}(\Bk,t)$, the authors obtained two coupled
1433: equations for $E(k,\mu,t)$ and $F(k,\mu,t)$ which they solved
1434: numerically for ``medium'' and ``strong'' anisotropic initial
1435: conditions. Their results indicate that in the medium anisotropy
1436: cases, Rotta's constant approaches a constant of the order of unity,
1437: qualitatively agreeing with Rotta's model and with the results
1438: \cite{her74}. On the other hand, in the strong anisotropy case this
1439: constant does not show any saturation, indicating the failure of
1440: Rotta's model. Additionally, their results support the idea that the
1441: decay isotropy strengthens when the $III$ invariant is negative.
1442: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1443: \subsection{Dimensional Analysis in the Presence of Strong Shear}
1444: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1445: \label{sec:lumley}
1446: An important discussion of the effects of strong shear on the energy
1447: spectrum was presented in \cite{lum67}. In this paper the author
1448: included anisotropic corrections in to the K41 framework extending
1449: the phenomenological dimensional reasonings leading to
1450: (\ref{eq:k41a}). He considered the dependence on {\it anisotropic}
1451: mean observables, like the large-scale Shear proportional to the
1452: large-scale mean gradient: ${\mathcal S} \propto
1453: \partial \la V \ra$:
1454: $$
1455: S^{(n)}(\Br) = (\eb r)^{n \over 3}f^{(n)}({r \over L_0},{\eta \over r},
1456: {\mathcal S}).
1457: %\case{1}{2}ill[eq:lumley]
1458: $$
1459: By further assuming that anisotropic corrections are ``small'' and
1460: {\it analytic} in the intensity of the shear ${\mathcal S}$, he proposed
1461: the following form for the anisotropic correction to the isotropic
1462: two-point longitudinal structure functions, in the inertial range
1463: \cite{lum67}:
1464: \begin{equation}
1465: S^{(2)}(\Br) \sim C^{(2)}(\eb r)^{2 \over 3} + D^{(2)}(\hat{\Br}){\mathcal
1466: S} r^{4
1467: \over 3}
1468: \label{eq:Lumley}
1469: \end{equation}
1470: where the coefficient $D^{(2(}(\hat{\Br})$ takes into account the
1471: dependence on the direction $\hat{\Br}$ in
1472: the anisotropic term. The counterpart of (\ref{eq:Lumley}) for the
1473: spectrum and co-spectrum in Fourier space is:
1474: \begin{equation}
1475: \la k^2 u_i(\Bk)u_l(-\Bk) \ra \sim k^{-{5\over 3}}(\delta_{il}
1476: -{k_ik_l
1477: \over k^2}) + A_{il} k^{-{7\over 3}}
1478: \label{eq:k41F}
1479: \end{equation}
1480: where the first term on the RHS is the isotropic K41 scaling and the
1481: second term is the anisotropic contribution with $A_{il}$ being a
1482: traceless matrix depending on the details of the large scale shear.\\
1483: In the past, most of the measurements of the anisotropic contributions to
1484: $S^{(2)}(r)$
1485: concentrated on the Fourier representation (\ref{eq:k41F}),
1486: \cite{sad94,tav81,sou95,lum65}.
1487: In \cite{sad94} the authors showed
1488: that the prediction (\ref{eq:Lumley} \ref{eq:k41F})
1489: is well verified in a wind tunnel
1490: flow. Later, many other experiments have confirmed this result in
1491: different experimental situations (see for example the recent results
1492: for an homogeneous shear in \cite{she00}). Only recently, a more extensive
1493: study of anisotropies has been carried out, considering also
1494: higher order statistical objects \cite{she00,she02,kur00,ara98}. The
1495: situation became immediately less clear:
1496: the prediction (\ref{eq:k41F}) is not the end of the story (see below the
1497: section on anomalous scaling for anisotropic fluctuations).
1498: We show later that in the jargon of the SO(3) decomposition the
1499: anisotropic part of a spherically averaged and solenoidal second-rank
1500: tensor is made from $j=2$ contribution only,
1501: for this reason the
1502: dimensional analysis is often viewed as predicting a $4/3$ exponent
1503: for the $j=2$ sector of the second-order structure function. This result
1504: was later derived by several authors in terms of Clebsch variables,
1505: but again by dimensional reasoning
1506: \cite{yak94,gro94,fal95}. Another, more systematic attempt
1507: to derive the scaling behavior of the second order structure function
1508: in a weakly anisotropic turbulent flow was presented in \cite{gro00}
1509: within a variable scale mean field theory. In that paper the authors
1510: reached the conclusion that all anisotropic contribution to the second order
1511: structure function must scale $\sim r^{4/3}$. To reach this result the
1512: authors
1513: had to simplify the tensorial structure of the equations for the second
1514: order correlation functions; we argue below that this uncontrolled
1515: simplification
1516: biased the estimate of the anisotropic exponents.
1517: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% CHAPTER %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1518: \section{The Modern Approach to Anisotropy}
1519: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% CHAPTER %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1520: \label{chap:modern}
1521: In the past 10 years, the subject of anomalous scaling in turbulence
1522: has gained a great deal of attention, as it became more and more
1523: accepted that in the infinite Re limit, the scaling
1524: exponents of the structure functions in the inertial range do not
1525: conform with the classical prediction of the Kolmogorov theory.
1526: The numerical values of these exponents, as
1527: well as the physical mechanism which is responsible for the anomalous
1528: scaling, have been the target of an extensive experimental, numerical,
1529: and theoretical research.
1530:
1531: On the theoretical side, important progress was made by studying
1532: Kraichnan's model of passive scalar advection \cite{kra68}. This model
1533: describes the advection of a passive scalar field by a synthetic,
1534: solenoidal velocity field with a Gaussian, white-in-time statistics. The
1535: linearity of the equations for the passive scalar field and
1536: the white-in-time statistics of the velocity field make it possible to
1537: write down a closed set of equations for the same-time correlation
1538: functions of the passive scalar \cite{kra68}. In
1539: \cite{gaw95,che95}, it was shown that
1540: the solution of these equations can lead to anomalous scaling. The
1541: key point is that the homogeneous solutions of these equations are
1542: scale invariant with nontrivial anomalous scaling exponents, which
1543: are different from the dimensional scaling exponents that
1544: characterize the inhomogeneous, ``forced'' solutions. Being usually
1545: smaller than the dimensional scaling exponent, the anomalous
1546: exponents dominate the small scales statistics of the passive scalar
1547: field. The homogeneous solutions are commonly referred to as ``zero
1548: modes'', and have been calculated to first order perturbatively in
1549: refs.~\cite{gaw95,che95} for the fourth order structure
1550: function and for all even structure functions in
1551: ref.~\cite{ber96}. Exact computer assisted calculations of the
1552: exponents of the third order structure functions were presented in
1553: \cite{gat}. Besides suggesting an elegant mechanism for anomalous
1554: scaling, Kraichnan's model also provided an example in which the
1555: scaling of the anisotropic parts of structure functions is different
1556: from the isotropic scaling. In the paper \cite{fai96} it was
1557: shown how such a thing can happen, by expanding the second-order
1558: structure function of the passive field in terms of spherical
1559: harmonics $Y_{j,m}(\hat r)$. It was found that this expansion
1560: leads to a set of decoupled $j$-dependent equations for the expansion
1561: prefactors. These equations can be easily solved by a power law whose
1562: exponent is an increasing function of $j$. These exponents are
1563: universal in the sense that they are independent of the forcing and
1564: boundary
1565: conditions.
1566:
1567: The authors in \cite{fai96} also noticed that the fact that the anisotropic
1568: exponents are
1569: higher than the isotropic exponent neatly explains the isotropization of
1570: the statistics as smaller and smaller scales are probed.
1571: Based on this example, it was suggested in
1572: \cite{lvo96c} that a similar mechanism may exist in a Navier-Stokes
1573: turbulence. The authors expanded the second-order structure function
1574: in terms of spherical harmonics
1575: \begin{equation}
1576: S^{(2)}(\B r) = \sum_{j,m} S^{(2)}_{j m}(r) Y_{j m}(\hat{\Br}) \ ,
1577: \label{SO3dec}
1578: \end{equation}
1579: and argued that in the case of weak anisotropy, one can linearize the
1580: equations for the anisotropic corrections of the second-order structure
1581: function around the isotropic solution. In such a case, the kernel of the
1582: linearized equation is invariant under rotations (isotropic), and as a
1583: result the equations for the different $(j,m)$ components decouple, and
1584: are $m$-independent - much as in the case of the second-order structure
1585: function in Kraichnan's model. In a scale-invariant situation, this leads
1586: to anisotropic, $j$-dependent exponents
1587: $$
1588: S^{(2)}_{j m}(r) \sim (\eb r)^{2/3}\left(\frac{r}{L}\right)^{\delta_j}
1589: \sim r^{\zeta^{(2)}_j} \ . $$
1590: If one accepts that homogeneous turbulence enjoys universal
1591: statistics in the inertial range, then the kernel of the above
1592: linearized equation is universal, and consequently so are the
1593: anisotropic scaling exponents $\zeta^{(2)}_j$. All of these
1594: statements could not have been proved rigorously (and still
1595: haven't been proved rigorously), yet they offered a new approach
1596: to understanding anisotropy in turbulence, an approach that is
1597: explored in the rest of this review.
1598: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1599: %
1600: \subsection{Mathematical Framework}
1601: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1602: %42
1603: Experiments in fluid turbulence are usually limited to the measurement
1604: of the velocity field at one single spatial point as a function of
1605: time. This situation has begun to improve recently, but still much of
1606: the analysis of the statistical properties of Navier-Stokes turbulence
1607: is influenced by this tradition: the Taylor hypothesis \cite{tay38} is
1608: used to justify the identification of velocity differences at different
1609: times with differences of longitudinal velocity components across a
1610: spatial length scale $r$. Most of the available statistical
1611: information is therefore about properties of longitudinal two-point
1612: differences of the Eulerian velocity field and their moments. Recent
1613: research \cite{lvo96} has pointed out the advantages of
1614: considering not only the longitudinal structure functions, but
1615: tensorial multi-point correlations of velocity field differences
1616: $$
1617: {\B w}({\B x},{\B x}^{\prime },t)\equiv {\B u}({\B x}^{\prime },t)-{\B
1618: u}({\B x},t), $$
1619: given by
1620: \begin{equation}
1621: F^{\alpha_1\dots\alpha_n}({\B x}_{1},{\B x}_{1}^{\prime},t_{1};;\dots ;{\B
1622: x}_{n},{\B x}_{n}^{\prime },t_{n})
1623: =\langle w^\alpha_1({\B x}_{1},{\B x}_{1}^{\prime },t_{1})
1624: \dots w^\alpha_n({\B x}_{n},{\B x}_{n}^{\prime},t_{n})\rangle
1625: \label{defFtilde}
1626: \end{equation}
1627: where all the coordinates are distinct. When the coordinates fuse to
1628: yield time-independent structure functions depending on one separation only,
1629: these are the so-called tensorial structure functions, denoted as
1630: \begin{equation}
1631: S^{\alpha_1 \dots \alpha_n}({\B r})\equiv \langle \lbrack u^{\alpha_1
1632: }({\Bx}+{\Br})-u^{\alpha_1 }({\B x})]
1633: \cdots[u^{\alpha_n }({\B x}+{\B r})-u^{\alpha_n }({\B x})]
1634: \rangle \ . \label{Sn}
1635: \end{equation}
1636: Needless to say, the tensorial information is partially lost in the
1637: usual measurements conducted at a single point. One of the main
1638: stresses of the present review is that keeping as much of tensorial
1639: information as possible can help significantly in disentangling
1640: different scaling contributions to the statistical objects. Especially
1641: when anisotropy implies different tensorial components with possible
1642: different scaling exponents characterizing them, careful control of
1643: the various contributions is called for.
1644:
1645: To understand why irreducible representations of the symmetry group
1646: may have an important role in determining the form of correlation
1647: functions, we need to discuss the equations of motion which they
1648: satisfy. We shall show that the isotropy of the Navier-Stokes equation
1649: and the incompressibility condition implies the isotropy of the
1650: hierarchical equations which the correlation functions satisfy. We
1651: will use this symmetry to show that every component of the general
1652: solution with a definite behavior under rotations (i.e., components of
1653: a definite {\em irreducible representation} of the $SO(3)$ group) has
1654: to satisfy these equations by itself - independently of components
1655: with different behavior under rotations. This ``foliation'' of the
1656: hierarchical equations may possibly lead to different scaling
1657: exponents for each component of the general solution which belong to a
1658: different $SO(3)$ irreducible representation.
1659: %%%%%%%%%%%%%%%%%%%%%%%
1660: %%%%%%%%%%%%%%%%%%%%%%%
1661: \subsection{Tensorial Correlation Functions and $SO(3)$ Irreducible
1662: Representations: General Theory}
1663: \label{sec:basis}
1664: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1665: %%%%%%%%%%%%%%%%%%%%%%%
1666: The physical objects that we deal with are the moments of the velocity
1667: field at different space-time locations. In this section we follow Ref.
1668: \cite{ara99b}
1669: which suggests a way of decomposing these objects into components with a
1670: definite
1671: behavior under rotations \cite{ara99b}. It will follow that components with
1672: different behavior under rotation are subject to different dynamical
1673: equations, and therefore, possibly, scale differently. Essentially, we
1674: are about to describe the tensorial generalization of the well-known
1675: procedure of decomposing a scalar function $\Psi ({\bf r})$ into
1676: components of different irreducible representations using the
1677: spherical harmonics:
1678: \begin{equation}
1679: \Psi ({\bf r})=\sum_{j,m}a_{jm}(r)Y_{jm}(\hat{{\bf r}})\ .
1680: \label{scalar-case}
1681: \end{equation}
1682: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1683: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1684: \subsubsection{Formal Definition}
1685: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1686: Consider the correlation function ${\bf F}^{(n)}$ of
1687: Eq.~(\ref{defFtilde}).
1688: This $n$-rank tensor is a
1689: function of $2n$ spatial variables and $n$ temporal
1690: variables. It transforms as a {\em tensor field}: if ${\bf F}^{(n)}$ is
1691: measured in two frames $I$ and $\overline{I}$ which are connected by
1692: the spatial transformation (say, a rotation)
1693: $
1694: \overline{x}^{\alpha }=\Lambda ^{\alpha \beta }x^{\beta }
1695: $
1696: then, the measured quantities in each frame will be connected by the
1697: relation:
1698: \begin{eqnarray}
1699: &&\overline{{ F}}^{\alpha_{1}\ldots \alpha _{n}}(\overline{{\bf x}}
1700: _{1},\overline{{\bf x}}_{1}^{\prime },\overline{t}_{1};\ldots ;\overline{
1701: {\bf x}}_{n},\overline{{\bf x}}_{n}^{\prime },\overline{t}_{n})
1702: =\Lambda ^{\alpha _{1}\beta _{1}}\cdots \Lambda ^{\alpha_n\beta _{n}}{
1703: F}^{\beta_{1}\ldots \beta_{n}}({\bf x}_{1}, {\bf x}_{1}^{\prime
1704: },t_{1};\ldots ;{\bf x}_{n},{\bf
1705: x}_{n}^{\prime },t_{n})
1706: \nonumber \\
1707: &&=\Lambda ^{\alpha_{1}\beta_{1}}\cdots \Lambda ^{\alpha
1708: _{n}\beta _{n}}{ F}^{\beta _{1}\ldots \beta _{n}}(\LL\overline{
1709: {\bf x}}_{1},\LL\overline{{\bf x}}_{1}^{\prime },\overline{t}_{1};\ldots
1710: ;\LL
1711: \overline{{\bf x}}_{n},\LL\overline{{\bf x}}_{n}^{\prime
1712: },\overline{t}_{n}). \label{tensor-trans}
1713: \end{eqnarray}
1714: We see that as we move from one frame to another,
1715: the {\em functional form}
1716: of the tensor field changes. We want to classify the different tensor fields
1717: according to the change in their functional form as we make that move. We
1718: can omit the time variables from our discussion since under rotation they
1719: merely serve as parameters. We thus define $\B T(\{\B x_i\})\equiv
1720: \B F (\{\B x_i\},\{t_i=0\})$.
1721: Consider coordinate transformations which are pure rotations. For such
1722: transformations we may simplify the discussion further by separating the
1723: dependence
1724: on the amplitude of ${\bf x}_{i}$ from the dependence on the directionality
1725: of ${\bf x}_{i}$:
1726: $$
1727: T^{\alpha _{1}\ldots \alpha _{n}}({\bf x}_{1},\ldots ,{\bf x}_{p}) \\
1728: =T^{\alpha _{1}\ldots \alpha _{n}}(x_{1},\ldots ,x_{p};\hat{{\bf x}}
1729: _{1},\ldots ,\hat{{\bf x}}_{p}),
1730: $$
1731: where here we have $p\le n$, i.e we consider also
1732: the possibility that $n-p$ spatial locations
1733: in (\ref{tensor-trans}) coincide.
1734: For pure rotations we may treat the amplitudes $x_{1},\ldots ,x_{p}$ as
1735: parameters: the transformations properties of $T^{\alpha _{1}\ldots \alpha
1736: _{n}}$ under rotation are determined only by the dependence of $T^{\alpha
1737: _{1}\ldots \alpha _{n}}$ on the unit vectors $\hat{{\bf x}}_{1},\ldots
1738: ,\hat{{\bf x}} _{p}$. Accordingly it seems worthwhile to discuss tensor
1739: fields
1740: which are functions of the unit vectors {\em only}. Notice that in the
1741: scalar case we follow the same procedure: by restricting our attention to
1742: scalar functions that depend only on the unit vector $\hat{{\bf x}}$, we
1743: construct the spherical harmonics. These functions are {\em defined} such
1744: that each one of them has unique transformation properties under rotations.
1745: We then represent the most general scalar function as a linear combination
1746: of the spherical harmonics with $x$-dependent coefficients, see
1747: Eq.~(\ref{scalar-case}).
1748:
1749: The classification of the tensor fields $T^{\alpha _{1}\ldots \alpha _{n}}(
1750: \hat{{\bf x}}_{1},\ldots ,\hat{{\bf x}}_{p})$ according to their functional
1751: change under rotations follows immediately from group representation
1752: theory \cite{Corn,Stern}. But in order to demonstrate that, we must
1753: first make some formal definitions. We define ${\mathcal S}_{p}^{n}$
1754: to be the space of all smooth tensor fields of rank $n$ which depend
1755: on $p$ unit vectors. This is obviously a linear space of infinite
1756: dimension. With each rotation $\Lambda \in SO(3)$, we may now
1757: associate a linear transformation ${\mathcal O}_{\Lambda }$ on that
1758: space via the relation (\ref{tensor-trans}):
1759: $$
1760: \left[ {\mathcal O}_{\Lambda }T\right] ^{\alpha _{1}\ldots
1761: \alpha_{n}}(\hat{{\bf x}}_{1},\ldots ,\hat{{\bf x}}_{p})
1762: \equiv \Lambda ^{\alpha _{1}\beta _{1}}\cdots \Lambda ^{\alpha
1763: _{n}\beta _{n}}T^{\beta _{1}\ldots \beta _{n}}(\LL\hat{{\bf x}}
1764: _{1},\ldots ,\LL\hat{{\bf x}}_{p}).
1765: $$
1766: Using this definition, it is easy to see that the set of linear operators
1767: $ {\mathcal O}_{\Lambda }$ furnishes a representation of the rotation group
1768: $SO(3)$
1769: since they satisfy the relations:
1770: $$
1771: {\mathcal O}_{\Lambda _{1}}{\mathcal O}_{\Lambda _{2}} ={\mathcal
1772: O}_{\Lambda
1773: _{1}\Lambda _{2}}, \qquad
1774: {\mathcal O}_{\Lambda }^{-1} ={\mathcal O}_{\Lambda ^{-1}}.
1775: $$
1776: General group theoretical considerations imply that it is possible to
1777: decompose ${\mathcal S}_{p}^{n}$ into subspaces which are invariant to the
1778: action of all the group operators ${\mathcal O}_{\Lambda }$. Moreover, we
1779: can
1780: choose these subspaces to be {\em irreducible} in the sense that they will
1781: not contain any invariant subspace themselves (excluding themselves and the
1782: trivial subspace of the zero tensor field). For the $SO(3)$ group each of
1783: these subspaces is conventionally characterized by an integer $
1784: j=0,1,2,\ldots $ and is of dimension $2j+1$ \cite{Corn,Stern}. It should be
1785: noted that unlike
1786: the scalar case, in the general space ${\mathcal S}_{p}^{n}$, there might be
1787: more than one subspace for each given value of $j$. We therefore use the
1788: index $q$ to distinguish subspaces with the same $j$. For each irreducible
1789: subspace $(q,j)$ we can now choose a basis with $2j+1$ components labeled by
1790: the index $m$:
1791: $$
1792: B_{q,jm}^{\alpha _{1},\ldots ,\alpha _{n}}(\hat{{\bf x}}_{1},\ldots ,\hat{
1793: {\bf x}}_{p})\;;\;m=-j,\ldots ,+j.
1794: $$
1795: In each subspace $(q,j)$, the group operators ${\mathcal O}_{\Lambda }$
1796: furnish
1797: a $2j+1$ dimensional irreducible representation of $SO(3)$. Using the basis
1798: $
1799: B_{q,jm}^{\alpha _{1},\ldots ,\alpha _{n}}(\hat{{\bf x}}_{1},\ldots ,\hat{
1800: {\bf x}}_{p})$, we can represent each operator ${\mathcal O}_{\Lambda }$ as
1801: a $
1802: (2j+1)\times (2j+1)$ matrix $D_{m^{\prime }m}^{(j)}(\Lambda )$ via the
1803: relation:
1804: \begin{eqnarray}
1805: \left[ {\mathcal O}_{\Lambda }B\right] _{q,jm}^{\alpha _{1},\ldots ,\alpha
1806: _{n}}(\hat{{\bf x}}_{1},\ldots ,\hat{{\bf x}}_{p})
1807: &=&\Lambda ^{\alpha _{1}\beta _{1}}\cdots \Lambda ^{\alpha
1808: _{n}\beta _{n}}B _{q,jm}^{\beta _{1}\ldots \beta _{n}}(\LL\hat{{\bf x}}
1809: _{1},\ldots ,\LL\hat{{\bf x}}_{p}) \nonumber\\&\equiv &\sum_{m^{\prime
1810: }=-j}^{+j}D_{m^{\prime }m}^{(j)}(\Lambda
1811: )B_{q,jm^{\prime }}^{\alpha _{1},\ldots ,\alpha _{n}}(\hat{{\bf x}}
1812: _{1},\ldots ,\hat{{\bf x}}_{p}).
1813: \nonumber \end{eqnarray}
1814: It is conventional to choose the basis ${\bf B}_{q,jm}$ such that the
1815: matrices $D_{m^{\prime }m}^{(j)}(\phi )$, that correspond to rotations of $
1816: \phi $ radians around the 3 axis, will be diagonal, and given by:
1817: $
1818: D_{m^{\prime }m}^{(j)}(\phi )=\delta _{mm^{\prime }}e^{im\phi }$.
1819: The ${\mathcal S}_{p}^{n}$ space possesses a natural inner-product:
1820: $$
1821: \left\langle {\bf T},{\bf U}\right\rangle \!\! \equiv \!\!\int \!\!d\hat{
1822: {\bf x}}_{1}\dots d\hat{{\bf x}}_{p}
1823: \cdot T^{\alpha _{1}\dots \alpha _{n}}(\hat{{\bf x}}_{1}\dots \hat{{\bf
1824: x}}
1825: _{p})g_{\alpha _{1}\beta _{1}}\ldots g_{\alpha _{n}\beta _{n}}U^{^{\beta
1826: _{1}\ldots \beta _{n}}}(\hat{{\bf x}}_{1}\dots \hat{{\bf x}}_{p})^{\ast }\
1827: $$
1828: where $g_{\alpha \beta }$ is the 3-dimensional Euclidean metric
1829: tensor. By definition, the rotation matrices $\Lambda ^{\alpha
1830: \beta}$ preserve this metric, and therefore it is easy to see that for
1831: each $\Lambda
1832: \in SO(3)$ we get:
1833: $$
1834: \left\langle {\mathcal O}_{\Lambda }{\bf T},{\mathcal O}_{\Lambda }{\bf U}
1835: \right\rangle =\left\langle {\bf T},{\bf U}\right\rangle
1836: $$
1837: so that, ${\mathcal O}_{\Lambda }$ are unitary operators. If we now choose
1838: the
1839: basis ${\bf B}_{q,jm}$ to be orthonormal with respect to the inner-product
1840: defined above, then the matrices $D_{m^{\prime }m}^{(j)}(\Lambda )$ will be
1841: unitary.
1842:
1843: Finally, consider {\em isotropic tensor fields}. An isotropic tensor field
1844: is a tensor field which preserves its functional form under any arbitrary
1845: rotation of the coordinate system. In other words, it is a tensor field
1846: which is invariant to the action of all operators ${\mathcal O}_{\Lambda }$.
1847: The
1848: one dimensional subspace spanned by this tensor-field is therefore invariant
1849: under all operators ${\mathcal O} _{\Lambda }$ and therefore it must be a
1850: $j=0$
1851: subspace.
1852:
1853: Once the basis ${\bf B}_{q,jm}$ has been selected, we may expand any
1854: arbitrary tensor field $F^{\alpha _{1}\ldots \alpha _{n}}({\bf x}_{1},\ldots
1855: ,{\bf x}_{p})$ in this basis. As mentioned above, for each fixed set of
1856: amplitudes $x_{1},\ldots ,x_{p}$, we can regard the tensor field $F^{\alpha
1857: _{1}\ldots \alpha _{n}}({\bf x}_{1},\ldots ,{\bf x}_{p})$ as a tensor field
1858: which depends only on the unit vectors $\hat{{\bf x}}_{1},\ldots ,\hat{{\bf
1859: x }}_{p}$, and hence belongs to ${\mathcal S}_{p}^{n}$. We can therefore
1860: expand
1861: it in terms of the basis tensor fields ${\bf B}_{q,jm}$ with coefficients
1862: that depend on $x_{1},\ldots ,x_{p}$:
1863: \begin{equation}
1864: F^{\alpha _{1}\ldots \alpha _{n}}({\bf x}_{1},\ldots ,{\bf x}_{p})
1865: =\sum_{q,j,m}
1866: {\mathcal F}_{q,jm}(x_{1},\ldots ,x_{p})B_{q,jm}^{\alpha _{1},\ldots
1867: ,\alpha _{n}}(\hat{{\bf x}}_{1},\ldots ,\hat{{\bf x}}_{p}) \ . \label{Texp}
1868: \end{equation}
1869: The goal of the following sections is to demonstrate the utility of such
1870: expansions for the study of the scaling properties of the correlation
1871: functions. For the important case of tensorial structure functions
1872: (\ref{Sn}) the basis function depend on one spatial vector only $\B r$,
1873: and we can expand
1874: \be
1875: \B S^{(n)}(\B r) = \sum_{q,jm} S^{(n)}_{q,jm}(r) \B B^{(n)}_{q,jm}(\hat{ \B
1876: r}).
1877: \label{eq:so3sn}
1878: \ee
1879: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1880: %%%%%%%%%%%%%%
1881: \subsubsection{Construction of the Basis Tensors}
1882: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1883: \label{sec:construction}
1884: \paragraph{The Clebsch-Gordan machinery.}
1885: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1886: A straightforward (although somewhat impractical) way to construct the basis
1887: tensors ${\bf B}_{q,jm}$ is to use the well-known Clebsch-Gordan machinery.
1888: In this approach we consider the ${\mathcal S}_{p}^{n}$ space as a {\em
1889: direct
1890: product space} of $n$ 3-dimensional Euclidean vector spaces with $p$
1891: infinite dimensional spaces of single-variable continuous functions on the
1892: unit sphere. In other words, we notice that ${\mathcal S}_{p}^{n}$ is given
1893: by:
1894: $$
1895: {\mathcal S}_{p}^{n}=\left[ {\mathcal S}_{0}^{1}\right] ^{n}\otimes
1896: \left[ {\mathcal S}
1897: _{1}^{0}\right] ^{p},$$
1898: and therefore every tensor $T^{\alpha_{1}\dots \alpha_{n}}(\hat{{\bf x}}
1899: _{1}\dots \hat{{\bf x}}_{p})$ can be represented as a linear combination of
1900: tensors of the form:$$
1901: v_{1}^{\alpha_{1}} \ldots v_{n}^{\alpha_{n}} \varphi_{1} \left(
1902: \hat{{\bf x}}_{1}\right) \cdot \ldots \cdot \varphi_{p}\left( \hat{{\bf x}}
1903: _{p}\right) .
1904: $$
1905: where
1906: $v_{i}^{\alpha _{i}}$ are constant Euclidean vectors and $\varphi _{i}(\hat{
1907: {\bf x}}_{i})$ are continuous functions over the unit sphere. The
1908: 3-dimensional Euclidean vector space, ${\mathcal S}_{0}^{1}$, contains
1909: exactly
1910: one irreducible representation of $SO(3)$ - the $j=1$ representation - while
1911: ${\mathcal S}_{1}^{0}$, the space of continuous functions
1912: $\otimes$ over the unit sphere,
1913: contains every irreducible representation exactly once. The statement that $
1914: {\mathcal S}_{p}^{n}$ is a direct product space may now be written in a
1915: group
1916: representation notation as:
1917: $$
1918: {\mathcal S}_{p}^{n}= \stackrel{n}
1919: {\overbrace{1 \otimes 1 \otimes
1920: \ldots \otimes 1}} \otimes \stackrel{p}{\overbrace{\left(
1921: 0\oplus 1\oplus 2\dots \right) \otimes \ldots \left( 0\oplus 1\oplus 2\dots
1922: \right) }}
1923: $$
1924: We can now choose an appropriate basis for each space in the product:
1925: \begin{itemize}
1926: \item For the 3-dimensional Euclidean space we may choose:
1927: $$
1928: {\bf e}_{1}={\frac{1}{\sqrt{2}}}\left(
1929: 1,
1930: i,
1931: 0
1932: \right) ,\quad {\bf e}_{0}={\frac{1}{\sqrt{2}}}\left(
1933: 0,
1934: 0,
1935: 1
1936: \right) ,\quad {\bf e}_{-1}={\frac{1}{\sqrt{2}}}\left(
1937: 1,
1938: -i,
1939: 0
1940: \right)
1941: $$
1942: \item For the space of continuous functions over the unit sphere we may
1943: choose the well-known spherical harmonic functions.
1944: \end{itemize}
1945:
1946: Once these bases have been chosen, we can construct a direct-product basis
1947: for ${\mathcal S}_{p}^{n}$:
1948: $$
1949: E_{i_{1}\ldots i_{n}\left( l_{1}\mu _{1}\right) \ldots \left( l_{p}\mu
1950: _{p}\right) }^{\alpha _{1}\ldots \alpha _{n}}\left( {\bf
1951: \hat{x}}_{1},\ldots ,{\bf \hat{x}}_{p}\right) \\
1952: \equiv e_{i_{1}}^{\alpha _{1}}\cdot \dots \cdot e_{i_{n}}^{\alpha_{n}}\cdot
1953: Y_{l_{1},\mu _{1}}(\hat{{\bf x}}_{1})\cdot \dots \cdot
1954: Y_{l_{p},\mu _{p}}(\hat{{\bf x}}_{p}).
1955: $$ The unitary matrix that connects the ${\bf E}_{i_{1}\ldots i_{n}\left(
1956: l_{1}\mu _{1}\right) \ldots \left( l_{p}\mu _{p}\right) }$ basis to
1957: the ${\bf B}_{q,jm}$ basis can be calculated using the appropriate
1958: Clebsch-Gordan coefficients. The calculation is straightforward but
1959: very long and tedious. However, the above analysis enables us to
1960: count and classify the different irreducible representations of a
1961: given $j$. By using the standard rules of angular-momentum addition:
1962: $$
1963: s\otimes l=\left| s-l\right| \oplus \ldots \oplus \left( s+l\right)
1964: $$
1965: we can count the number of irreducible representations of a given
1966: $j$. For example, consider the space ${\mathcal S}_{1}^{2}$ of second-rank
1967: tensors with one variable over the unit sphere. Using the
1968: angular-momentum addition rules we see:
1969: \begin{eqnarray}
1970: {\mathcal S}_{1}^{2} &=&1\otimes 1\otimes \left( 0\oplus 1\oplus 2\oplus
1971: 3\oplus
1972: \ldots \right) \label{dps21} \\
1973: &=&\left( 0\oplus 1\oplus 2\right) \otimes \left( 0\oplus 1\oplus
1974: 2\oplus 3\oplus \ldots \right) \nonumber \\ &=&\left( 3\times
1975: 0\right)\oplus \left( 7\times 1\right) \oplus \left( 9\times 2\right)
1976: \oplus
1977: \left( 9\times 3\right) \oplus \ldots \nonumber
1978: \end{eqnarray}
1979: We see that there are exactly three $j=0$ representations, seven $j=1$
1980: representations and 9 representations for each $j>1$. It can be
1981: further argued that the symmetry properties of the basis tensors with
1982: respect to their indices come from the $1\otimes 1=0\oplus 1\oplus 2$
1983: part of the direct product (\ref{dps21}). Therefore, out of the 9
1984: irreducible representation of $j>1$, 5 will be symmetric and
1985: traceless, 3 will be anti-symmetric and 1 will be trace-full and
1986: diagonal. Similarly, the parity of the resulting tensors (with respect
1987: to the single variable) can be calculated.
1988:
1989: Once we know how many irreducible representations of each $j$ are
1990: found in ${\mathcal S}_{p}^{n}$, we can construct them ``by hand'', in
1991: some other, more practical method which will be demonstrated in the
1992: following subsection.
1993: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1994: \paragraph{Alternative derivation of the basis functions.}
1995: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1996: \label{alternative}
1997: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1998: The method found most useful in application is based on the
1999: simple idea that contractions with $r^{\alpha },\delta ^{\alpha \beta
2000: },\epsilon ^{\alpha \beta \gamma }$ and differentiation with respect
2001: to $r^{\alpha }$ are all {\em isotropic} operations \cite{ara99b}. Isotropic
2002: in the
2003: sense that the resulting expression will have the {\em same}
2004: transformation properties under rotation as the expression we started
2005: with. The proof of the last statement follows directly from the
2006: transformation properties of $r^{\alpha },\delta ^{\alpha \beta
2007: },\epsilon ^{\alpha \beta \gamma }$.
2008:
2009: The construction of all ${\bf B}_{q,jm}$ that belongs to ${\mathcal
2010: S}_{1}^{n}$
2011: now becomes a rather trivial task. We begin by defining a scalar tensor
2012: field with a definite $j,m$. An obvious choice will be the well-known
2013: spherical harmonics $Y_{jm}(\hat{{\bf r}})$, but a better one will be:
2014: $$
2015: \Phi _{jm}({\bf r})\equiv r^{j}Y_{jm}(\hat{{\bf r}}).
2016: $$
2017: The reason that we prefer $\Phi _{jm}({\bf r})$ to $Y_{jm}(\hat{{\bf r}})$,
2018: is that $\Phi _{jm}({\bf r})$ is polynomial in ${\bf r}$ (while
2019: $Y_{jm}(\hat{{\bf r}})$ is polynomial in $\hat{{\bf r}}$) and therefore it
2020: is easier to
2021: differentiate it with respect to ${\bf r}$. Once we have defined $\Phi
2022: _{jm}(
2023: {\bf r})$, we can construct the ${\bf B}_{q,jm}$ by ``adding indices'' to $
2024: \Phi _{jm}({\bf r})$ using the isotropic operations mentioned above. For
2025: example, we may now construct:
2026:
2027: \begin{itemize}
2028: \item $r^{-j}\delta ^{\alpha \beta }\Phi _{jm}({\bf r})$,
2029:
2030: \item $r^{-j+2}\delta ^{\alpha \beta }\partial ^{\tau }\partial ^{\gamma
2031: }\Phi _{jm}({\bf r})$,
2032:
2033: \item $r^{-j-1}r^{\alpha }\Phi _{jm}({\bf r})$, \ etc...
2034: \end{itemize}
2035:
2036: Notice that we should always multiply the resulting expression with an
2037: appropriate power of $r$, in order to make it $r$-independent, and thus a
2038: function of $\hat{{\bf r}}$ only.
2039:
2040: The crucial role of the Clebsch-Gordan analysis is to tell us how many
2041: representations from each type we should come up with. First, it tells
2042: us the highest power of $\hat{{\bf r}}$ in each representation, and
2043: then it can also give us the symmetry properties of ${\bf B}_{q,jm}$
2044: with respect to their indices. For example, consider the irreducible
2045: representations of ${\mathcal S} _{1}^{2}$ - second rank tensors which
2046: depend on one unit vector $\hat{{\bf r} }$. The Clebsch-Gordan
2047: analysis shows us that this space contains the following irreducible
2048: representations spelled out in (\ref{dps21}).
2049: That is, for each $j>1$ we are going to have 9 irreducible representations.
2050: The indices symmetry of the tensor comes from the ${\mathcal
2051: S}_{0}^{1}\otimes
2052: {\mathcal S}_{0}^{1}=1\otimes 1=0\oplus 1\oplus 2$ part of the direct
2053: product.
2054: This is a direct product of two Euclidean spaces, so its a second rank
2055: constant
2056: tensor. We can mark the representation number in this space with the letter
2057: $s$, and the representation number of the ${\mathcal S}_{1}^{0}=0\oplus
2058: 1\oplus
2059: 2\oplus 3\oplus \ldots $ space with the letter $l$. This way each
2060: representation in ${\mathcal S}_{1}^{2}$ of a given $j$ will have two
2061: additional
2062: numbers $(s,l)$, which actually serve as the index $q$ that distinguishes
2063: irreducible representations of the same $j$. The $s$ index will determine
2064: the indices symmetry of the tensor, while the $l$ index will determine the
2065: highest power of $\hat{{\bf r}}$ in the tensor. If we now recall that in the
2066: space of constant second-rank tensors, ${\mathcal S}_{0}^{1}\otimes
2067: {\mathcal S}
2068: _{0}^{1}=0\oplus 1\oplus 2$, the $s=0,2$ representations are symmetric while
2069: the $s=1$ representation is anti-symmetric, we can easily construct the $
2070: { B}^{\alpha \beta}_{q,jm}$:
2071: \begin{equation}
2072: \begin{array}{ll}
2073: \left( s,l\right) =\left( 0,j\right) & B^{\alpha \beta}_{1,jm}(\hat{{\bf
2074: r}})\equiv
2075: r^{-j}\delta ^{\alpha \beta }\Phi _{jm}({\bf r}), \\
2076: \left( s,l\right) =\left( 1,j-1\right) & B^{\alpha \beta}_{2,jm}(\hat{{\bf
2077: r}})\equiv
2078: r^{-j+1}\epsilon ^{\alpha \beta \mu }\partial _{\mu }\Phi _{jm}({\bf r}), \\
2079: \left( s,l\right) =\left( 1,j\right) & B^{\alpha \beta}_{3,jm}(\hat{{\bf
2080: r}})\equiv r^{-j}
2081: \left[ r^{\alpha }\partial ^{\beta }-r^{\beta }\partial ^{\alpha }\right]
2082: \Phi _{jm}({\bf r}), \\
2083: \left( s,l\right) =\left( 1,j+1\right) & B^{\alpha \beta}_{4,jm}(\hat{{\bf
2084: r}})\equiv
2085: r^{-j-1}\epsilon ^{\alpha \beta \mu }r_{\mu }\Phi _{jm}({\bf r}), \\
2086: \left( s,l\right) =\left( 2,j-2\right) & B^{\alpha \beta}_{5,jm}(\hat{{\bf
2087: r}})\equiv
2088: r^{-j+2}\partial ^{\alpha }\partial ^{\beta }\Phi _{jm}({\bf r}), \\
2089: \left( s,l\right) =\left( 2,j-1\right) & B^{\alpha \beta}_{6,jm}(\hat{{\bf
2090: r}})\equiv
2091: r^{-j+1}\left[ \epsilon ^{\alpha \mu \nu }r_{\mu }\partial _{\nu }\partial
2092: ^{\beta }+\epsilon ^{\beta \mu \nu }r_{\mu }\partial _{\nu }\partial
2093: ^{\alpha }\right] \Phi _{jm}({\bf r}), \\
2094: \left( s,l\right) =\left( 2,j\right) & B^{\alpha \beta}_{7,jm}(\hat{{\bf
2095: r}})\equiv r^{-j}
2096: \left[ r^{\alpha }\partial ^{\beta }+r^{\beta }\partial ^{\alpha }\right]
2097: \Phi _{jm}({\bf r}), \\
2098: \left( s,l\right) =\left( 2,j+1\right) & B^{\alpha \beta}_{8,jm}(\hat{{\bf
2099: r}})\equiv
2100: r^{-j-1}\left[ r^{\beta }\epsilon ^{\alpha \mu \nu }r_{\mu }\partial _{\nu
2101: }+r^{\alpha }\epsilon ^{\beta \mu \nu }r_{\mu }\partial _{\nu }\right] \Phi
2102: _{jm}({\bf r}), \\
2103: \left( s,l\right) =\left( 2,j+2\right) & B^{\alpha \beta}_{9,jm}(\hat{{\bf
2104: r}})\equiv
2105: r^{-j-2}r^{\alpha }r^{\beta }\Phi _{jm}({\bf r}).
2106: \end{array}
2107: \label{eq:second-rank-tensors}
2108: \end{equation}
2109: It should be stressed that these $B^{\alpha \beta}_{q,jm}$ are not exactly
2110: the same
2111: one we would have gotten from the Clebsch-Gordan machinery. For example,
2112: they are not orthogonal among themselves for the same values of $j,m$
2113: (although, they are orthogonal for different values of $j$ or $m$).
2114: Nevertheless, they are linearly independent and thus span a given $(j,m)$
2115: sector in the ${\mathcal S}_{1}^{2}$ space.
2116: The set of eigenfunction, $B^{\alpha \beta}_{q,jm}$, can be further
2117: classified in terms of its properties under permutation of tensorial
2118: indices,
2119: $\alpha \beta$ and in terms of their parity properties,
2120: i.e. how do they transform under the $\B r \rightarrow -\B r$ operation.
2121: Taking in to account both properties we may distinguish:
2122: \begin{description}
2123: \item{Subset I} Symmetric in $\alpha,\beta$ and with parity $(-1)^j$:
2124: $$B^{\alpha\beta}_{9,jm}(\hat{\B r}),
2125: B^{\alpha\beta}_{7,jm}(\hat{\B r}),
2126: B^{\alpha\beta}_{1,jm}(\hat{\B r}),
2127: B^{\alpha\beta}_{5,jm}(\hat{ \B r})$$
2128: \item{Subset II} Symmetric to $\alpha,\beta$ exchange and with parity
2129: $(-1)^{j+1}$:
2130: $$ B_{8,jm}^{\alpha \beta }(\hat{ \B r}),B_{6,jm}^{\alpha\beta}(\hat{ \B
2131: r})$$
2132: \item{Subset III} Antisymmetric to $\alpha,\beta$ exchange and with parity
2133: $(-1)^{j+1}$:
2134: $$ B_{4,jm}^{\alpha \beta }(\hat{ \B r}),B_{2,jm}^{\alpha\beta}(\hat{ \B
2135: r})$$
2136: \item{Subset IV} Antisymmetric to $\alpha,\beta$ exchange and with parity
2137: $(-1)^{j}$:
2138: $$B_{3,jm}^{\alpha \beta }(\hat{ \B r})$$
2139: \end{description}
2140:
2141: The reader may find more details on the algebra of SO(3) decomposition of
2142: second order tensor
2143: in the Appendix ~\ref{sec:general-form}.
2144: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2145: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2146: \subsection{The Isotropy of the Hierarchy of Equations and
2147: its Consequences}
2148: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2149: \label{sec:hierachy}
2150: In this section we follow Ref. \cite{ara99b} in deriving equations of motion
2151: for the
2152: statistical averages of the velocity and pressure fields
2153: differences. We start from the Navier-Stokes equations, and show that
2154: its isotropy implies the isotropy of the equations for the statistical
2155: objects. Finally, we demonstrate the foliation of these equations to
2156: different sectors of $j,m$.
2157: Consider the Navier-Stokes equations (\ref{NS}) in a bounded domain
2158: $\Omega$.
2159: In principle these equations can be the basis for deriving infinite linear
2160: hierarchy
2161: of equations for the Eulerian correlation functions and to study its
2162: properties
2163: under rotation.
2164: Unfortunately the relevant dynamical time scales are revealed only when
2165: the effect of sweeping is removed. Therefore we choose to work here
2166: with the transformation
2167: proposed in \cite{bel87} in which the flow is observed from the point of
2168: view of one specific fluid particle which is located at ${\bf x}_{0}$ at
2169: time $t_{0}$. Let ${\bf \rho }({\bf x}_{0},t_{0}|t)$ be the particle's
2170: translation at time $t$:
2171: $$
2172: {\bf \rho }({\bf x}_{0},t_{0}|t)=\int\limits_{t_{0}}^{t}ds{\bf u}[{\bf x}
2173: _{0}+{\bf \rho }({\bf x}_{0},t_{0}|s),s]\ . $$
2174: We then redefine the velocity and pressure fields to be those seen from an
2175: inertial frame whose origin sits at the current particle's position:
2176: $$
2177: {\bf v}({\bf x}_{0},t_{0}|{\bf x},t) \equiv {\bf u}[{\bf x}+{\bf \rho }(
2178: {\bf x}_{0},t_{0}|t),t]\ , $$
2179: $$\pi ({\bf x}_{0},t_{0}|{\bf x},t) \equiv p[{\bf x}+{\bf \rho }({\bf x}
2180: _{0},t_{0}|t),t]\ .$$
2181: Next, we define the differences of these fields:
2182: $$
2183: {\mathcal W}^{\alpha }({\bf x}_{0},t_{0}|{\bf x},{\bf x}^{\prime },t)
2184: \equiv v^{\alpha }({\bf x}_{0},t_{0}|{\bf x},t)-v^{\alpha }({\bf
2185: x}_{0},t_{0}|{\bf
2186: x}^{\prime },t)\ , $$
2187: $$
2188: \Pi ({\bf r}_{0},t_{0}|{\bf x},{\bf x}^{\prime },t) \equiv \pi (
2189: {\bf x}_{0},t_{0}|{\bf x},t)-\pi ({\bf x}_{0},t_{0}|{\bf x}^{\prime },t)\ .
2190: $$
2191: A straightforward calculation shows that the dynamical equations for $
2192: {\bf {\mathcal W}}$ are:
2193: \begin{eqnarray}
2194: &&\partial _{t}{\mathcal W}^{\alpha }({\bf x},{\bf x}^{\prime },t) =-\left(
2195: \partial _{\alpha }+\partial ^{\prime}_{ \alpha }\right) \Pi ({\bf x}
2196: _{0},t_{0}|{\bf x},{\bf x}^{\prime },t) \label{eq:WDyn}
2197: +\nu \left( \partial ^{2}+\partial ^{\prime 2}\right) {\mathcal W}^{\alpha
2198: }(
2199: {\bf x}_{0},t_{0}|{\bf x,x}^{\prime },t) \nonumber \\
2200: &&-\partial _{\mu }{\mathcal W}^{\mu }({\bf x}_{0},t_{0}|{\bf
2201: x,x}_{0},t){\mathcal
2202: W}
2203: ^{\alpha }({\bf x}_{0},t_{0}|{\bf x,x}^{\prime },t)
2204: -\partial _{\mu }^{\prime }{\mathcal W}^{\mu }({\bf x}_{0},t_{0}|{\bf x}
2205: ^{\prime }{\bf ,x}_{0},t){\mathcal W}^{\alpha }({\bf x}_{0},t_{0}|{\bf x,x}
2206: ^{\prime },t)\ , \nonumber \\
2207: &&\partial _{\alpha }{\mathcal W}^{\alpha }({\bf x}_{0},t_{0}|{\bf x},{\bf
2208: x}
2209: ^{\prime },t) =\partial _{\alpha }^{\prime }{\mathcal W}^{\alpha }({\bf x}
2210: _{0},t_{0}|{\bf x},{\bf x}^{\prime },t)=0\ .
2211: \end{eqnarray}
2212: By inspection, $t_{0}$ merely serves as a parameter, and therefore we will
2213: not denote it explicitly in the following discussion. Also, in order to make
2214: the equations easier to understand, let us introduce some shorthand notation
2215: for the variables $({\bf x}_{k},{\bf x}_{k}^{\prime },t_{k})$:
2216: $$
2217: {\bf X}_{k} \equiv ({\bf x}_{k},{\bf x}_{k}^{\prime },t_{k})\ , \quad
2218: X_{k} \equiv (x_{k},x_{k}^{\prime },t_{k})\ , \quad
2219: {\hat{{\bf X}}}_{k} \equiv (\hat{{\bf x}}_{k},\hat{{\bf x}}_{k}^{\prime
2220: })\ .
2221: $$
2222: Using (\ref{eq:WDyn}), we can now derive the dynamical equations for the
2223: statistical moments of ${\bf {\mathcal W}}$,$\Pi $: Let $\left\langle \cdot
2224: \right\rangle $ denote a suitable ensemble averaging. We define two types of
2225: statistical moments:
2226: \begin{eqnarray*}
2227: &&{\mathcal F}^{\alpha _{1}\ldots \alpha _{n}}({\bf x}_{0}|{\bf
2228: X}_{1},\ldots ,{\bf
2229: X}_{n})
2230: \equiv \left\langle {\mathcal W}^{\alpha _{1}}({\bf x}_{0}|{\bf
2231: X}_{1})\ldots
2232: {\mathcal W}^{\alpha _{n}}({\bf x}_{0}|{\bf X}_{n})\right\rangle \ , \\
2233: &&{\mathcal H}^{\alpha _{2}\ldots \alpha _{n}}({\bf x}_{0}|{\bf
2234: X}_{1},\ldots ,{\bf
2235: X}_{n})
2236: \equiv \left\langle \Pi (x_{0}|{\bf X}_{1}){\mathcal W}^{\alpha _{2}}({\bf
2237: x}
2238: _{0}|{\bf X}_{2})\ldots {\mathcal W}^{\alpha _{n}}({\bf x}_{0}|{\bf X}
2239: _{n})\right\rangle \ .
2240: \end{eqnarray*}
2241: Equation (\ref{eq:WDyn}) implies:
2242: \begin{eqnarray}
2243: &&\partial _{t_{1}}{\mathcal F}^{\alpha _{1}\ldots \alpha _{n}}({\bf
2244: x}_{0}|{\bf X}
2245: _{1},\ldots ,{\bf X}_{n})
2246: =-\left( \partial _{(x_{1})}^{\alpha _{1}}+\partial _{(x_{1}^{\prime
2247: })}^{\alpha _{1}}\right) {\mathcal H}^{\alpha _{2}\ldots \alpha _{n}}({\bf
2248: x}_{0}|
2249: {\bf X}_{1},\ldots ,{\bf X}_{n}) \nonumber \\
2250: &&-\partial _{\mu }^{(x_{1})}{\mathcal F}^{\mu \alpha _{1}\ldots \alpha
2251: _{n}}\left( {\bf x}_{0}|{\tilde{{\bf X}}},{\bf X}_{1},\ldots ,{\bf
2252: X}_{n}\right)
2253: -\partial _{\mu }^{(x_{1}^{\prime })}{\mathcal F}^{\mu \alpha _{1}\ldots
2254: \alpha
2255: _{n}}\left( {\bf x}_{0}|{\tilde{{\bf X}}}^{\prime }{\bf ,X}_{1},\ldots ,{\bf
2256: X}_{n}\right) \nonumber \\
2257: &&+\nu \left( \partial _{(x_{1})}^{2}+\partial _{(x_{1}^{\prime
2258: })}^{2}\right) {\mathcal F}^{\alpha _{1}\ldots \alpha _{n}}({\bf x}_{0}|{\bf
2259: X}
2260: _{1},\ldots ,{\bf X}_{n})\ , \label{eq:FDyn}
2261: \end{eqnarray}
2262: with ${\tilde{{\bf X}}} \equiv {(}{\bf x}_{0},{\bf x}^{\prime
2263: },t)\;;\;{\tilde{
2264: {\bf X}}}^{\prime }{\equiv (}{\bf x},{\bf x}_{0},t)\ ,$ with the
2265: further constraint:
2266: $$
2267: \partial _{\alpha _{1}}^{(x_{1})}{\mathcal F}^{\alpha _{1}\ldots \alpha
2268: _{n}}({\bf x
2269: }_{0}|{\bf X}_{1},\ldots ,{\bf X}_{n}) =0\ , \qquad
2270: \partial _{\alpha _{1}}^{(x_{1}^{\prime })}{\mathcal F}^{\alpha _{1}\ldots
2271: \alpha
2272: _{n}}({\bf x}_{0}|{\bf X}_{1},\ldots ,{\bf X}_{n}) =0\ . $$
2273: Equations (\ref{eq:FDyn}), are linear and homogeneous.
2274: Therefore their solutions form a linear space. The most general solution to
2275: these equations is given by a linear combination of a suitable basis of the
2276: solutions space. To construct a specific solution, we must use the boundary
2277: conditions in order to set the linear weights of the basis solutions. We
2278: shall now show that the isotropy of these equations implies that our
2279: basis of solutions can be constructed such that every solution will have a
2280: definite behavior under rotations (that is, definite $j$ and $m$).
2281: But before we do that, note that in many aspects the situation
2282: described here is similar to the well-known problem of Laplace equation in a
2283: closed domain $\Omega $:
2284: $$
2285: \nabla ^{2}\Psi =0\, \qquad
2286: \Psi |_{\partial \Omega } =\sigma \ .
2287: $$
2288: The Laplace equation is linear, homogeneous and isotropic. Therefore its
2289: solutions form a linear space. One possible basis for this space is:
2290: $$
2291: \Psi_{jm}({\bf r})\equiv r^{j}Y_{jm}(\hat{{\bf r}})\ ,
2292: $$
2293: in which the solutions have a definite behavior under rotations (belong
2294: to an irreducible representation of $SO(3)$ ). The general solution of the
2295: problem is given as a linear combination of the $\Psi_{jm}({\bf r})$, cf.
2296: Eq. (\ref{scalar-case}).
2297: For a specific problem, we use the value of $\Psi ({\bf r})$ on the boundary
2298: (i.e., we use $\sigma ({\bf r})$) in order to set the values of $a_{l,m}$.
2299:
2300: To see that the same thing happens with the hierarchy equations
2301: (\ref{eq:FDyn}), we consider an arbitrary solution
2302: $\{{\B {\mathcal F}}^{(n)},{\B {\mathcal H}}^{(n)}|\;n=2,3,\ldots \}$ of
2303: these
2304: equations. We may write the tensor fields
2305: ${\B{\mathcal F}}^{(n)},{\B {\mathcal H}}^{(n)}$ in terms of a basis ${\bf
2306: B}_{q,jm}$:
2307: \begin{equation}
2308: \label{def:jmExpansion}
2309: {\mathcal F}^{\alpha _{1}\ldots \alpha _{n}}({\bf x}_{0}|{\bf X}_{1},\ldots
2310: ,{\bf
2311: X}_{n})
2312: \equiv \sum_{q,j,m} F_{q,jm}^{(n)}(x_{0},X_{1},\ldots ,X_{n})
2313: \B B_{q,jm}^{(n)}(\hat{{\bf x}}_{0},{
2314: \hat{{\bf X}}}_{1},\ldots {\hat{{\bf X}}}_{n})\ ,
2315: \end{equation}
2316: \begin{equation}
2317: {\mathcal H}^{\alpha _{2}\ldots \alpha _{n}}({\bf x}_{0}|{\bf X}_{1},\ldots
2318: ,{\bf
2319: X}_{n})
2320: \equiv \sum_{q,j,m} H_{q,jm}^{(n)}(x_{0},X_{1},\ldots ,X_{n})
2321: \B B_{q,jm}^{(n-1)}(\hat{{\bf r}}_{0},{
2322: \hat{{\bf X}}}_{1},\ldots {\hat{{\bf X}}}_{n})\ ;
2323: \end{equation}
2324: where here and below we use the shorthand notation, $\B B_{q,jm}^{(n)}$
2325: to denote the SO(3) basis of $n$th order tensors,
2326: $ B_{q,jm}^{\alpha_1,\dots,\alpha_n}$.
2327: Now all we have to show is that the pieces of
2328: ${\B {\mathcal F}}^{(n)},{\B {\mathcal
2329: H}}^{(n)}$ with
2330: definite $j,m$ solve the hierarchy equations {\em by themselves} -
2331: independently of pieces with different $j,m$. The proof of the last
2332: statement is straightforward though somewhat tedious. We therefore only
2333: sketch it in general lines. The isotropy of the hierarchy equations implies
2334: that pieces of ${\B {\mathcal F}}^{(n)},{\B {\mathcal H}}^{(n)}$ with
2335: definite $j,m$,
2336: maintain their
2337: transformations properties under rotation {\em after} the linear and
2338: isotropic operations of the equation have been performed. For example, if $
2339: {\mathcal F}^{\alpha _{1}\ldots \alpha _{n}}({\bf x}_{0}|{\bf X}_{1},\ldots
2340: ,{\bf X}
2341: _{n})$ belongs to the irreducible representation $(j,m)$, then so will the
2342: tensor fields:
2343: $
2344: \partial _{\alpha _{i}}^{(x_{k})}{\mathcal F}^{\alpha _{1}\ldots \alpha
2345: _{n}},
2346: \partial _{\alpha _{i}}^{(x_{k})}\partial_{\alpha _{i}}^{(x_{k})}{\mathcal
2347: F}^{\alpha _{1}\ldots \alpha
2348: _{n}},
2349: $
2350: although, they may belong to different ${\mathcal S}_{p}^{n}$ spaces
2351: (i.e., have one less or one more indices). Therefore, if we choose the
2352: bases $\left\{ {\bf B}_{q,jm}^{(n)}\right\} $ to be orthonormal, plug
2353: the expansion (\ref {def:jmExpansion}) into the hierarchy equations
2354: equations (\ref{eq:FDyn}), and take the inner product with ${\bf
2355: B}_{q,jm}^{(n)}$, we
2356: will obtain new equations for the scalar functions
2357: $F_{q,jm}^{(n)},H_{q,jm}^{(n)}$:
2358: \begin{eqnarray}
2359: &&\partial _{t_{1}}F_{q,jm}^{(n)}(r_{0},X_{1},\ldots ,X_{n})
2360: =\nonumber\\&&-\sum_{q^{\prime }}\left\langle \left( \partial
2361: _{(x_{1})}^{\alpha
2362: _{1}}+\partial _{(x_{1}^{\prime })}^{\alpha _{1}}\right) H_{q^{\prime
2363: }jm}^{(n)}(r_{0},X_{1},\ldots ,X_{n}){\bf B}_{q^{\prime }jm}^{(n-1)},{\bf B}
2364: _{q,jm}^{(n)}\right\rangle \label{eq:ScalarDyn} \\
2365: &&-\sum_{q^{\prime }}\left\langle \partial _{\mu }^{(x_{1})}F_{q^{\prime
2366: }jm}^{(n+1)}(r_{0},\tilde{X},X_{1},\ldots ,X_{n}){\bf B}_{q^{\prime
2367: }jm}^{(n+1)},{\bf B}_{q,jm}^{(n)}\right\rangle\nonumber\\&&
2368: -\sum_{q^{\prime }}\left\langle \partial _{\mu }^{(x_{1}^{\prime
2369: })}F_{q^{\prime }jm}^{(n+1)}(r_{0},\tilde{X}^{\prime },X_{1},\ldots ,X_{n})
2370: {\bf B}_{q^{\prime }jm}^{(n+1)},{\bf B}_{q,jm}^{(n)}\right\rangle
2371: \nonumber
2372: \\
2373: &&+\nu \sum_{q^{\prime }}\left\langle \left( \partial
2374: _{(x_{1})}^{2}+\partial _{(x_{1}^{\prime })}^{2}\right) F_{q^{\prime
2375: }jm}^{(n)}(r_{0},X_{1},\ldots ,X_{n}){\bf B}_{q^{\prime }jm}^{(n)},{\bf B}
2376: _{q,jm}^{(n)}\right\rangle , \nonumber
2377: \end{eqnarray}
2378: \begin{eqnarray}
2379: \sum_{q^{\prime }}\left\langle \partial _{\alpha _{1}}^{(x_{1})}F_{q^{\prime
2380: }jm}^{(n)}(x_{0},X_{1},\ldots ,X_{n}){\bf B}_{q^{\prime }jm}^{(n)},{\bf B}
2381: _{q,jm}^{(n-1)}\right\rangle &=&0\ , \label{eq:ScalarIncomp} \\
2382: \sum_{q^{\prime }}\left\langle \partial _{\alpha _{1}}^{(x_{1}^{\prime
2383: })}F_{q^{\prime }jm}^{(n)}(r_{0},X_{1},\ldots ,X_{n}){\bf B}_{q^{\prime
2384: }jm}^{(n)},{\bf B}_{q,jm}^{(n-1)}\right\rangle &=&0\ . \nonumber
2385: \end{eqnarray}
2386: Note that in the above equations, $\left\langle \cdot \right\rangle $ denote
2387: the inner-product in the ${\mathcal S}_{p}^{n}$ spaces. Also, the sums over
2388: $
2389: q^{\prime },j^{\prime },m^{\prime }$ from (\ref{def:jmExpansion}) was
2390: reduced to a sum over $q^{\prime }$ only - due to the isotropy. We thus see
2391: explicitly from (\ref{eq:ScalarDyn},\ref{eq:ScalarIncomp}) the decoupling of
2392: the equations for different $j,m$.\\
2393: At this point we remind the reader that in the case of the most used
2394: statistical objects
2395: in the analysis of experimental and numerical data are the
2396: longitudinal $n$th order structure functions:
2397: $$
2398: S^{(n)}(\B r) = \langle (\delta u_{\ell}(\B r))^n \rangle .
2399: $$
2400: For these objects the basis functions are simply the spherical harmonics
2401: and the SO(3) decomposition reads:
2402: \be
2403: S^{(n)}(\B r) = \sum_{j,m} S^{(n)}_{jm}(r) Y_{jm}(\hat{\B r}).
2404: \label{eq:fundamental}
2405: \ee
2406: A question of major interest for all that follows are the numerical values
2407: of the scaling
2408: exponents which are defined by the power laws
2409: $$
2410: S^{(n)}_{jm}(r) \propto r^{\zeta_j^{(n)}} $$
2411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2412: \subsection{Dimensional Analysis of Anisotropic Fluctuations}
2413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2414: \label{sec:dim_ana}
2415: The actual calculation of scaling exponents in the anisotropic
2416: sectors is difficult, and will be considered in the rest of this
2417: review. It is worthwhile to have a phenomenological guess based on
2418: dimensional analysis. Unfortunately, once anisotropies are considered,
2419: dimensional considerations becomes tricky. Historically, the first
2420: successful attempt to introduce dimensional considerations in
2421: anisotropic turbulence was the approach discussed in
2422: Sect.~\ref{sec:lumley}. There, the key role was played by the large-scale
2423: mean-shear. However this work is limited to the analysis of second order
2424: correlations, without discriminating among
2425: different $j$ sectors. In light of the SO(3)
2426: decomposition it should be considered as a prediction for
2427: $\zeta_{2}^{(2)}$.
2428: Another dimensional argument was presented in
2429: \cite{bif02} extending the consideration of Sec.~\ref{sec:lumley}.
2430: This argument takes into
2431: account also the particular angular structure entering in the
2432: interaction between small-scale fluctuations and large-scale shear.
2433: By decomposing the velocity field, $\B{u}$,
2434: in a small-scale component, $\B{v}$, and a large-scale
2435: anisotropic component, $\B{U}$, one finds the following equation for
2436: the time evolution of $\B{v}$:
2437: $$
2438: \partial_t v_{\alpha} + v_{\beta}\partial_{\beta} v_{\alpha} +
2439: U_{\beta}\partial_{\beta} v_{\alpha} + v_{\beta}\partial_{\beta}
2440: U_{\alpha} = -\partial_{\alpha} p + \nu\Delta v_{\alpha}.
2441: $$
2442: The major effect of the large-scale field is given by the
2443: instantaneous shear $\partial_{\beta} U_{\alpha}$ which acts as an
2444: anisotropic forcing term on small scales.
2445: We can write the balance
2446: equation for two
2447: point quantities $\la v_{\delta}(\Bx')v_{\alpha}(\Bx)\ra$ in the
2448: stationary regime:
2449: $$
2450: \la v_{\delta}(\Bx') v_{\beta}(\Bx)
2451: \partial_{\beta} v_{\alpha}(\Bx) \ra \sim \la
2452: \partial_{\mu} U_{\alpha}
2453: v_{\delta}(\Bx') v_{\mu}(\Bx) \ra.
2454: $$
2455: The shear
2456: term is a large-scale ``slow'' quantity and therefore, as far as
2457: scaling properties are concerned, can be safely estimated as: $\la
2458: \partial_{\mu} U_{\alpha}
2459: v_{\delta}(\Bx') v_{\mu}(\Bx) \ra \sim
2460: D_{\alpha \mu} \la v_{\delta}(\Bx')v_{\mu}(\Bx)
2461: \ra$. The tensor $D_{\alpha \beta}$ is associated to the joint
2462: probability to have a given shear and a given small scale velocity
2463: configuration. The $D_{\alpha \beta}$ being a constant tensor can possess at
2464: most angular momentum
2465: up to $j=2$.
2466: Similarly for three point quantities
2467: we may write: $\la v v v \partial v \ra \sim \la \partial U vvv \ra$,
2468: which can
2469: be easily generalized to velocity correlation of any order. One may
2470: therefore argue, by using simple composition
2471: of angular momenta, ($j= 2 \oplus j-2$), the following dimensional
2472: matching for structure functions in different anisotropic sectors:
2473: \be {S}^{(n)}_{jm}(r) \;\sim\; r\, |D| \, \cdot \,
2474: S_{j-2,m}^{(n-1)}(r),
2475: \label{eq:gen}
2476: \ee
2477: where $ S_{j,m}^{(n)}(r)$ is a shorthand notation for the projection
2478: on the $j$-{\it th} sector of the $n$-{\it th} order correlation
2479: function introduced in the previous section, ${F}^{(n)}_{q jm}(r)$.
2480: In (\ref{eq:gen}) with $|D|$ we denote the typical intensity of the
2481: shear term $D_{\alpha \beta}$ in the $j=2$ sector. From
2482: Eq. (\ref{eq:gen}) one can obtain higher $j$ exponents of the higher
2483: order structure functions from the lower order structure functions of
2484: lower anisotropic sectors which appear on the RHS. For example, the
2485: dimensional prediction for the third order scaling exponent in the
2486: $j=2$ sector, $\zeta^{(3)}_2$ can be obtained by the
2487: matching: $S_{2,m}^{(3)}(r) \sim r\, |D| S_{0,m}^{(2)}(r) \sim
2488: r^{\zeta_{2}^{(3)}}$. By using the same argument and the known scaling of
2489: the third order correlation for $j=0,2$, the scaling exponents of the
2490: fourth order correlation for $j=2,4$ can be estimated. The following
2491: expression is readily obtained for any order\,:
2492: \begin{equation}
2493: \zeta_{j}^{(n)}=\frac{(n+j)}{3} \quad \mbox{(dimensional prediction).}
2494: \label{lumleygen}
2495: \end{equation}
2496: This formula coincides with the prediction (\ref{eq:Lumley})
2497: for $n=2$ and $j=2$. We will see below that
2498: both measurements and closure calculations exhibit
2499: exponents which are anomalous, i.e.
2500: different from these dimensional predictions.
2501: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2502: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2503: \section{Exactly Solvable Models}
2504: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2505: \label{chap:analytical}
2506: In this section we review the work done on anomalous scaling in
2507: the anisotropic sectors of exactly solvable models. The first of
2508: these models is the Kraichnan model of passive scalar advection
2509: in which the velocity field is rapidly varying in time
2510: \cite{kra68,kra94,fal01}. This
2511: model offers detailed understanding of the anomalous scaling in all the
2512: anisotropic sectors both from the Lagrangian and the Eulerian
2513: points of view. The scaling exponents can be calculated however
2514: only in perturbation theory. The second model that we consider is
2515: of passive advection of a magnetic field \cite{ver96}. In this case one can
2516: compute non-perturbatively the scaling exponents of the second order
2517: correlation function in all the sectors of the symmetry group.
2518: These two models show that the spectrum of scaling exponents is
2519: discrete and strictly increasing as a function of $j$. If this
2520: is true for systems with pressure, like the Navier-Stokes
2521: equation, it may lead to problems of convergence of the integrals
2522: induced by the existence of the pressure terms. To this aim we
2523: review below a third exactly solvable model in which pressure is
2524: used explicitly to keep an advected vector solenoidal.It was shown
2525: that also here the spectrum is discrete and strictly increasing,
2526: and it was explained how the putative divergences are avoided. The
2527: mechanism discovered here is most likely also operating in the
2528: Navier-Stokes case. The last model reviewed in this section
2529: is the the second order structure function in the Navier-Stokes problem,
2530: linearized for small anisotropies. Also in this case we find
2531: a discrete spectrum of strictly increasing scaling exponents
2532: as a function of $j$.
2533: Most of the results here presented can also be reproduced within
2534: the Renormalization Group approach. We do not enter here in this
2535: subject which would deserve a whole review by itself.
2536: The interested reader can find the most important results
2537: for passive scalar advection in \cite{anto01,anto01a}, for magnetic fields
2538: in \cite{anto00,anto99} and for passive vectors in \cite{anto03}.\\
2539: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2540: \subsection{Anomalous Scaling in the Anisotropic Sectors
2541: of the Kraichnan Model of Passive Scalar Advection}
2542: \label{passcalar}
2543: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2544: Kraichnan's model of passive scalar advection in
2545: which the driving (Gaussian) velocity field has fast temporal
2546: decorrelation turned out to be a very important case model for
2547: understanding the anomalous scaling behavior in turbulent
2548: advection, including the anisotropic sectors of turbulent scalar fields.
2549: We review here the derivation that shows that the solutions of the Kraichnan
2550: equation for the
2551: $n$ order correlation functions foliate into sectors that are
2552: classified by the irreducible representations of the SO($d$)
2553: symmetry group. A discrete spectrum of universal
2554: anomalous exponents is found, with a different exponent characterizing the
2555: scaling behavior in every sector. Generically the correlation
2556: functions and structure functions appear as sums over all these
2557: contributions, with non-universal amplitudes which are determined
2558: by the anisotropic boundary conditions. The isotropic sector is
2559: always characterized by the smallest exponent, and therefore for
2560: sufficiently small scales local isotropy is always restored.
2561: %We start the discussion with a Lagrangian calculation that is not
2562: %aimed at computing the scaling exponents, but to reveal the
2563: %intimate connection between the anomalous exponents and the
2564: %dynamics of configurations of Lagrangian trajectories.
2565: We start by
2566: presenting the Eulerian calculation which results in actual values
2567: of the scaling exponents (in perturbation theory) \cite{ara00}. The Eulerian
2568: calculation of the anomalous exponents is done in two
2569: complementary ways. In the first they are obtained from the
2570: analysis of the correlation functions of {\em gradient fields}.
2571: The theory of these functions involves the control of logarithmic
2572: divergences which translate into anomalous scaling with the ratio
2573: of the inner {\em and} the outer scales appearing in the final
2574: result. In the second way one computes the exponents from the zero
2575: modes of the Kraichnan equation for the correlation functions of
2576: the scalar field itself. In this case the renormalization scale
2577: is the outer scale. The two approaches lead to the same scaling
2578: exponents for the same statistical objects, illuminating the
2579: relative role of the outer and inner scales as renormalization
2580: scales. To clarify this further, Ref.~\cite{ara00} presented
2581: an exact derivation of
2582: fusion rules which govern the small scale asymptotic of the
2583: correlation
2584: functions in all the sectors of the symmetry group and in all
2585: dimensions. The purpose of the Eulerian calculation is twofold. On the one
2586: hand we are interested in the effects of anisotropy on the
2587: universal aspects of scaling behavior in turbulent systems. On the other
2588: hand
2589: we are interested in clarifying the relationship between
2590: ultraviolet and infrared anomalies in turbulent systems.
2591: The two issues discussed in this subsection have an importance
2592: that transcends the particular example that we treat here in
2593: detail. Having below a theory of anomalous scaling in all the various
2594: sectors of
2595: the symmetry group allows us to explain clearly the relationship
2596: between the two renormalization scales and the anomalous
2597: exponents that are implied by their existence.
2598: Since we expect
2599: that Kolmogorov type theories, which assume that no
2600: renormalization scale appears in the theory, are generally
2601: invalidated by the appearance of both the outer and the inner
2602: scales as renormalization scales, the clarification of the
2603: relation between the two is important also for other cases of
2604: turbulent statistics.
2605:
2606: The central quantitative result of the Eulerian calculation is the
2607: expression for the scaling exponent $\xi^{(n)}_{j}$ which is
2608: associated with the scaling behavior of the $n$-order correlation
2609: function (or structure function) of the scalar field in the
2610: $j$'th sector of the symmetry group. In other words, this is
2611: the scaling exponent of the projection of the correlation
2612: function on the $j$'th irreducible representation of the
2613: SO($d$) symmetry group, with $n$ and $j$ taking on even values
2614: only, $n=0,2, \dots$ and $j=0,2,\dots$:
2615: \begin{equation}
2616: \xi^{(n)}_{j}= n-\epsilon\Big[\frac{n(n+d)}{2(d+2)}
2617: -\frac{(d+1)j(j+d-2)}{2(d+2)(d-1)}\Big] +O(\epsilon^2) \ .
2618: \label{eq:perturbative}
2619: \end{equation}
2620: The result is valid for any even $j\le n$, and to $O(\epsilon)$
2621: where $\epsilon$ is the scaling exponent of the eddy diffusivity
2622: in the Kraichnan model (and see below for details). In the
2623: isotropic sector ($j=0$) we recover the result of
2624: \cite{ber96}. It is noteworthy that for higher values of $j$
2625: the discrete spectrum is a strictly increasing function of
2626: $j$. This is important, since it shows that for diminishing
2627: scales the higher order scaling exponents become irrelevant, and
2628: for sufficiently small scales only the isotropic contribution
2629: survives. As the scaling exponent appear in power laws of the
2630: type $(r/\Lambda)^\xi$, with $\Lambda$ being some typical outer scale
2631: and $r \ll \Lambda$, the larger is the exponent, the faster is the
2632: decay of the contribution as the scale $r$ diminishes. This is
2633: precisely how the isotropization of the small scales takes place,
2634: and the higher order exponents describe the rate of
2635: isotropization. Nevertheless for intermediate scales or for finite
2636: values of the Reynolds and Peclet numbers the lower lying scaling
2637: exponents will appear in measured quantities, and understanding
2638: their role and disentangling the various contributions cannot be
2639: avoided.
2640: %%%%%%%%%%%%%%%%%%%%%%%%%%
2641: \subsubsection{Kraichnan's Model of Turbulent Advection and the
2642: Statistical Objects} \label{s:Kra} The model of passive scalar
2643: advection with rapidly decorrelating velocity field was introduced in
2644: \cite{kra68}. In recent years
2645: \cite{gaw95,che95,ber96,fai96,kra94,lvo94c,ben97} it was shown to be a
2646: fruitful case model for understanding multi-scaling in the statistical
2647: description of turbulent fields. The basic dynamical equation in this
2648: model is for a scalar field $T({\B r},t)$ advected by a random
2649: velocity field ${\B u}({\B x},t)$:
2650: \begin{equation}
2651: \label{advect}
2652: \big[\partial_t - \kappa_0 \nabla^2 +
2653: {\B u}({\B x},t) \cdot {\bf \nabla}\big]
2654: T({\B x},t) = f({\B x},t)\ .
2655: \end{equation}
2656: In this equation $f({\B x},t)$ is the forcing. In Kraichnan's
2657: model the advecting field ${\B u}({\B x},t)$ as well as the
2658: forcing field $f({\B x},t)$ are taken to be Gaussian, time and
2659: space homogeneous, and delta-correlated in time:
2660: $$
2661: \overline{ f({\B x},t) f({\B x}',t') } =
2662: \Phi({\B x}-{\B x}') \delta(t - t')\,, \quad
2663: \langle u^\alpha({\B x},t) u^\beta({\B x}',t') \rangle
2664: = \C W^{\alpha\beta}({\B x}-{\B x}') \delta(t - t').
2665: $$
2666: Here the symbols ~$\overline{\cdots}$~ and ~$\langle \cdots
2667: \rangle$~ stand for independent ensemble averages with respect to
2668: the statistics of $f$ and ${\B u}$ which are given {\em a priori}. We
2669: will study this model in the limit of large Peclet (Pe) number,
2670: Pe$\equiv U_{\Lambda} \Lambda/\kappa_0$, where $U_\Lambda$ is the
2671: typical size of the velocity fluctuations on the outer scale $\Lambda$
2672: of the velocity field. We stress that the forcing is {\em not} assumed
2673: isotropic, and actually the main goal of this section is to study the
2674: statistic of the scalar field under anisotropic forcing.
2675:
2676: The correlation function of the advecting velocity needs further
2677: discussion. It is customary to introduce $\C W^{\alpha \beta}(\B
2678: r)$ via its $\B k$-representation: \be
2679: \label{W1} \C
2680: W^{\alpha\beta}(\B r)=\frac{\e\, D}{\Omega_d} \int \limits
2681: _{\Lambda ^{-1}}^{\lambda^{-1}} \frac{ d^d p} {p^{d+\e}} \,
2682: P^{\alpha\beta}(\B p)\,
2683: \exp (-i \B p\cdot \B r)\,, \quad
2684: P^{\alpha\beta}(\B p) = \Big[\delta_{\alpha\beta}
2685: -\frac {p^\alpha p^\beta}{p^2} \Big] \ee
2686: where
2687: $P^{\alpha\beta}(\B p)$ is the transversal projector, $\Omega_d=(d-1)
2688: \Omega(d)/d $ and $\Omega(d)$ is the volume of the sphere in $d$
2689: dimensions. Equation (\ref{W1}) introduces the four important parameters
2690: that determine the
2691: statistics of the driving velocity field: $\Lambda$ and $\lambda$ are
2692: the outer and inner scales of the driving velocity field
2693: respectively. The scaling exponent $\e$ characterizes the correlation
2694: functions of the velocity field, lying in the interval $[0,2]$. The
2695: factor $D$ is related to the correlation function as
2696: follows:
2697: \be\label{W3} \C W^{\alpha\beta}(0)=D\delta_{\a\b}(\Lambda^\e
2698: -\lambda^\epsilon)\ . \ee The most important property of the driving
2699: velocity field from the point of view of the scaling properties of the
2700: passive scalar is the ``eddy diffusivity" tensor \cite{kra68}
2701: \be
2702: \label{eddy-diff} K^{\alpha\beta}({\B r})
2703: \equiv 2[\C W^{\alpha\beta}(0)-\C W^{\alpha\beta}(\B r)]\ . \ee
2704: The scaling properties of the scalar depend sensitively on the
2705: scaling exponent $\e$ that characterizes the $r$ dependence of
2706: $K^{\alpha\beta}({\B r})
2707: \propto [\Lambda^\e
2708: -\lambda^\epsilon]\delta_{\alpha\beta}$, for
2709: $ r \gg \Lambda$, namely:
2710: \be
2711: K^{\alpha\beta}({\B r})\propto r^\e
2712: \Big[\delta_{\alpha\beta} -\frac{\epsilon}{d-1+\epsilon}{r^\alpha
2713: r^\beta
2714: \over r^2}\Big], \quad \lambda\ll r\ll\Lambda \ .
2715: \label{kappa1}
2716: \ee
2717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2718: \subsubsection{The Statistical Objects}
2719: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2720: In the statistical theory we are interested in the power laws
2721: characterizing the $r$ dependence of the various correlation and
2722: response functions of $T({\B x},t)$ and its gradients. We will focus
2723: on three types of quantities:
2724:
2725: 1) ``Unfused" structure functions of $T({\B x},t)$ are defined as
2726: \begin{eqnarray}
2727: \label{Sunfused}F_T^{(n)}(\B x_1,\B x'_1,\dots \B x_{n},\B x'_{n})
2728: &\equiv &\langle [T(\B x_1,t)-T(\B x'_1,t)]\\ \nonumber \times [T(\B
2729: x_2,t)-T(\B x'_2,t)]&\dots& [T(\B x_n,t)-T(\B x'_n,t)]\rangle \ ,
2730: \end{eqnarray}
2731: and in particular the standard ``fused" structure functions are
2732: $$
2733: S_T^{(n)}({\B r}) \equiv \langle [ T({\B x}+{\B r}, t) - T({\B
2734: x}, t)]^{n} \rangle \ .
2735: $$
2736: In writing this equation we used the fact that the stationary and
2737: space-homogeneous statistics of the velocity and the forcing fields
2738: lead to a stationary and space homogeneous ensemble of the scalar
2739: $T$. If the statistics is also isotropic, then $S^{(n)}_T(\B r)$ becomes a
2740: function of $r$ only, independent of the direction of ${\B r}$. The
2741: ``isotropic scaling exponents'' $\xi^{(n)}_0$ of the structure functions
2742: $$
2743: S_T^{(n)}(r) \propto r^{\xi^{(n)}_0} ,
2744: $$
2745: characterize their $r$ dependence in the limit of large Pe, when
2746: $r$ is in the ``inertial" interval of scales. This range is
2747: $\lambda,\eta\ll r\ll\Lambda, \,L$ where
2748: $\eta = \Lambda \left(\frac{\kappa_0}{D}\right) ^{1/\epsilon}$
2749: is the dissipative
2750: scale of the scalar field,.
2751: When the ensemble is not isotropic we define the exponents
2752: (\ref{eq:perturbative})
2753: by expanding $S^{(n)}_T(\B r)$ according to:
2754: $$
2755: S^{(n)}_T(\B r) = \sum_{jm} S^{(n)}_{T,jm}(r) Y_{jm}(\hat{\Br}) \ ; \qquad
2756: S^{(n)}_{T,jm}(r) \propto r^{\xi^{(n)}_j}
2757: $$
2758: 2) In addition to structure functions we are also interested in
2759: the simultaneous $n\,$th order correlation functions of the
2760: temperature field which is time independent in stationary
2761: statistics:
2762: \be
2763: \label{Fn} \C T^{(n)}(\{\B x_l \}) \equiv \la T(\B
2764: x_1,t)\,T(\B x_2,t)\dots T(\B x_{n},t)\ra \,,
2765: \ee
2766: where we used
2767: the shorthand notation $\{\B x_l \}$ for the whole set of
2768: arguments of $n$th order correlation function $\C T^{(n)}$, $\B
2769: x_1,\B x_2\dots \B x_{n}$.
2770:
2771: 3) Finally, we are interested in correlation functions of the
2772: gradient field $\B \nabla T$. There can be a number of these, and
2773: we denote
2774: $$ \C H^{\alpha_1\dots \alpha_n} (\{\B x_l\}) \equiv
2775: \Big\langle \prod_{i=1}^{n} \Big[\nabla^{\a_i}T(\B x_i,t)\Big]\Big\rangle
2776: \,, $$
2777: The tensor $\C H^{\alpha_1 \dots \alpha_n}$
2778: can be contracted in various ways. For example,
2779: binary contractions $\alpha_1=\alpha_2, \alpha_3=\alpha_4$, {\em
2780: etc.} with $\B x_1=\B x_2, \B x_3=\B x_4$ {\em etc.} produces
2781: the correlation functions of dissipation field $|\B \nabla T|^2$.
2782: Of particular interest is the coordinate independent tensor $\B H^{(n)}$
2783: obtained by taking
2784: all $\B x_i=\B x$:
2785: \begin{equation}
2786: H^{\alpha_1\ldots\alpha_n}={\C H}^{\alpha_1\dots \alpha_n} (\{\B x_i=\B
2787: x\}) \ . \label{defHn}
2788: \end{equation}
2789:
2790: When the ensemble is not isotropic we need to take into account
2791: the angular dependence of $\B x$, and the scaling behavior
2792: consists of multiple contributions arising from anisotropic
2793: effects.
2794: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2795: \subsubsection{The Eulerian calculation}
2796: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2797: The correlation functions of the gradient field $\B H^{(n)}$ of Eq.
2798: (\ref{defHn}) are tensors independent of the coordinates. Nevertheless
2799: their calculation is somewhat heavy, and we do not reproduce it here;
2800: we refer the reader to \cite{ara00} where the calculation is
2801: presented in full detail. The final result of the calculation is for
2802: the projection of $\B H^{(n)}$ onto the $j$ sector of the SO(3)
2803: symmetry group reads
2804: \begin{equation}
2805: \B H^{(n)}_{j}\propto \left(\frac{\Lambda}{\eta}\right)^{n-\xi^{(n)}_j} \ ,
2806: \label{Hscale}
2807: \end{equation}
2808: where the proportionality constant is a tensor in the limit $\eta \ll
2809: \Lambda$. The exponents $\xi^{(n)}_j$ are the same as those found
2810: below for the correlation function in which all the scales are within
2811: the inertial range. The appearance of both renormalization lengths
2812: and the identity of the exponents in inertial and gradient objects is
2813: a consequence of the fusion rules that were explored in \cite{ara00}
2814: with
2815: some care.
2816: The correlation functions ${\mathcal T}^{(n)}$ satisfy the Kraichnan's
2817: equation \cite{kra68}
2818: \begin{eqnarray}
2819: & & \Big[- \kappa_0 \sum_{i=1}^{n} \nabla^2_i+ \frac{1}{2}
2820: \sum_{i,k=1}^{n} K^{\a\b} (\B x_i-\B
2821: x_k)\nabla^\a_i\nabla^\b_k \Big] {\mathcal T}^{(n)}(\{{ \B x}_{l}
2822: \}) \nonumber \\ &&= \frac{1}{2} \sum_{\{i\ne k\}=1}^{n}
2823: \Phi({\B x}_i-{\B x}_k) {\mathcal T}^{(n-2)}(\{{\B x}_l\}_{l\ne
2824: i,k}) \,,
2825: \label{bob1}
2826: \end{eqnarray}
2827: where $\{{\B x}_{l}\}_{l\ne i,k}$ is the set off all ${\B x}_{l}$
2828: with $l$ from 1 to $n$, except of $l= i$ and $l= j$. Substituting
2829: $ K^{\a\b} (\B x)$ from Eqs.~(\ref{W3},
2830: \ref{eddy-diff}) one gets:
2831: \begin{eqnarray}
2832: & & \Big[- \kappa \sum_{i=1}^{n} \nabla^2_i+ \sum_{\{i\ne
2833: k\}=1}^{n} \C W^{\a\b} (\B x_i-\B x_k)\nabla^\a_i\nabla^\b_k
2834: \Big] {\mathcal T}^{(n)}(\{{ \B x}_{l} \}) \nonumber \\
2835: &&=\frac{1}{2} \sum_{\{i\ne k\}=1}^{n} \Phi({\B x}_i-{\B x}_k)
2836: {\mathcal T}^{(n-2)}(\{{\B x}_l\}_{l\ne i,k}) \,,
2837: \label{bob0}
2838: \end{eqnarray}
2839: where $
2840: \kappa =\kappa _0+
2841: D[\Lambda^\e-\lambda^\epsilon]$.
2842: Here we used that in space
2843: homogeneous case $ \sum_{i=1}^{n} \B \nabla_i=0$ and therefore
2844: $$\Big|\sum_{i=1}^{n} \B \nabla_i\Big|^2=
2845: \sum_{i=1}^{n}\nabla^2_i+\sum_{\{i\ne k\}=1}^{n}
2846: \nabla^\a_i\nabla^\b_k=0\ .
2847: $$
2848:
2849: In this section we consider the zero-modes of Eq.~(\ref{bob1}). In
2850: other words we seek solutions $Z^{(n)}(\{\B x_l\})$ which in the
2851: inertial interval solve the homogeneous equation
2852: \begin{equation}
2853: \sum_{i\ne k=1}^{n}
2854: K^{\a\b} (\B x_i-\B x_k)\nabla^\a_i\nabla^\b_k
2855: Z^{(n)}(\{{ \B x}_{l} \}) =0 \ . \label{zeromodes}
2856: \end{equation}
2857: We allow anisotropy on the large scales. Since all the operators here
2858: are isotropic and the equation is linear, the solution space foliate
2859: into sectors ${j,m}$ corresponding the the irreducible
2860: representations of the SO($d$) symmetry group. Accordingly we write
2861: the wanted solution in the form
2862: $$
2863: Z^{(n)}(\{{ \B r}_{l} \})= \sum_{j,m}Z^{(n)}_{jm}(\{{ \B r}_{l}
2864: \})
2865: \ , $$
2866: where $Z^{(n)}_{jm}$ are functions composed of irreducible
2867: representations of SO($d$) with definite $j,m$. Each of these
2868: components is now expanded in $\epsilon$. In other words, we write, in
2869: the notation of Ref.\cite{ber96},
2870: $$
2871: Z^{(n)}_{jm} =E^{(n)}_{jm}+\epsilon
2872: G^{(n)}_{jm}+O(\epsilon^2) \ .
2873: $$
2874: For $\epsilon=0$ Eq.~(\ref{zeromodes}) simplifies to
2875: \begin{equation}
2876: \sum_{i=1}^{n} \nabla_i^2 E^{(n)}_{jm}(\{{ \B r}_{l} \}) =0
2877: \ , \label{EqE}
2878: \end{equation}
2879: for any value of $j,m$. Next we expand the operator in
2880: Eq.~(\ref{zeromodes}) in $\epsilon$ and collect the terms of
2881: $O(\epsilon)$:
2882: \begin{equation}
2883: \sum_{i=1}^{n} \nabla_i^2 G^{(n)}_{jm}(\{{ \B r}_{l}
2884: \})=V_{n}E^{(n)}_{jm}(\{{ \B r}_{l} \}) \ , \label{firsto}
2885: \end{equation}
2886: where $\epsilon V_{n}$ is the first order term in the expansion
2887: of the operator in (\ref{zeromodes}):
2888: \begin{equation}
2889: V_{n} \equiv \sum_{i\ne
2890: k=1}^{n}\Big[\delta^{\alpha\beta}\log(r_{ik})-{r_{ik}^\alpha
2891: r_{ik}^\beta\over
2892: (d-1)r_{ik}^2}\Big]\nabla_i^\alpha\nabla_k^\beta \ , \label{defVn}
2893: \end{equation}
2894: where $\B r_{ik}\equiv \B x_i-\B x_k$.
2895:
2896: In solving Eq.~(\ref{EqE}) we are led by the following
2897: considerations: we want scale invariant solutions, which are
2898: powers of $\B r_{ik}$. We want analytic solutions, and thus we
2899: are limited to polynomials. Finally we want solutions that
2900: involve all the $n$ coordinates for the function
2901: $E^{(n)}_{jm}$; solutions with fewer coordinates do not
2902: contribute to the structure functions (\ref{Sunfused}). To see
2903: this note that the structure function is a linear combination of
2904: correlation functions. This linear combination can be represented
2905: in terms of the difference operator $\delta_l(\B x,\B x')$
2906: defined by:
2907: \begin{equation}
2908: \delta_l(\B x,\B x') {\mathcal T}^{(n)}(\{\B x_k\})\equiv {\mathcal
2909: T}^{(n)}(\{\B
2910: x_k\})|_{\B x_l=\B x} -{\mathcal T}^{(n)}(\{\B x_k\})|_{\B x_l=\B x'} \ .
2911: \label{diffop}
2912: \end{equation}
2913: Then,
2914: \begin{equation}
2915: S_T^{(n)}(\B x_1,\B x'_1\dots \B x_{n}\B x'_{n}) = \prod_l \delta_l(\B
2916: x_l,\B x'_l) {\mathcal T}^{(n)}(\{\B x_k\}) \ . \label{SinF}
2917: \end{equation}
2918:
2919: Accordingly, if ${\mathcal T}^{(n)}(\{\B x_k\})$ does not depend on $\B
2920: x_i$, then $\delta_i (\B r_i,\B r'_i){\mathcal T}^{(n)}(\{\B r_k\})=0$
2921: identically. Since the difference operators commute, we can have
2922: no contribution to the structure functions from parts of ${\mathcal
2923: T}^{(n)}$ that depend on less than $n$ coordinates. Finally we want the
2924: minimal polynomial because higher order ones are negligible in
2925: the limit $r_{ik}\ll \Lambda$. Accordingly, $E^{(n)}_{jm}$
2926: with $j\le n$ is a polynomial of order $n$. Following the procedure
2927: outlined in Appendix~(\ref{app:d-dim})
2928: we can write the most general form of
2929: $E^{(n)}_{jm}$, up to an arbitrary factor, as
2930: \begin{equation}
2931: E^{(n)}_{jm}=x_1^{\alpha_1}\dots x_{n}^{\alpha_{n}}
2932: B^{\alpha_1\dots \alpha_{n}}_{n,jm} +[\dots] \ ,
2933: \end{equation}
2934: where $[\dots]$ stands for all the terms that contain less than
2935: $n$ coordinates; these do not appear in the structure functions
2936: but maintain the translational invariance of our quantities. Note that in
2937: this case we carry
2938: the index $n$ in the tensor basis functions since the theory mixes basis
2939: functions
2940: of different orders. The
2941: appearance of the tensor $B^{\alpha_1\dots
2942: \alpha_n}_{n,jm}$
2943: is justified by the fact
2944: that $E^{(n)}_{jm}$ must be symmetric to permutations of any
2945: pair of coordinates on the one hand, and it has to belong to the
2946: $jm$ sector on the other hand. This requires the
2947: appearance of the fully symmetric tensor (\ref{Birr}).
2948: In light of Eqs.~(\ref{firsto}-\ref{defVn}) we seek solution for
2949: $G^{(n)}_{j} (\{{ \B r}_{k}\})$ of the form
2950: \begin{equation}
2951: G^{(n)}_{jm}(\{{ \B r}_{k}\})=\sum_{i\ne
2952: l}H^{il}_{jm}(\{{ \B r}_{k}\}) \log(r_{il})
2953: +H_{jm}(\{{ \B r}_{k}\}) \ , \label{ansatzG}
2954: \end{equation}
2955: where $H^{il}_{jm}(\{{ \B r}_{k}\})$ and
2956: $H_{jm}(\{{ \B r}_{k}\})$ are polynomials of degree $n$.
2957: The latter is fully symmetric in the coordinates. The former is
2958: symmetric in $r_i$, $r_l$ and separately in all the other
2959: $\{r_k\}_{k\ne l,i}$.
2960: Substituting Eq.~(\ref{ansatzG}) into Eq.~(\ref{firsto}) and
2961: collecting terms of the same type yields three equations:
2962: \begin{eqnarray}
2963: &&\sum_i \nabla_i^2 H^{lk}_{jm}=\nabla_l\cdot \nabla_k
2964: E_{jm}^{(n)}\ ,
2965: \label{zero1}\\
2966: &&\big[d-2+\B r_{lk}\cdot (\B \nabla_l-\B
2967: \nabla_k)\big]H^{lk}_{jm} +\frac{r_{lk}^\alpha
2968: r_{lk}^\beta \nabla_l^\alpha \nabla_k ^\beta}{2d-2} E_{jm}^{(n)}
2969: =-\frac{r_{lk}^2 K_{jm}^{lk}}{2}\ ,
2970: \label{zero2}\\ &&\sum_i \nabla_i^2 H_{jm} =\sum_{l\ne
2971: k}K_{jm}^{lk} \ . \nonumber
2972: \end{eqnarray}
2973: Here $K_{jm}^{lk}$ are polynomials of degree $n-2$ which
2974: are separately symmetric in the $l,k$ coordinates and in all the
2975: other coordinates except $l,k$. In Ref. \cite{ber96} it was proven
2976: that for $j=0$ these equations possess a unique solution. The
2977: proof follows through unchanged for any $j \ne 0$, and we thus
2978: proceed to finding the solution.
2979:
2980: By symmetry we can specialize the discussion to $l=1$, $k=2$. In
2981: light of Eq.~(\ref{zero2}) we see that $H^{12}_{jm}$
2982: must have at least a quadratic contribution in $r_{12}$. This
2983: guarantees that (\ref{ansatzG}) is nonsingular in the limit
2984: $r_{12}\to 0$. The only part of $H^{12}_{jm}$ that will
2985: contribute to structure functions must contain $\B r_3\dots \B
2986: r_{n}$ at least once. Since $H^{12}_{jm}$ has to be a
2987: polynomial of degree $n$ in the coordinates, it must be of the
2988: form
2989: \begin{equation}
2990: H^{12}_{jm}=r_{12}^{\alpha_1}r_{12}^{\alpha_2}
2991: r_3^{\alpha_3}\dots r_{n}^{\alpha_{n}} C^{\alpha_1\alpha_2\dots
2992: \alpha_{n}} +[\dots]_{1,2} \ , \label{H}
2993: \end{equation}
2994: where $[\dots]_{1,2}$ contains terms with higher powers of
2995: $r_{12}$ and therefore do not contain some of the other
2996: coordinates $r_3\dots r_{n}$. Obviously such terms are
2997: unimportant for the structure functions. Since
2998: $H^{12}_{jm}$ has to be symmetric in $\B r_1,\B r_2$ and
2999: $\B r_3\dots \B r_{n}$ separately, and it has to belong to an
3000: $jm$ sector, we conclude that the constant tensor $\B C$
3001: must have the same symmetry and must belong to the same sector.
3002: Consulting Appendix~(\ref{app:d-dim}) the most general form of $\B C$ is
3003: \begin{equation}
3004: C^{\alpha_1\alpha_2\dots \alpha_{n}} =
3005: aB_{n,jm}^{\alpha_1\alpha_2\dots
3006: \alpha_{n}}+b\delta^{\alpha_1\alpha_2}
3007: B_{n-2,jm}^{\alpha_3\alpha_4\dots
3008: \alpha_{n}}+c\sum_{i\ne
3009: l>2}\delta^{\alpha_1\alpha_i}\delta^{\alpha_2\alpha_l}
3010: B_{n-4,jm}^{\alpha_3\alpha_4\dots \alpha_{n}} \ .
3011: \end{equation}
3012: Substituting in Eq.~(\ref{zero2}) one find
3013: $$
3014: (d+2)H^{12}_{jm}+\frac{r_{12}^{\alpha_1}r_{12}^{\alpha_2}r_3^{
3015: \alpha_3} \dots r_{n}^{\alpha_{n}}}{2d-2} B_{n,jm}^{\alpha_1\dots
3016: \alpha_{n}} +\frac{1}{2}r_{12}^{\alpha_1}r_{12}^{\alpha_2}
3017: \delta^{\alpha_1\alpha_2}K^{1,2}_{jm}=[\dots]_{1,2} \ .
3018: $$
3019: Substituting Eq.~(\ref{H}) and demanding that coefficients of the
3020: term $r_1^{\alpha_1}\dots r_{n}^{\alpha_{n}}$ will sum up to
3021: zero, we obtain
3022: $$
3023: -2(d+2)a -\frac{2}{2d-2}=0 \ , \quad -2(d+2)c=0\ ;
3024: \Longrightarrow c=0 \ . $$
3025: The coefficient $b$ is not determined from this equation due to
3026: possible contributions from the unknown last term. We determine
3027: the coefficient $b$ from Eq.~(\ref{zero1}). After substituting
3028: the forms we find
3029: $$
3030: 4\delta^{\alpha_1\alpha_2} r_3^{\alpha_3}\dots r_{n}^{\alpha_{n}}
3031: [a B_{n,jm}^{\alpha_1\dots
3032: \alpha_{n}}+b\delta^{\alpha_1\alpha_2}
3033: B_{n-2,jm}^{\alpha_3\alpha_4\dots\alpha_{n}}]=\delta^{\alpha_1
3034: \alpha_2}
3035: r^{\alpha_3}\dots r^{\alpha_{n}} B_{n,jm}^{\alpha_1\dots
3036: \alpha_{n}}+ [\dots]_{1,2} \ .
3037: $$
3038: Recalling the identity (\ref{iden1}) we obtain
3039: $
3040: b=\frac{z_{n,j}}{4d}[1-4a] \ . $
3041: Finally we find that $a$ is $n,j$-independent,
3042: $
3043: a=-\frac{1}{2(d+2)(d-1)}\ , $
3044: whereas $b$ does depend on $n$ and $j$, and we therefore
3045: denote it as $b_{n,j}$
3046: $$
3047: b_{n,j}=\frac{(d+1)}{4(d+2)(d-1)} z_{n,j} \ . $$
3048: In the next Subsect. we compute from these results the scaling
3049: exponents in all the sectors of the SO($d$) symmetry group.
3050: \subsubsection{The Scaling Exponents of the
3051: Structure Functions}
3052: We now wish to show that the solution for the zero modes of the
3053: correlation functions $F^{(n)}_T$ (i.e $Z^{(n)}$) result in
3054: homogeneous structure functions $S^{(n)}_T$. In every sector
3055: $j,m$ we compute the scaling exponents, and show that
3056: they are independent of $m$. Accordingly the scaling
3057: exponents are denoted $\xi^{(n)}_{j}$, and we compute them to
3058: first order in $\epsilon$.
3059: Using (\ref{diffop}) and (\ref{SinF}), the structure function is
3060: given by:
3061: \begin{eqnarray} S^{(n)}_{T,jm}(\B r_1,\Obr_1;
3062: \ldots ; \B r_{n},\Obr_{n} )= \Delta_1^{\alpha_1} \ldots
3063: \Delta_{n}^{\alpha_{n}} B_{n,jm}^{\alpha_1\dots
3064: \alpha_{n}} + \nonumber \\
3065: \epsilon \sum_{i \neq l}
3066: \stackrel{\mbox{no } i,l} {\overbrace{ \Delta_1^{\alpha_1} \ldots
3067: \Delta_{n}^{\alpha_{n}} } } f^{\alpha_i\alpha_l} (\B r_i,\Obr_i,
3068: \B r_l,\Obr_l) [a B_{n,jm}^{\alpha_1\dots \alpha_{n}} +b\Ctensor{i}{l}]
3069: \ , \nonumber \end{eqnarray}
3070: where $\Delta_i^{\alpha_i} \equiv
3071: r_i^{\alpha_i}-\Or_i^{\alpha_i}$, and the function $f$ is defined
3072: as:
3073: \begin{eqnarray}
3074: f^{\alpha_i\alpha_l} (\B r_i,\Obr_i,\B r_l,\Obr_l) & \equiv &
3075: (r_i-r_l)^{\alpha_i}(r_i-r_l)^{\alpha_l} \ln|\B r_i -\B r_l| \nonumber\\&+&
3076: (\Or_i - \Or_l)^{\alpha_i}(\Or_i - \Or_l)^{\alpha_l} \ln|\Obr_i -
3077: \Obr_l| \\ & - & (r_i - \Or_l)^{\alpha_i}(r_i - \Or_l)^{\alpha_l}
3078: \ln |\B r_i - \Obr_l| \nonumber\\&-& (\Or_i - r_l)^{\alpha_i}(\Or_i -
3079: r_l)^{\alpha_l}\ln |\Obr_i-{\B r}_l| \ .
3080: \end{eqnarray}
3081: The scaling exponent of $S^{(n)}_{T,jm}$ can be found
3082: by multiplying all its coordinates by $\mu$. A direct calculation
3083: yields:
3084: \begin{eqnarray*}
3085: S^{(n)}_{T,jm}(\mu \B r_1, \mu \Obr_1 ; \ldots ) =
3086: \mu^{n}S^{(n)}_{T,jm}(\B r_1,\Obr_1 ; \ldots) -
3087: 2\epsilon \mu^{n}\ln \mu \sum_{i\neq l}\stackrel{\mbox{no } i,l} {
3088: \overbrace{\Delta_1^{\alpha_1} \ldots \Delta_{n}^{\alpha_{n}} } }
3089: \Delta_i^{\alpha_i}\Delta_l^{\alpha_l} \\ \times [a\Btensor +
3090: b_{n,j}\Ctensor{i}{l}] + O(\epsilon^2), \\ = \mu^{n}S^{(n)}_{jm}(\B
3091: r_1,\Obr_1 ; \ldots ) - 2\epsilon \mu^{n}
3092: \ln \mu \Delta_1^{\alpha_1} \ldots \Delta_{n}^{\alpha_{n}}
3093: \times \sum_{i\neq l} [a\Btensor + b{n,j}\Ctensor{i}{l} +
3094: O(\epsilon^2) \ .
3095: \end{eqnarray*}
3096: Using (\ref{iden2}), we find that $ \sum\limits_{i\neq l}
3097: [a\Btensor + b_{n,j}\Ctensor{i}{l}] = [n(n-1)a+b_{n,j}] \Btensor $, and
3098: therefore, we finally obtain:
3099: \begin{eqnarray*}
3100: S^{(n)}_T(\mu \B r_1, \mu \Obr_1 ;\ldots ) = \mu^{n} \left\{
3101: 1 - 2\epsilon [n(n-1)a+b_{n,j}] \ln \mu \right\} S^{(n)}_T(\B r_1,
3102: \Obr_1 ;\ldots ) + O(\epsilon^2) \\ = \mu^{\xi^{(n)}_{j}}
3103: S^{(n)}_T(\B r_1, \Obr_1 ; \ldots ) + O(\epsilon^2)\ ,
3104: \end{eqnarray*}
3105: The result of the scaling exponent is:
3106: $$
3107: \xi_{j}^{(n)} =
3108: n-2\epsilon[-\frac{n(n-1)}{2(d+2)(d-1)}+\frac{(d+1)}{4(d+2)(d-1)}
3109: z_{n,j}] +O(\epsilon^2)
3110: $$
3111: from which follows (\ref{eq:perturbative}).
3112: This is the
3113: final result of this calculation.\\
3114: It is noteworthy that this result is in full agreement with
3115: (\ref{Hscale}), even though the scaling
3116: exponents that appear in these result refer to different
3117: quantities. The way to understand this is the fusion rules that
3118: are discussed next.
3119: %%%%%%%%%%%%%%%%%%%%%
3120: \subsubsection{Fusion Rules}
3121: The fusion rules address the asymptotic properties of the fully
3122: unfused structure functions when two or more of the coordinates
3123: are approaching each other, whereas the rest of the coordinates
3124: remain separated by much larger scales. A full discussion of the
3125: fusion rules for the Navier-Stokes and the Kraichnan model can be
3126: found in \cite{lvo96,fai96,ben98}. In this section we quote the fusion rules
3127: that
3128: were
3129: derived in Ref. \cite{ara00} directly from the zero modes that were computed
3130: to $O(\epsilon)$, in all the sectors of the symmetry group. In
3131: other words, we are after the dependence of the structure
3132: function $ S^{(n)}_T({\bf r}_{1},\overline{{\bf r}}_{1};\ldots
3133: )$ on its first $p$ pairs of coordinates ${\bf
3134: r}_{1},\overline{{\bf r}}_{1};\ldots ;{\bf r}_{p}, \overline{{\bf
3135: r}}_{p}$ in the case where these points are very close to each
3136: other compared to their distance from the other $n-p$ pairs of
3137: coordinates. Explicitly, we consider the case where ${\bf r}_{1},
3138: \overline{{\bf r}}_{1};\ldots ;{\bf r}_{p},\overline{{\bf
3139: r}}_{p}\ll {\bf r} _{p+1},\overline{{\bf r}}_{p+1};\ldots ;{\bf
3140: r}_{n},\overline{{\bf r}}_{n}$. (We have used here the property
3141: of translational invariance to put the center of mass of the
3142: first $2p$ coordinates at the origin. The full calculation is
3143: presented in \cite{ara00}, with the final result (to $O(\epsilon)$)
3144: $$
3145: S^{(n)}_{T,jm}({\bf r}_{1},\overline{{\bf
3146: r}}_{1};\ldots ;{\bf r}_{n}, \overline{{\bf r}}_{n})
3147: =\sum_{l=l_{\rm max}}^p\sum_{m^{\prime }}\psi_{l,m^{\prime
3148: }}S^{(p)}_{T,lm^{\prime }}({\bf r}_{1},\overline{{\bf
3149: r}}_{1};\ldots ;{\bf r}_{p},\overline{ {\bf r}}_{p}) \ .
3150: $$
3151: In this expression the quantity $ \psi_{l,m^{\prime }}$
3152: is a tensor function of all the coordinates that remain separated
3153: by large distances, and
3154: $$
3155: l_{\rm max}={\rm max}\{0,p+j-n\} \ , \quad j \le n.
3156: $$
3157: We have shown that the LHS has a homogeneity exponent
3158: $\xi^{(n)}_j$. The RHS is a product of functions with
3159: homogeneity exponents $\xi^{(p)}_l$ and the functions $
3160: \psi_{l,m^{\prime}}$. Using the linear independence of the
3161: functions $S^{(p)}_{T,lm'}$ we conclude that $
3162: \psi_{l,m^{\prime}}$ must have homogeneity exponent
3163: $\xi^{(n)}_j-\xi^{(p)}_l$. This is precisely the
3164: prediction of the fusion rules, but in each sector separately.
3165: One should stress the intuitive meaning of the fusion rules. The
3166: result shows that when $p$ coordinates approach each other, the
3167: homogeneity exponent corresponding to these coordinates becomes
3168: simply $\xi^{(p)}_l$ as if we were considering a $p$-order
3169: correlation function. The meaning of this result is that $p$
3170: field amplitudes measured at $p$ close-by coordinates in the
3171: presence of $n-p$ field amplitudes determined far away behave
3172: scaling-wise like $p$ field amplitudes in the presence of
3173: anisotropic boundary conditions.
3174: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3175: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3176: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3177: \subsubsection{The Lagrangian Approach to Anomalous Scaling}
3178: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3179: An elegant approach to the correlation functions is furnished by
3180: Lagrangian dynamics \cite{ap01,sra74,gat98,gat98b,fmnv99}. In this
3181: formalism one recognizes that the actual value of the scalar at
3182: position $\B x$ at time $t$ is determined by the action of the
3183: forcing along the Lagrangian trajectory from $t=-\infty$ to $t$:
3184: \begin{equation}
3185: T(\B x_0,t_0)=\int_{-\infty}^{t_0} dt
3186: \left< f(\B x(t),t) \right>_{\B \eta} \ ,
3187: \label{path}
3188: \end{equation}
3189: with the trajectory $\B x(t)$ obeying
3190: \be
3191: \B x(t_0) = \B x_0 \ ,\qquad
3192: \partial_t \B x(t) = \B u(\B x(t),t) +
3193: \sqrt{2 \kappa}{\B \eta}(t) \ ,
3194: \label{traject1}
3195: \ee
3196: and $\B \eta$ is a vector of zero-mean independent Gaussian
3197: white random variables, $\left< \eta^\alpha (t) \eta^\beta (t')
3198: \right> = \delta^{\alpha\beta} \delta(t-t')$. With this in mind,
3199: we can rewrite $S_T^{(2n)}$ of Eq. (\ref{Sunfused}) by substituting each
3200: factor
3201: of $T(\B x_i)$ by its representation (\ref{path}). Performing the
3202: averages over the random forces, we end up with
3203: \begin{eqnarray}
3204: S_T^{(2n)}(\B x_1, \ldots, \B x_{2n},t_0) =
3205: \Big< \int_{-\infty}^{t_0} dt_1 \cdots dt_n
3206: \Big[ \phi(\B x_1(t_1)-\B x_2(t_1)) \cdots \\
3207: \times \phi(\B x_{2n-1}(t_n)-\B x_{2n}(t_n))
3208: + \hbox{permutations} \Big] \Big>_{\B u,\{\B \eta_i\}} \ ,
3209: \label{FnXi}
3210: \end{eqnarray}
3211: To understand the averaging procedure recall that each of the
3212: trajectories $\B x_i$ obeys an equation of the form
3213: (\ref{traject1}), where $\B u$ as well as $\{\B
3214: \eta_i\}_{i=1}^{2n}$ are independent stochastic variables whose
3215: correlations are given above. Alternatively, we refer the reader
3216: to section II of \cite{fmnv99}, where the above analysis is carried
3217: out in detail. Here we follow the derivation of Ref. \cite{ap01}.
3218: In considering Lagrangian trajectories of {\em groups} of
3219: particles, we should note that every initial configuration is
3220: characterized by a {\em center of mass}, say $\B R$, a {\em
3221: scale} $s$ (say the radius of gyration of the cluster of
3222: particles) and a {\em shape} $\B Z$. In ``shape" we mean here all
3223: the degrees of freedom other than the scale and $\B R$: as many
3224: angles as are needed to fully determine a shape, in addition to
3225: the Euler angles that fix the shape orientation with respect to a
3226: chosen frame of coordinates. Thus a group of $2n$ positions $\{\B
3227: x_i\}$ will be sometime denoted below as $\{\B R,s,\B Z\}$.
3228:
3229: One component in the evolution of an initial configuration is a
3230: rescaling of all the distances which increase on the average like
3231: $t^{1/\xi_2}$; this rescaling is analogous to Richardson
3232: diffusion. The exponent $\xi_2$ which determines the scale
3233: increase is also the characteristic exponent of the second order
3234: structure function \cite{kra68}. This has been related to the
3235: exponent $\epsilon$ of (\ref{kappa1}) according to $\xi_2=2-\epsilon$.
3236: After factoring out this overall expansion we are left with a
3237: normalized `shape'. It is the evolution of this shape that
3238: determines the anomalous exponents.
3239:
3240: Consider a final shape ${\B Z}_{0}$ with an overall scale $s_0$
3241: which is realized at $t=0$. This shape has evolved during
3242: negative times. We fix a scale $s>s_0$ and examine the shape when
3243: the configuration reaches the scale $s$ for the last time before
3244: reaching the scale $s_0$. Since the trajectories are random the
3245: shape $\B Z$ which is realized at this time is taken from a
3246: distribution $\gamma(\B Z;\B Z_0,s\to s_0)$. As long as the
3247: advecting velocity field is scale invariant, this distribution
3248: can depend only on the ratio
3249: $s/s_0$.
3250:
3251: Next, we use the shape-to-shape transition probability to define
3252: an operator $\hat \gamma (s/ s_0)$ on the space of functions
3253: $\Psi(\B Z)$ according to
3254: $$
3255: [\hat {\B \gamma} (s/ s_0) \Psi](Z_0) =
3256: \int d\B Z \gamma(\B Z;\B Z_0,s \to s_0) \Psi(\B Z).
3257: $$
3258: We will be interested in the eigenfunction and eigenvalues of
3259: this operator. This operator has two important properties. First,
3260: for an isotropic statistics of the velocity field the operator is
3261: isotropic. This means that this operator commutes with all
3262: rotation operators on the space of functions $\Psi(\B Z)$. In
3263: other words, if ${\mathcal O}_\Lambda$ is the rotation operator that
3264: takes the function $\Psi(\B Z)$ to the new function
3265: $\Psi(\Lambda^{-1} \B Z)$, then
3266: $ {\mathcal O}_\Lambda \hat{\B \gamma}
3267: =\hat{\B \gamma} {\mathcal O}_\Lambda $.
3268: This property follows from the obvious symmetry of the Kernel
3269: $\gamma(\B Z;\B Z_0,s\to s_0)$ to rotating $\B Z$ and $\B Z_0$
3270: simultaneously. Accordingly the eigenfunctions of $\hat {\B
3271: \gamma}$ can be classified according to the irreducible
3272: representations of SO($3$) symmetry group.
3273: Because in this section we are not computing explicitly the exponents
3274: we do not need to present the precise form of the eigenfunctions
3275: and we will denote them for simplicity as
3276: $B_{qj m}(\B Z)$.
3277: The second important property of $\hat {\B \gamma}$ follows from
3278: the $\delta$-correlation in time of the velocity field. Physically
3279: this means that the future trajectories of $n$ particles are
3280: statistically independent of their trajectories in the past.
3281: Mathematically, it implies for the kernel that
3282: $$
3283: \gamma(\B Z;\B Z_0,s\to s_0) = \int d\B Z_1 \gamma(\B Z;\B Z_1,s\to s_1)
3284: \gamma(\B Z_1;\B Z_0,s_1\to s_0) \ , \quad s>s_1>s_0
3285: $$
3286: and in turn, for the operator, that
3287: $$
3288: \hat {\B \gamma} (s/ s_0) =\hat {\B \gamma} (s/ s_1)\hat {\B
3289: \gamma} (s_1/ s_0) \ .
3290: $$
3291: Accordingly, by a successive application of $\hat{\B
3292: \gamma}(s/s_0)$ to an arbitrary eigenfunction, we get that the
3293: eigenvalues of $\hat{\B \gamma}$ have to be of the form
3294: $\alpha_{q,j}=(s/s_0)^{\xi^{(2n)}_{j}}$:
3295: \begin{equation}
3296: ({s \over s_0})^{\xi^{(2n)}_{j}} B_{qj m}(\B Z_0) =
3297: \int d \B Z\gamma(\B Z; \B Z_0, s \to s_0) B_{qj m}(\B Z)
3298: \label{eig}
3299: \end{equation} From Schur's lemmas one can prove that the eigenvalues do not
3300: depend on
3301: $m$. On the other hand they can still be a function of $q$
3302: but for simplicity of notation we do not explicitly carry the $q$
3303: index in $\xi$.\\
3304: To proceed we want to introduce into the averaging process in
3305: (\ref{FnXi}) by averaging over Lagrangian trajectories of the
3306: $2n$ particles. This will allow us to connect the shape dynamics
3307: to the statistical objects. To this aim consider any set of
3308: Lagrangian trajectories that started at $t=-\infty$ and end up at
3309: time $t=0$ in a configuration characterized by a scale $s_0$ and
3310: center of mass $\B R_0=0$. A full measure of these have evolved
3311: through the scale $L$ or larger. Accordingly they must have
3312: passed, during their evolution from time $t=-\infty$ through a
3313: configuration of scale $s>s_0$ at least once. Denote now
3314: $
3315: \mu_{2n}(t,R,\B Z;s\to s_0,\B Z_0)dtd\B Rd\B Z
3316: $
3317: as the probability that this set of $2n$ trajectories crossed the
3318: scale $s$ for the last time before reaching $s_0,\B Z_0$, between
3319: $t$ and $t+dt$, with a center of mass between $\B R$ and $\B
3320: R+d\B R$ and with a shape between $\B Z$ and $\B Z+d\B Z$.
3321:
3322: In terms of this probability we can rewrite Eq.(\ref{FnXi})
3323: (displaying, for clarity, $\B R_0=0$ and $t=0$) as
3324: \begin{eqnarray}
3325: && S_T^{(2n)}(\B R_0=0,s_0,\B Z_0,t=0) = \int d\B Z\int_{-\infty}^0 dt \int
3326: d\B
3327: R
3328: \mu_{2n}(t,R,\B Z; s \to s_0, \B Z_0) \nonumber \\
3329: &&\times\left< \int_{-\infty}^{0}dt_1\cdots dt_n
3330: \Big[ \phi(\B x_1(t_1)-\B x_2(t_1)) \cdots
3331: \phi(\B x_{2n-1}(t_n) - \B x_{2n}(t_n)) +\hbox{perms}\Big]
3332: \Big \vert (s; \B R, \B Z, t) \right> _{\B u,\B \eta_i} \ .
3333: \label{Fntraj}
3334: \end{eqnarray}
3335: The meaning of the conditional averaging is an averaging over all
3336: the realizations of the velocity field and the random $\B \eta_i$
3337: for which Lagrangian trajectories that ended up at time $t=0$ in
3338: $\B R=0,s_0, \B Z_0$ passed through $\B R ,s,\B Z$ at time $t$.
3339:
3340: Next, the time integrations in the above equation are split to the
3341: interval $[-\infty,t]$ and $[t,0]$ giving rise to $2^n$ different
3342: contributions:
3343: $$
3344: \int_{-\infty}^t dt_1 \cdots \int_{-\infty}^t dt_n +
3345: \int_{t}^0 dt_1 \int_{-\infty}^{t} dt_2 \cdots \int_{-\infty}^{t} dt_n
3346: + \dots
3347: $$
3348: Consider first the contribution with $n$ integrals in the domain
3349: $[-\infty,t]$. It follows from the delta-correlation in time of
3350: the velocity field, that we can write
3351: \begin{eqnarray}
3352: && \left< \int_{-\infty}^t dt_1\cdots dt_n
3353: \Big[ \phi(\B x_1(t_1)-\B x_2(t_1)) \cdots
3354: \phi(\B x_{2n-1}(t_n)-\B x_{2n}(t_n)) + \hbox{perms} \Big]
3355: \Big \vert (s; \B R, \B Z,t) \right> _{\B u,\B \eta_i} \nonumber \\
3356: && = \left< \int_{-\infty}^t dt_1 \cdots dt_n
3357: \Big[\phi(\B x_1(t_1)-\B x_2(t_1)) \cdots \phi(\B x_{2n-1}(t_n) -
3358: \B x_{2n}(t_n)) + \hbox{perms} \Big]
3359: \right> _{\B u,\B \eta_i} \nonumber \\
3360: && = S^{(2n)}_T(\B R, s, \B Z, t)=S_T^{(2n)}(s, \B Z) \ .
3361: \end{eqnarray}
3362: The last equality follows from translational invariance in
3363: space-time. Accordingly the contribution with $n$ integrals in
3364: the domain $[-\infty,t]$ can be written as
3365: $$
3366: \int d\B Z S_T^{(2n)}(s,\B Z) \int_{-\infty}^0 dt \int d\B R~~
3367: \mu_{2n}(t,R,\B Z;s\to s_0,\B Z_0) \ .
3368: $$
3369: We identify the shape-to-shape transition probability:
3370: \begin{equation}
3371: \gamma(\B Z;\B Z_0,s \to s_0)=\int_{-\infty}^0 dt\int d\B R~~
3372: \mu_{2n}(t,R,\B Z;s \to s_0, \B Z_0) \ .
3373: \end{equation}
3374: Finally, putting all this added wisdom back in Eq.(\ref{Fntraj})
3375: we end up with
3376: \begin{equation}
3377: S_T^{(2n)}(s_0, \B Z_0) = I + \int d\B Z \gamma(\B Z; \B Z_0, s \to s_0)
3378: S_T^{(2n)}(s,\B Z)\ . \label{crucial}
3379: \label{split}
3380: \end{equation}
3381: Here $I$ represents all the contributions with one or more time
3382: integrals in the domain $[t,0]$. The key point now is that only
3383: the term with $n$ integrals in the domain $[-\infty,t]$ contains
3384: information about the evolution of $2n$ Lagrangian trajectories
3385: that probed the forcing scale $L$. Accordingly, the term denoted
3386: by $I$ cannot contain information about the leading anomalous
3387: scaling exponent belonging to $F_{2n}$, but only of lower order
3388: exponents. The anomalous scaling dependence of the LHS of
3389: Eq.(\ref{crucial}) has to cancel against the integral containing
3390: $F_{2n}$ without the intervention of $I$.
3391:
3392: Representing now
3393: \begin{eqnarray}
3394: S_T^{(2n)}(s_0,\B Z_0) &=& \sum_{qj m} a_{q,j m}(s_0)
3395: B_{qj m}(\B Z_0)\ ,\nonumber \\
3396: S_T^{(2n)}(s,\B Z) &=& \sum_{qj m} a_{q, j m}(s)
3397: B_{qj m}(\B Z) \ ,\nonumber \\
3398: I&=& \sum_{qj m} I_{q j m} B_{qj m}(\B Z_0)
3399: \end{eqnarray}
3400: and substituting on both sides of Eq.(\ref{crucial}) and using
3401: Eq.(\ref{eig}) we find, due to the linear independence of the
3402: eigenfunctions $B_{qj m}$
3403: $$ a_{q, j m}(s_0) = I_{q j m} +
3404: \left(\frac{s}{s_0}\right)^{\xi^{(2n)}_{j}} a_{q, j m}(s).
3405: $$ To leading order the contribution of $I_{q j m}$ is neglected,
3406: leading to the conclusion that {\em the spectrum of anomalous
3407: exponents of the correlation functions is determined by the
3408: eigenvalues of the shape-to-shape transition probability
3409: operator}. Calculations show that the leading exponent in the
3410: isotropic sector is always smaller than the leading exponents in
3411: all other sectors. This gap between the leading exponent in the
3412: isotropic sector to the rest of the exponents determines the rate
3413: of decay of anisotropy upon decreasing the scale of observation.
3414:
3415: The derivation presented above has used explicitly the properties
3416: of the advecting field, in particular the $\delta$-correlation in
3417: time. Accordingly it cannot be immediately generalized to more
3418: generic situations in which there exist time correlations.
3419: Nevertheless we find it pleasing that at least in the present
3420: case we can trace the physical origin of the exponents anomaly,
3421: and connect it to the underlying dynamics. In more generic cases
3422: the mechanisms may be more complicated, but one should still keep
3423: the lesson in mind - higher order correlation functions depend on
3424: many coordinates, and these define a configuration in space. The
3425: scaling properties of such functions may very well depend on how
3426: such configurations are reached by the dynamics. Focusing on
3427: static objects like structure functions of one variable may be
3428: insufficient for the understanding of the physics of anomalous
3429: scaling. Important confirmation of this picture have been found recently
3430: also for the case of passive scalars advected by a $2d$ turbulent flow
3431: in the inverse cascade regime \cite{cel01a} and for the case of
3432: shell models for passive scalars advection \cite{ara01a}.
3433: \subsubsection{Summary and Discussion}
3434: The main lesson from this subsection is that the scaling exponents
3435: form a discrete and strictly increasing spectrum as a function of
3436: $j$. This is the first example where this can be shown rigorously.
3437: The meaning of this result is that for higher $j$ the anisotropic
3438: contributions to the statistical objects decay faster upon decreasing
3439: scales. The rate of isotropization is determined by the difference
3440: between the $j$ dependent scaling exponents, and is of course a power
3441: law. The result shows that to first order in $\epsilon$ the
3442: $j$-dependent part is independent of the order of the correlation
3443: function. This means that the rate of isotropization of all the
3444: moments of the distribution function of field differences across a
3445: given scale is the same. This is a demonstration of the fact that,
3446: to $O(\epsilon)$ the
3447: distributions function itself tends toward a locally isotropic
3448: distribution function. We note in passing that to
3449: first order in $\epsilon$ the $j$ dependent part is also the same for
3450: $\xi^{(2)}$, a quantity whose isotropic value is {\em not}
3451: anomalous. For all $j>1$ also $\xi^{(2)}_j$ is anomalous, and in
3452: agreement with the $n=2$ value of Eq.~(\ref{eq:perturbative}).
3453: Significantly, for
3454: $\xi_j^{(2)}$ we have a nonperturbative result that was derived in
3455: \cite{fai96}, namely
3456: $$
3457: \xi^{(2)}_j=\frac{1}{2}\Big(2-d-\epsilon+\sqrt{(2-d-\epsilon)^2
3458: +\frac{4(d+\epsilon-1)j(d+j-2)}{d-1}}\Big) \ , \quad
3459: j\ge 2$$
3460: valid for all values of $\epsilon$ in the interval (0,2) and for
3461: all $j \ge 2$. This exact result agrees after expanding to
3462: $O(\epsilon)$ with (\ref{eq:perturbative}) for $n=2$ and $j=2$.
3463:
3464: The second lesson from this first exactly solvable example
3465: was the correspondence between
3466: the scaling exponents of the zero modes in the inertial interval
3467: and the corresponding scaling exponents of the gradient fields.
3468: The latter do not depend on any inertial scales, and the exponent
3469: appears in the combination
3470: $\left(\Lambda/\eta\right)^{\xi^{(n)}_j}$ where $\eta$ is the
3471: appropriate ultraviolet inner cutoff. We found exact agreement with
3472: the exponents of the zero modes in all the sectors of the
3473: symmetry group and for all values of $n$. The deep reason behind
3474: this agreement is the linearity of the fundamental equation of
3475: the passive scalar (\ref{advect}). This translates to the fact
3476: that the viscous cutoff $\eta$ is $n$ and $j$
3477: independent, and also does not depend on the inertial separations
3478: in the unfused correlation functions. This point has been
3479: discussed in detail in \cite{fai96,lvo96a}. In the case of
3480: Navier-Stokes statistics we expect this ``trivial" correspondence
3481: to fail. Nevertheless, many attempts have been done
3482: to describe the matching between the inertial and dissipative
3483: scaling properties \cite{lvo96,fri91,ben96b}
3484: for the isotropic sector. Finally we note that in the present
3485: case we have displayed the fusion rules in all the $j$
3486: sectors, using the $O(\epsilon)$ explicit form of the zero modes.
3487: We expect the fusion rules to have a nonperturbative validity for
3488: any value of $\epsilon$.\\
3489: An interesting modification of Kraichnan models has been recently
3490: proposed in \cite{cel04} where the scaling properties of a passive
3491: scalar advected by a Kraichnan-like shear flow are investigated.
3492: The anisotropy introduced by the shear breaks the foliations of the
3493: correlation functions equations. Nevertheless, the authors have been able
3494: to explain the existence of a scaling range in the passive spectrum
3495: with anomalous slope (i.e. different from the result obtained
3496: in absence of shear), for scales larger than the typical shear-length
3497: in the system. This anomalous slope is due to the fast advection of passive
3498: particles in the mean shear direction.
3499: %%%%%%%%%%%%%%%%%%%
3500: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3501: \subsection{Passively Advected Magnetic Field}
3502: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3503: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3504: \label{chap:magnetic} Another exactly solvable system of some interest
3505: is the case of passively advected magnetic field. This model was first
3506: proposed in \cite{ver96}. It describes the advection of a
3507: magnetic field $\BB(\Bx,t)$ by the same Kraichnan stochastic velocity
3508: field described in Eq.~(\ref{W1}). The equation of motion for the
3509: magnetic field is
3510: \begin{eqnarray}
3511: \partial_t \BB(\Bx,t) &&+ \Big[\Bu(\Bx,t)\cdot\B{\nabla}\Big]\BB(\Bx,t)
3512: - \Big[\BB(\Bx,t)\cdot\B{\nabla}\Big] \Bu(\Bx,t)
3513: \nonumber\\&&= \kappa \nabla^2\BB(\Bx,t) + \B{f}(\Bx,t) \ , \label{eq:B}
3514: \end{eqnarray}
3515: which has to be supplemented by the solenoidality condition
3516: $\B{\nabla}\cdot\BB(\Bx,t)=0$. The source (``forcing'') term $\B{f}(\Bx,t)$
3517: is a solenoidal vector field that is responsible for injecting the magnetic
3518: field into the system at large scales.
3519: The second-order moment of the source field here is a
3520: second-order solenoidal tensor
3521: \begin{equation}
3522: \Big< f^\alpha(\Bx+\Br,t')f^\beta(\Bx,t) \Big> \equiv
3523: \delta(t-t')A^{\alpha\beta}\left(\frac{\Br}{L}\right) \ ,
3524: \label{eq:ff}
3525: \end{equation}
3526: instead of a scalar. The tensor
3527: $A^{\alpha\beta}(\By)$ is used to mimic large-scale anisotropic boundary
3528: conditions and is therefore taken to be anisotropic, analytic in $\By$ and
3529: vanishing rapidly for $y \gg 1$.
3530: Finally, the dissipative term
3531: $\kappa \nabla^2\BB(\Bx,t)$ dissipates the magnetic field out of the system
3532: at small scales.
3533:
3534: Notice that in order to keep the magnetic field solenoidal, \Eq{eq:B}
3535: contains a ``stretching'' term
3536: $\Big[\BB(\Bx,t)\cdot\B{\nabla}\Big]\Bu(\Bx,t)$. This term may cause a
3537: ``dynamo effect'', which is what happens when the magnetic field amplifies
3538: itself by extracting kinetic energy from the velocity field \cite{chi95}.
3539: Such effect
3540: can destabilize the system, and prevents it from reaching a stationary
3541: state.
3542:
3543: Just as in the Kraichnan passive scalar case, we can use the fact that
3544: both the velocity and source fields are white-noise Gaussian
3545: processes, and derive a closed set of equations for the simultaneous
3546: $n$th-order correlation-functions of the magnetic field. For example,
3547: the equation of motion for the second order magnetic correlation
3548: function $$ C^{\alpha\beta}(\Br,t) \equiv \Big< B^\alpha(\Bx+\Br,t)
3549: B^\beta(\Bx,t) \Big> \, $$ can be easily derived
3550: \cite{ver96}:
3551: \begin{eqnarray}
3552: \partial_t C^{\alpha\beta}&=& K^{\mu\nu}\partial_\mu \partial_\nu
3553: C^{\alpha\beta} - [(\partial_\nu K^{\mu\beta})\partial_\mu
3554: C^{\alpha\nu} + (\partial_\nu K^{\alpha\mu}) \partial_\mu
3555: C^{\nu\beta}] \nonumber\\ &+&(\partial_\mu \partial_\nu
3556: K^{\alpha\beta}) C^{\mu\nu} + 2\kappa\nabla^2 C^{\alpha\beta} +
3557: A^{\alpha\beta} \equiv {\hat
3558: T}^{\alpha\beta}_{~~\sigma\rho}C^{\sigma\rho} + A^{\alpha\beta} \
3559: , \label{EqC} \end{eqnarray}
3560: where one has to add also the solenoidal condition for the magnetic field,
3561: $\partial_\alpha C^{\alpha\beta} = 0$
3562: and the tensor $K^{\mu \nu}$
3563: is the two-point velocity correlation (\ref{kappa1}).
3564: The solution of (\ref{EqC}) was found in
3565: \cite{ver96}. It was shown there that for $0<\epsilon <1$ no dynamo
3566: occurs, while for $\epsilon >1$ a dynamo is developed. Consequently
3567: for $0<\epsilon<1$ the system may reach a stationary state where the
3568: correlation function of the magnetic field behaves like a power law in
3569: the inertial range. In \cite{ver96} the zero modes of the
3570: second-order correlation-function was calculated and its anomalous
3571: scaling in the isotropic sector was found for any $0 \le \epsilon \le
3572: 1$. Notice that for this passive vector model, the absence of any
3573: conservation law for the magnetic energy allows for anomalous scaling
3574: already for the second order correlation in the isotropic sector, at
3575: difference from what happens in the passive scalar case discussed in
3576: section (\ref{s:Kra}). This was the first case where a fully
3577: non-perturbative analytical solutions was presented demonstrating the
3578: possibility to have anomalous scaling in hydrodynamic problems.
3579:
3580: In ref.~\cite{lan99,ara00a} this analysis was generalized to all
3581: the sectors of the SO($3$) group using the SO($3$) decomposition. Here we
3582: review the results presented in ref.~\cite{ara00a} where a
3583: systematic non-perturbative study of the solutions of (\ref{EqC}) was
3584: given in all $(j,m)$ sectors of the SO($3$) group. As usual, it is
3585: advantageous to decompose the covariance $C^{\alpha\beta}$ in terms of
3586: basis functions that block-diagonalize the angular part of the
3587: operator $\hat{ \B T}$,which is invariant to all
3588: rotations. In addition, $\hat{ \B T}$ is invariant to the
3589: parity transformation $\B r \to -\B r$, and to the index permutation
3590: $(\alpha,\mu)\Leftrightarrow (\beta,\nu)$. Accordingly, $\hat{ \B T}$
3591: can be further block-diagonalized into blocks with definite parity and
3592: symmetry under permutations.
3593:
3594: In light of these consideration we seek solutions
3595: in terms of the decomposition given in (\ref{Texp}):
3596: \begin{equation}
3597: C^{\alpha\beta}(\Br,t) =\sum_{q,j,m} C^{(2)}_{q,jm}(r,t) \,\,
3598: B^{\alpha\beta}_{q,jm}(\hat{\Br}).
3599: \label{c-expand}\end{equation}
3600: As discussed in sec. (\ref{sec:construction}) the nine basis functions
3601: can be grouped in four sub-groups depending on their symmetries under
3602: parity and index permutation (\ref{eq:second-rank-tensors}).
3603: It should be noted that not all
3604: subsets contribute for every value of $j$.
3605: Space homogeneity implies the obvious symmetry of the covariance:
3606: $
3607: C^{\alpha\beta}(\B r,t)=C^{\beta\alpha}(-\B r,t) $.
3608: Therefore representations symmetric to $\alpha,~\beta$ exchange must
3609: also have even parity, while antisymmetric representations must have
3610: odd parity. Accordingly, even $j$'s are associated with subsets I and
3611: III, and odd $j$'s are associated with subset II. Subset IV cannot
3612: contribute to this theory due to the solenoidal constraint.
3613: %%%%%%%%%%%%%%
3614: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3615: %
3616: % Section III - The matrix representation of the operator $\hat {\bf T}$
3617: %
3618: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3619: %%
3620: \subsubsection{The Matrix Representation of the Operator $\hat {\bf T}$}
3621: Having the angular basis functions we seek the representation of the
3622: operator $\hat{ \B T}$ in this basis. In such a representation $\hat{ \B T}$
3623: is a
3624: differential operator with respect to $r$ only. In Appendix A
3625: of \cite{ara00a} it is shown
3626: how $\hat{ \B T}$ mixes basis functions within a given subset, but
3627: not between the subsets - as is expected in the last section. In finding the
3628: matrix representation of $\hat{ \B T}$ we are aided by the incompressibility
3629: constraint. Consider first subset I made of the four
3630: symmetric and with $(-)^j$ parity
3631: basis functions: $B^{\alpha\beta}_{q,jm}(\hat{\Br})$ with $q=1,5,7,9$
3632: in a given $j,m$ sector. To simplify the notation we
3633: denote the $a$'s coefficients according to
3634: $a(r) \equiv C^{(2)}_{9jm}(r)$, $b(r) \equiv C^{(2)}_{7jm}(r)$,
3635: $c(r) \equiv C^{(2)}_{1jm}(r) $ and $d(r) \equiv C^{(2)}_{5jm}(r)$.
3636: Primes will denote
3637: differentiation with respect to $r$. \\
3638: In this basis the operator $\hat{ \B T}$ takes on the form
3639: \begin{equation}
3640: \hat{ \B T} \left [\left(
3641: \begin{array}{c} a\\ b \\ c \\ d \end{array} \right)\right]
3642: =\B T_1 \left( \begin{array}{c} a''\\ b'' \\ c'' \\ d'' \end{array} \right)
3643: + \B T_2\left( \begin{array}{c} a'\\ b' \\ c' \\ d'\end{array} \right)
3644: + \B T_3 \left(\begin{array}{c} a\\ b \\ c \\ d \end{array} \right) \ .
3645: \label{T123}
3646: \end{equation}
3647: On the RHS we have matrix products. In addition, the solenoidal condition
3648: implies the following two constrains on $a,~b,~c$ and $d$ (cf. the Appendix
3649: of \cite{ara99b}):
3650: \begin{eqnarray}
3651: 0 &=&a' + 2\frac{a}{r} + jb' - j^2\frac{b}{r} + c' - j\frac{c}{r}
3652: \nonumber\\
3653: 0 &=&b' + 3\frac{b}{r} + \frac{c}{r} + (j-1)d' - (j-1)(j-2)\frac{d}{r}
3654: \nonumber \end{eqnarray}
3655: Using these conditions one can bring $\B T_1$ and $\B T_2$ to diagonal
3656: forms,
3657: $$
3658: \B T_1=2(Dr^\epsilon+\kappa) \B 1 \; ;
3659: %\left(
3660: %\begin{array}{cccc}
3661: %1 & & & \\
3662: %& 1 & & \\
3663: %& & 1 & \\
3664: %& & & 1
3665: %\end{array}
3666: %\right)
3667: \qquad
3668: \B T_2=\frac{4}{r}[(Dr^\epsilon+\kappa)+\epsilon D r^\epsilon] \B 1
3669: %\left(
3670: %\begin{array}{cccc} 1
3671: %& & & \\
3672: %& 1 & & \\
3673: %& & 1 & \\
3674: %& & & 1
3675: %\end{array}
3676: %\right) \ .
3677: $$
3678: where $\B 1$ is the unit matrix.
3679: $\B T_3$ can be written in the form
3680: $$
3681: \B T_3 = Dr^{\epsilon-2} \B Q(j,\epsilon)+\kappa r^{-2}\B Q(j,0).
3682: $$
3683: The explicit expression for the four columns of $\B Q(j,\epsilon)$
3684: can be found in \cite{ara00a}
3685: %\begin{eqnarray}
3686: %\left(
3687: %\begin{array}{c}
3688: %-(2+\epsilon )(j+2)(j+3)+2\epsilon \lbrack (j+1)(2+\epsilon )+8]+\epsilon
3689: %^{2}(1-\epsilon )\\ (2+\epsilon
3690: %)(2-\epsilon )\\ (2+\epsilon )(2-\epsilon )(1-\epsilon )\\0\end{array}
3691: %\right)\nonumber\\
3692: %\left(\begin{array}{c} -2j(j+1-\epsilon )\epsilon (2-\epsilon ) \\
3693: %-j(2+\epsilon )(j+1)+2\epsilon
3694: %(7-\epsilon )\\ -2j\epsilon (2+\epsilon )(2-\epsilon )\\ 2(2+\epsilon
3695: %)(2-\epsilon
3696: %)\end{array}\right)\nonumber\\ \left(\begin{array}{c} -\epsilon (2-\epsilon
3697: %)(2j-3-\epsilon )\\ \epsilon (2-\epsilon )\\ -j(2+\epsilon
3698: %)(j+1)+\epsilon^{2}(3+\epsilon )\\
3699: %0\end{array}\right)\nonumber\\ \left(\begin{array}{c} -j(j-1)(2-\epsilon
3700: %)(4-\epsilon
3701: %)\epsilon \\ -\epsilon(j-1)(2-\epsilon)(j-4) \\ -j(j-1)(2-\epsilon
3702: %)(2+\epsilon )\epsilon \\ -(2+\epsilon )(j-2)(j-1+2\epsilon )-2\epsilon
3703: %\end{array}
3704: %\right)
3705: %\end{eqnarray}
3706: In Appendix B of (\cite{ara00a})
3707: the two remaining blocks (subsets II, III after the list
3708: (\ref{eq:second-rank-tensors}) ), in the
3709: matrix representation of $\hat{ \B T}$ as a function of $j$ have been also
3710: investigated. The single basis
3711: $B_{3,jm}$ (subset IV) cannot appear in the theory since $C^{(2)}_{3jm}=0$
3712: by the
3713: solenoidal condition (cf. Appendix of \cite{ara99b}).
3714: Lastly, there are no solutions belonging to the $j=1$ sector. This is due
3715: to the fact that such solutions correspond to subset II. In this subset the
3716: $j=1$ solenoidal condition implies the equation:
3717: $\frac{d}{dr} C^{(2)}_{81m}+\frac{3C^{(2)}_{81m}}{r}=0 $,
3718: or $C^{(2)}_{81m} \propto r^{-3}$ which is not an admissible solution.
3719: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3720: %%
3721: %
3722: % Section IV - Absence of dynamo effect
3723: %
3724: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3725: %%
3726:
3727: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3728: %%
3729: %
3730: % Section V - calculation of the scaling exponents
3731: %
3732: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3733: %%
3734: \subsubsection{Calculation of the Scaling Exponents}
3735: Before turning to the computation of the exponents one should consider
3736: the existence of a stationary solution for
3737: $t\to \infty$. In \cite{ver96}
3738: it was showed that there is not dynamo in the isotropic
3739: sector as long as $\epsilon<1$. In \cite{ara00a}
3740: it has been demonstrated that for the same values of $\epsilon$,
3741: the dynamo effect is absent also in the anisotropic sectors.
3742: The reader is referred to \cite{ara00a} for details
3743: on this subject.
3744: In the absence of a dynamo effect, we can consider a stationary state of the
3745: system, maintained by the forcing term $\B f(\B r,t)$. The covariance in
3746: such
3747: a case will obey the following equation:
3748: $$
3749: {\hat T}^{\alpha\beta}_{~~\sigma\rho}C^{\sigma\rho}+A^{\alpha\beta}=0 \ .
3750: $$
3751: Deep in the inertial range we look for scale invariant solutions of
3752: the above equation
3753: neglecting the dissipative terms.
3754: The most general scale invariant solution can be expressed as a linear
3755: superposition of homogeneous (zero-modes)
3756: and non-homogeneous solutions of the above equation:
3757: $$C^{\sigma\rho}(\B r) = C_{hom}^{\sigma\rho}(\B r) +
3758: C_{non-h}^{\sigma\rho}(\B r).$$
3759: In particular, only zero-modes can carry anomalous scaling, being the
3760: scaling properties of the non-homogeneous solutions fixed by
3761: the dimensional matching $ {\hat T}^{\alpha\beta}_{~~\sigma\rho}
3762: C^{\sigma\rho} \sim A^{\alpha\beta} $. Therefore,
3763: the existence of a leading anomalous scaling contribution to small scales
3764: magnetic fluctuations is connected to the existence of one, some,
3765: zero-modes with scaling exponents smaller than the dimensional estimate.
3766:
3767: The calculation of the scale-invariant solutions becomes rather immediate
3768: once
3769: we know the functional form of the operator $\hat{\B T}$ in the basis of the
3770: angular tensors $\B B_{q,jm}$. Using the expansion (\ref{c-expand}), and the
3771: fact that $\hat{\B T}$ is block diagonalized by such an expansion, we get a
3772: set of 2nd order coupled ODE's for each block. To demonstrate this point,
3773: consider the four dimensional block of $\hat{\B T}$, created by the four
3774: basis
3775: tensors $\B B_{q,jm}$ of subset I. According to the notation of the last
3776: section, we denote the coefficients of these angular tensors in
3777: (\ref{c-expand}), by the four functions $a(r),b(r),c(r),d(r)$:
3778: $$
3779: C^{\alpha\beta}(\B r) \equiv a(r)B^{\alpha\beta}_{9,jm} +
3780: b(r)B^{\alpha\beta}_{7,jm} + c(r)B^{\alpha\beta}_{1,jm} +
3781: d(r)B^{\alpha\beta}_{5,jm} + \dots \ ,
3782: $$
3783: where ($\dots$) stand for terms with other $j,m$ and other symmetries with
3784: the same $j,m$. Let us first consider the case where $\xi>0$. According to
3785: (\ref{T123}), well within the inertial range, these functions obey:
3786: \begin{equation}
3787: \B T_1(\kappa=0)
3788: \left(\begin{array}{c}a''\\ b'' \\ c'' \\ d'' \end{array} \right) +
3789: \B T_2(\kappa=0)\left(\begin{array}{c} a'\\ b'\\ c'\\ d'\end{array}\right) +
3790: \B T_3(\kappa=0) \left( \begin{array}{c} a\\ b\\ c\\ d \end{array}\right)=0
3791: \ .
3792: \label{zero-eq}
3793: \end{equation}
3794: Due to the scale-invariance of these equations, we look for scale-invariant
3795: solutions in the form:
3796: \begin{equation}
3797: a(r)=ar^{\xi}, \quad b(r)=br^{\xi},\quad d(r)=cr^{\xi},
3798: \quad d(r)=dr^{\xi} \ .
3799: \label{si-form}
3800: \end{equation}
3801: Where $a,b,c,d$ are complex constants. Substituting (\ref{si-form}) into
3802: (\ref{zero-eq}) results in a set of four
3803: linear homogeneous equations for the unknowns $a,b,c,d$ :
3804: $$
3805: \left[
3806: \xi(\xi-1) \B T_1(\kappa=0) + \xi \B T_2(\kappa=0) + \B
3807: T_3(\kappa=0)
3808: \right]
3809: \left(\begin{array}{c} a\\ b \\ c \\ d \end{array} \right) =0 \ .
3810: $$
3811: The last equation admits non-trivial solutions only when
3812: $$
3813: \det\left[
3814: \xi(\xi-1) \B T_1(\kappa=0) + \xi \B T_2(\kappa=0) + \B
3815: T_3(\kappa=0)
3816: \right]=0 \ .
3817: $$
3818: This solvability condition allows us to express $\xi$ as a function of $j$
3819: and $\epsilon$. Using MATHEMATICA one finds eight possible values of $\xi$,
3820: out-of-which, only four are in agreement with the solenoidal condition:
3821: \begin{eqnarray}
3822: \xi_{j}^{(2)}(i)&=&-\frac{1}{2}\epsilon -\frac{3}{2} \pm \frac{1}{2}
3823: \sqrt{H(\epsilon ,j) \pm 2 \sqrt{K(\epsilon ,j)}} \ , \quad{i=1,2,3,4}
3824: \ ,\label{exponents-I} \\
3825: K(\epsilon,j) &\equiv& \epsilon^4- 2\epsilon^3 + 2\epsilon^3j +
3826: 2\epsilon^3j^2 - 4\epsilon^2j -
3827: 3\epsilon^2 - 4\epsilon^2 j^2 - 8 \epsilon j^2-8\epsilon j + 4\epsilon +
3828: 16j + 16j^2 + 4
3829: \nonumber \\
3830: H(\epsilon ,j) &\equiv& -\epsilon^2 - 8\epsilon + 2\epsilon j^2 + 2\epsilon
3831: j + 4j^2 + 4j + 5
3832: \nonumber \ .
3833: \end{eqnarray}
3834: Not all of these solutions are physically acceptable because not all of them
3835: can be matched to the zero mode solutions in the dissipative regime. To see
3836: why this is so, consider the zero-mode equation for $\epsilon=0$:
3837: \begin{equation}
3838: (2\kappa+2D) \nabla^2 \B C = 0 \ .
3839: \label{zero-eq-zero-epsilon}
3840: \end{equation}
3841: The main difference between the $\epsilon=0$ case and the $\epsilon>0$ case
3842: is that
3843: in the
3844: former the same scale-invariant equation holds {\em both} for the inertial
3845: range and the dissipative range. As a result, for $\epsilon=0$, the zero
3846: modes
3847: scale
3848: with the same exponents in the two regimes. These exponents are given
3849: simply by
3850: (\ref{exponents-I}) with $\epsilon=0$, because for $\epsilon=0$ the zero
3851: modes
3852: equation
3853: with $\kappa=0$ is the same as (\ref{zero-eq-zero-epsilon}) up to the
3854: overall
3855: factor
3856: $\frac{D}{D+\kappa}$ which does not change the exponent. For
3857: $\epsilon=0$ th solutions should be valid for the dissipative regime as
3858: well as
3859: for
3860: the inertial regime, ruling out the two solutions with negative exponents in
3861: (\ref{exponents-I}), for they will give a non-physical divergence as
3862: $r \rightarrow 0$. Assuming now that the solutions (including the
3863: exponents) are
3864: continuous in $\epsilon$, (and not necessarily analytic!), one finds that
3865: also
3866: for
3867: finite $\epsilon$ only the positive exponents appear in the inertial range
3868: (an
3869: exception to that is the $j=0$, to be discussed below). Finally there
3870: are two branches of solutions corresponding to the ($-$) and ($+$) in the
3871: square
3872: root.
3873: $$
3874: \xi_{j}^{(2)}=-\frac{3}{2}-\frac{1}{2}\epsilon + \frac{1}{2}
3875: \sqrt{H(\epsilon ,j)
3876: \pm 2 \sqrt{K(\epsilon ,j)}}\ , \quad \mbox {subset I}.
3877: $$
3878: Note that for $j=0$,
3879: only the branch with the $+$ sign under the square root exists
3880: since the other exponent is not admissible,
3881: being negative for $\epsilon\to 0$, and therefore excluded by continuity.
3882: $\xi_{0}^{(2)}$ however becomes negative as $\epsilon$ increases.
3883: For $j\ge 2$ both solutions are admissible, and the
3884: leading is that one with the minus sign in the square root.\\
3885: Let us also discuss
3886: the behavior of the zero modes in the dissipative regime for
3887: $\epsilon>0$. Here the dissipation terms become dominant and we can neglect
3888: all
3889: other
3890: terms in $\hat{\B T}$. The zero mode equation in this regime becomes
3891: $2\kappa
3892: \nabla^2 C^{\alpha\beta} = 0$, which is again, up to an overall factor,
3893: identical to the zero mode equation with $\kappa=0, \epsilon=0$. The
3894: solutions in
3895: this region are once again scale invariant with scaling exponents
3896: $\xi_{j}^{(2)}|_{\epsilon=0}=j,j-2$. As expected, the correlation
3897: function
3898: $C^{\alpha\beta}(\B r)$ becomes smooth in the dissipative regime.
3899:
3900: In \cite{ara00a} the computation of the exponents corresponding to
3901: subsets II and III is also presented. The result is:
3902: $$
3903: \xi_{j}^{(2)}=-\frac{3}{2}-\frac{1}{2}\epsilon +\frac{1}{2}
3904: \sqrt{1-10\epsilon+ \epsilon ^{2}+2j^{2}\epsilon +2j\epsilon
3905: +4j+4j^{2}}\ ,
3906: \quad \mbox {subset II}
3907: $$
3908: $$
3909: \xi_{j}^{(2)}=-\frac{3}{2}-\frac{1}{2}\epsilon +\frac{1}{2}\sqrt{ \epsilon
3910: ^{2}
3911: + 2\epsilon + 1 + 4j^{2} + 2j^{2}\epsilon + 4j + 2\epsilon j}\ ,
3912: \quad \mbox {subset III}.
3913: $$
3914: For $j=0$ there is no contribution from this subset, as the exponent is
3915: negative.
3916: After matching the zero modes to the dissipative range, one has to
3917: guarantee matching at the outer scale $L$. The condition to be fulfilled is
3918: that
3919: the sum of the zero-modes with the inhomogeneous solutions (whose exponents
3920: are
3921: 2-$\epsilon$) must give $\B C(\B r)\to 0$ as $|\B r|\to L$. Obviously this
3922: means
3923: that
3924: the forcing must have a projection on any sector $\B B_{q,jm}$ for which
3925: $C^{(2)}_{q,jm}$ is nonzero.
3926: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3927: %%
3928: %
3929: % Section VI - Summary and conclusions
3930: %
3931: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3932: %%
3933: \subsubsection{Summary and Conclusions}
3934: The results of this section should be examined in the light of the
3935: previous section on passive scalars. That passive scalar case afforded
3936: only perturbative calculation of anomalous exponents in all
3937: anisotropic sectors. The present example offers exact,
3938: non-perturbative calculations, of the whole spectrum of scaling
3939: exponents that determines the covariance of a vector field in the
3940: presence of anisotropy. The main conclusions are: (i) scaling
3941: exponents of the second order magnetic correlation functions are
3942: anomalous; (ii) they are strictly increasing with the index of $j$ of
3943: the sector, meaning that there is a tendency toward isotropization
3944: upon decreasing the scales of observation.
3945: The equations for the magnetic covariance foliate into
3946: independent closed equations for each set of irreducible
3947: representations of the SO(3) group. Moreover, scaling properties of
3948: the zero-modes do not show any dependence on the $q$ index labeling
3949: projections on different irreducible representations of the SO(3)
3950: groups for each fixed $(j,m)$. The consequence of the latter
3951: property is that transversal and longitudinal correlation have
3952: the same scaling exponents within each anisotropic sector.
3953:
3954: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3955: %%
3956: %
3957: % Chapter 8 - Analytical analysis: the linear pressure model
3958: %
3959: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3960: %%
3961: \subsection{The Linear Pressure Model}
3962: \label{chap:linearP}
3963: In this subsection we discuss the scaling exponents
3964: characterizing the power-law behavior of
3965: the anisotropic components of correlation functions in turbulent systems
3966: with pressure, exploring the fundamental question whether
3967: also for such systems the scaling exponents increase as $j$
3968: increases, or they are bounded from above.
3969: The equations of motion in systems with pressure contain nonlocal
3970: integrals over all space. One could argue that the requirement of
3971: convergence of these integrals bounds the exponents from above. It is
3972: shown here on the basis of a solvable model (the ``Linear Pressure
3973: Model"), that this is not necessarily the case. The model described here
3974: is of a passive vector advection by a rapidly varying velocity
3975: field \cite{ara01}. The advected vector field is divergent free and the
3976: equation contains a pressure term that maintains this condition.
3977: The zero modes of the second order correlation function are found
3978: in all the sectors of the symmetry group. We show that the
3979: spectrum of scaling exponents can increase with $j$ without
3980: bounds, while preserving finite integrals. The conclusion is that
3981: contributions from higher and higher anisotropic sectors can disappear
3982: faster and faster upon decreasing the scales also in systems with
3983: pressure.
3984: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3985: To demonstrate that, consider a
3986: typical integral term of the form,
3987: \begin{equation}
3988: \label{eq:ex}
3989: \int\!\! d\B y\, G(\B r-\B y) C(\B y) \ .
3990: \end{equation}
3991: Here $G(\B r) = -1/(4\pi r)$ is the infinite domain Green
3992: function of the Laplacian operator, and $C(\B r)$ is some
3993: statistical object which is expected to be scale invariant in the
3994: inertial range. If $C(\B r)$ has an infrared cross over at scale
3995: $L$ (or equivalently, the integral has an infrared cutoff at
3996: scale $L$), then the above expression will not be a pure power
3997: law of $r$, not even inside the inertial range. Then how is it
3998: possible that such an expression will cancel out a local term of
3999: $C(\B r)$, as is required by the typical equations of motion?
4000: This puzzle has led in the past to the introduction of the
4001: concept of ``window of locality'' \cite{lvo95c,fuk00}. The window
4002: of locality is the range for the scaling exponents in which no
4003: divergence occurs, even if the cross over length $L$ is taken to
4004: infinity. For these exponents integrals of type (\ref{eq:ex}) are
4005: dominated by the range of integration $y\approx r$ and are
4006: therefore termed ``local". In a ``local" theory no infrared
4007: cutoff is called for.
4008:
4009: In this subsection we present solutions for the scaling exponents
4010: in the anisotropic sectors of a linear model of turbulence with
4011: pressure.
4012: This model reveals two
4013: mechanisms that allow an unbounded spectrum of scaling exponents.
4014: First, a careful analysis of the window of locality in the
4015: anisotropic sectors shows that it widens as $j$ increases. We
4016: always have a leading scaling exponent within the window of
4017: locality. Secondly, there is a more subtle mechanism that comes
4018: to play when sub-leading exponents exist outside the window of
4019: locality. In these cases we show that there exist counter-terms
4020: in the exact solution (not the zero modes!) which maintain the
4021: locality of the integrals. The bottom line is that in these
4022: models the anisotropic exponents are unbounded from above leading
4023: to a fast decay of the anisotropic contributions in the inertial
4024: range.
4025: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4026: The Linear Pressure model captures some of the
4027: aspects of the pressure term in Navier-Stokes turbulence, while being
4028: a linear and therefore much simpler problem. The non linearity of the
4029: Navier-Stokes equation is replaced by an advecting Kraichnan field
4030: $\B u(\B x,t)$ and an advected field $\B v(\B x,t)$. The advecting
4031: field $\B u(\B x, t)$ is taken, as before, the Kraichnan field
4032: (\ref{W1}). Both fields are assumed incompressible. The equation
4033: of motion for the vector field $v^\alpha(\B x, t)$ is:
4034: \begin{eqnarray}
4035: \partial_t v^\alpha + u^\mu \partial_\mu v^\alpha +
4036: \partial^\alpha p - \kappa \partial^2 v^\alpha &=& f^\alpha \ ,
4037: \label{eq:v-p} \\
4038: \partial_\alpha v^\alpha = \partial_\alpha u^\alpha &=& 0 \ .
4039: \end{eqnarray}
4040: In this equation, $\B f(\B x,t)$
4041: is the same as the one in Eq.~(\ref{eq:B}).
4042: Analyticity of $\B f(\B x,t)$ is an
4043: important requirement. It means that $A^{\alpha\beta}(\B x)$ can
4044: be expanded for small $|\B x|$ as a power series in $x^\alpha$;
4045: as a result its leading contribution in the $j$-sector is
4046: proportional to $x^{j-2}$, given by
4047: $\partial^\alpha\partial^\beta x^j Y_{j m}(\hat{\B x})$. To
4048: see that this is the leading contribution the reader can consult
4049: the general discussion of the construction of the irreducible
4050: representations in Ref.\cite{ara99b}. All other analytic
4051: contributions contain less derivatives and are therefore of
4052: higher order in $x$.
4053:
4054: In order to derive the statistical equations of the correlation
4055: function of $v^\alpha(\B x,t)$, we need a version of (\ref{eq:v-p})
4056: without the pressure term. Following the standard treatment of the
4057: pressure term in Navier-Stokes equation, we take the divergence of
4058: (\ref{eq:v-p}) and arrive at,
4059: $$
4060: \partial_\nu \partial_\mu u^\mu v^\nu +
4061: \partial^2 p = 0 \ .
4062: $$
4063: The Laplace equation is now inverted using the Green function of
4064: infinite domain with zero-at-infinity boundary conditions:
4065: $$
4066: p(\B x) = -\int\!\! d\B y\, G (\B x - \B y)
4067: \partial_\nu \partial_\mu u^\mu(\B y) v^\nu(\B y) \ ,$$
4068: with
4069: $ G(\B x) \equiv -1/4\pi x $.
4070: With this expression for $p(\B x)$, Eq.~(\ref{eq:v-p}) can be
4071: rewritten as:
4072: \begin{eqnarray}
4073: \partial_t v^\alpha(\B x, t)
4074: &+& u^\mu(\B x,t) \partial_\mu v^\alpha(\B x,t)
4075: - \partial^\alpha_{(\B x)} \int\!\! d\B y\,
4076: G(\B x-\B y)\partial_\nu\partial_\mu u^\mu(\B y) v^\nu(\B y)
4077: \nonumber\\&-& \kappa \partial^2 v^\alpha(\B x,t) = f^\alpha(\B x,t) \ .
4078: \nonumber
4079: \end{eqnarray}
4080: In \cite{ara01} the
4081: equation of motion for the 2-point
4082: correlation function,$
4083: C^{\alpha\beta}(\B r) \equiv
4084: \left<v^\alpha(\B x+\B r) v^\beta(\B x) \right>$
4085: was found:
4086: \begin{eqnarray}
4087: && \partial_t C^{\alpha\beta}(\B r) - T^{\alpha\beta}(\B r)
4088: -T^{\beta\alpha}(-\B r)
4089: + \int\!\!d\B y\, G(\B r - \B y) \partial^\beta\partial_\nu
4090: T^{\alpha\nu}(\B y)\nonumber\\
4091: &&+ \int\!\!d\B y\, G(-\B r - \B y) \partial^\alpha\partial_\nu
4092: T^{\beta\nu}(\B y) -2\kappa\partial^2 C^{\alpha\beta}(\B
4093: r)\nonumber\\
4094: && = \left<v^\alpha(\B x+\B r) f^\beta(\B x)\right> +
4095: \left<v^\beta(\B x) f^\alpha(\B x+\B r)\right> \ . \nonumber
4096: \end{eqnarray}
4097: where
4098: to simplify the equations we have defined an auxiliary function
4099: $T^{\alpha\beta}(\B r)$:
4100: $$
4101: T^{\alpha\beta}(\B r) \equiv
4102: \partial_\mu^{(r)} \left< v^\alpha(\B x+\B r) u^\mu(\B x) v^\beta(\B x)
4103: \right> \ .
4104: $$
4105: This equation is identical to the equation for the second
4106: order correlation function in the usual Navier-Stokes turbulence,
4107: provided that $u^\mu$ is replaced with $v^\mu$ in
4108: the expression above. Indeed, the vexing problem that we face is
4109: being made very clear: if the triple correlation function has a
4110: power law dependence on $\B r$ with an arbitrarily large
4111: exponent, how can the integral converge in the infrared? One
4112: possibility is that the scaling exponent of $T^{\alpha\beta}(\B
4113: r)$ is sufficiently low, making the integral convergent. The
4114: other possibility is that the correlation function is scale
4115: invariant only in the inertial range and vanishes quickly after
4116: that, which is equivalent to the introduction of an infrared
4117: cutoff. However the integral terms in the equation probe the
4118: correlation function throughout the entire space. Therefore, a
4119: cross over behavior of the correlation function at the outer
4120: scale $L$, seems to contradict a pure scaling behavior of the
4121: correlation function in the inertial range itself. This in turn
4122: implies the saturation of the anisotropic scaling exponents.
4123:
4124: To proceed, we use the fact that the field $\B u(\B x,t)$, as well as
4125: the forcing, are Gaussian white noises with correlation given by
4126: Eq.~(\ref{kappa1}). This enables us to express
4127: $T^{\alpha\beta}(\B r)$ and the correlation of the force in terms of
4128: $C^{\alpha\beta}(\B r)$ and $A^{\alpha\beta}(\B r)$. One can use
4129: the well known method of Gaussian integration
4130: by parts \cite{fri95} which leads to the final equations (see also
4131: appendix of (\cite{ara01}):
4132: \begin{eqnarray}
4133: \partial_t C^{\alpha\beta}(\B r) &=& T^{\alpha\beta}(\B r) +
4134: T^{\beta\alpha}(-\B r) - \int\!\!d\B y\, G(\B r - \B y)
4135: \partial^\beta\partial_\nu T^{\alpha\nu}(\B y) \nonumber \\ &-&
4136: \int\!\!d\B y\, G(-\B r - \B y) \partial^\alpha\partial_\nu
4137: T^{\beta\nu}(\B y) + 2\kappa\partial^2 C^{\alpha\beta}(\B r) +
4138: A^{\alpha\beta}(\B r) \ , \label{eq:dtC} \\ T^{\alpha\beta}(\B r)
4139: &=& -\frac{1}{2}K^{\mu\nu}\partial_\mu\partial_\nu
4140: C^{\alpha\beta}(\B r) + \frac{1}{2}\partial^\alpha_{(\B r)} \int\!\!
4141: d\B y\, G(\B r-\B y) \partial_\tau\Big[ K^{\mu\nu}(\B y)
4142: \partial_\mu \partial_\nu C^{\tau\beta}(\B y) \Big] \nonumber \\ &-&
4143: \frac{1}{2} \int \!\! d\B y\, G(\B y) \partial^\beta\partial_\tau
4144: \Big[ K^{\mu\nu}(\B y)\partial_\mu\partial_\nu C^{\alpha\tau}(\B
4145: r-\B y)\Big] \ . \label{eq:Tab}
4146: \end{eqnarray}
4147: These equations have to be supplemented with two more equations that
4148: follow directly from the definition of $C^{\alpha\beta}(\B r)$:
4149: $$
4150: \partial_\alpha C^{\alpha\beta}(\B r) = 0 \ ,
4151: \quad
4152: C^{\alpha\beta}(\B r) = C^{\beta\alpha}(-\B r) \ .
4153: $$
4154: Finally we note that Eqs.~(\ref{eq:dtC},\ref{eq:Tab}) can be
4155: interpreted in a transparent way, utilizing two projection
4156: operators which maintain the RHS of \Eq{eq:dtC} divergence free
4157: in both indices. To define them, let us consider a tensor field
4158: $X^{\alpha\beta}(\B r)$ which vanishes sufficiently fast at
4159: infinity. Then the two projection operators $\PLO$ and $\PRO$ are
4160: defined by:
4161: \begin{eqnarray}
4162: \PLO X^{\alpha\beta}(\B r) &\equiv& X^{\alpha\beta}(\B r) -
4163: \partial^\alpha_{(r)}\int d\B y\, G(\B r-\B y)
4164: \partial_\mu X^{\mu\beta}(\B y) \ , \nonumber \\
4165: \PRO X^{\alpha\beta}(\B r) &\equiv& X^{\alpha\beta}(\B r) -
4166: \partial^\beta_{(r)}\int d\B y\, G(\B r-\B y)
4167: \partial_\mu X^{\alpha\mu}(\B y) \ .\nonumber
4168: \end{eqnarray}
4169: We observe that $\PLO X^{\alpha\beta}$ and $\PRO X^{\alpha\beta}$
4170: are divergence free in the left and right indices respectively.
4171: Using these operators we can rewrite
4172: Eqs.(\ref{eq:dtC}-\ref{eq:Tab}) in the form
4173: \be
4174: \partial_t C^{\alpha\beta}(\B r) =
4175: \PRO T^{\alpha\beta}(\B r) + \PRO T^{\beta\alpha}(-\B r)
4176: + 2 \kappa\partial^2 C^{\alpha\beta}(\B r)
4177: + A^{\alpha\beta}(\B r) \ , \label{eq:dtC1} \ee
4178: \be
4179: T^{\alpha\beta}(\B r) = -\frac{1}{2} \PLO
4180: K^{\mu\nu}\partial_\mu\partial_\nu
4181: C^{\alpha\beta}(\B r)
4182: - \frac{1}{2} \int \!\! d\B y\, G(\B y) \partial^\beta\partial_\tau
4183: \Big[ K^{\mu\nu}(\B y)\partial_\mu\partial_\nu
4184: C^{\alpha\tau}(\B r-\B y)\Big] \ . \label{eq:Tab1}
4185: \ee
4186: The projection in \Eq{eq:Tab1} guarantees that
4187: $T^{\alpha\beta}(\B r)$ is divergence free in its left index,
4188: while the projection in \Eq{eq:dtC1} guarantees divergence
4189: freedom in the right index.
4190:
4191: Not all the terms in these equations are of the same nature. The
4192: integrals due to the projection operator are easy to deal with by
4193: applying a Laplacian on them. For example,
4194: $
4195: \partial^2 \PRO T^{\alpha\beta}(\B r) = \partial^2 T^{\alpha\beta}(\B r)
4196: - \partial^\beta \partial_\nu T^{\alpha\nu}(\B r) \ $.
4197: On the other hand, there seems to be no way to eliminate the last
4198: integral in \Eq{eq:Tab1}, and therefore we shall refer to it as
4199: the ``non-trivial integral''. Only when the velocity
4200: scaling exponents in (\ref{kappa1}) are $\epsilon=0$ and $\epsilon=2$ it
4201: trivializes: the integral vanishes when $\epsilon=0$ and is
4202: proportional to $C^{\alpha\beta}(\B r)$ when $\epsilon=2$.
4203: Unfortunately, in these extreme cases also the projection
4204: operator trivializes, and the effect of the pressure cannot be
4205: adequately assessed. We prefer to study the problem for a generic
4206: value $\epsilon$ for which the incompressibility constraint and the
4207: pressure terms are non-trivial.
4208:
4209: We deal with the this problem head-on in
4210: Sect.\ref{sec:zeromodes}. Due to the non-trivial integral, we
4211: will not be able to provide a full solution of
4212: $C^{\alpha\beta}(\B r)$, but only of the zero modes. However
4213: before doing so we would like to study a model that affords an
4214: exact solution in order to understand in detail the issues at
4215: hand. In the next section we therefore consider a simplified
4216: model of the Linear Pressure model, yet posing much of the same
4217: riddle.
4218: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4219: \subsubsection{An exactly Solvable Toy model} \label{sec:Toy}
4220:
4221: We construct a toy model which is inspired by equations
4222: (\ref{eq:dtC}, \ref{eq:Tab}) for the correlation function in the
4223: Linear Pressure model. Within this model we demonstrate the
4224: strategy of dealing with the non-local pressure term. Since it is
4225: a simplification of the {\em statistical} equation of the Linear
4226: Pressure model, the toy model has no obvious underlying dynamical
4227: equation.
4228:
4229: In the toy model, we are looking for a ``correlation function''
4230: $C^\alpha(\B r)$, whose equations of motion are:
4231: \begin{eqnarray}
4232: && \partial_t C^{\alpha}(\B r) =
4233: -K^{\mu\nu}(\B r)\partial_\mu\partial_\nu C^\alpha(\B r) \label{eq:C}
4234: -\partial^\alpha_{(r)} \int \!\! d\B x G(\B r-\B x)
4235: \partial_\tau
4236: K^{\mu\nu}(\B x)\partial_\mu\partial_\nu C^\tau(\B x)\nonumber\\&&
4237: +\ \kappa \partial^2 C^\alpha(\B r)+A^\alpha(\B r/L) \ , \nonumber
4238: \\
4239: && \partial_\alpha C^\alpha(\B r) = 0 \label{eq:toy-incomp} \ .
4240: \end{eqnarray}
4241: Here $A^\alpha(\B x)$ is a one-index analog of the correlation
4242: function of the original forces $A^{\alpha\beta}(\B x)$.
4243: Accordingly, we take it anisotropic, analytic in $x^\alpha$ and
4244: rapidly vanishing for $ x \gg 1$. As in the previous model,
4245: also here analyticity requires that the leading contribution for
4246: small $ x$ is proportional to $\partial^\alpha x^j Y_{j
4247: m}(\hat{\B x})$ in the $j$-sector. Accordingly it is of order
4248: $x^{j-1}$.
4249:
4250: The toy model is simpler than the Linear Pressure model in two
4251: aspects: First, the ``correlation function'', $C^\alpha(\B r)$
4252: has one index instead of two and therefore can be represented by
4253: a smaller number of scalar functions. Second, the unpleasant
4254: non-trivial term of the Linear Pressure model is absent. This
4255: will allow us to solve the model exactly for every value of
4256: $\epsilon$. Nevertheless, the toy model confronts us with the same
4257: conceptual problems that exist in the Linear Pressure model and
4258: in NS: can a scale invariant solution in the inertial range with
4259: a cross over to a decaying solution at scale $L$, be consistent
4260: with the integral term? If not, is there a saturation of the
4261: anisotropic exponents?
4262:
4263: \Eq{eq:C} can be rewritten in terms of a new projection operator
4264: $\PO$, which projects a vector $X^\alpha(\B r)$ on its divergence
4265: free part:
4266: $$
4267: \partial_t C^{\alpha} =
4268: -\PO \Big[ K^{\mu\nu}\partial_\mu\partial_\nu C^\alpha \Big]
4269: + \kappa \partial^2 C^\alpha +A^\alpha \ ,
4270: $$
4271: where
4272: $$
4273: \PO X^\alpha(\B r) \equiv X^\alpha(\B r) -
4274: \partial^\alpha
4275: \int\!\! d\B y\, G(\B r-\B y)\partial_\mu X^\mu(\B y)
4276: $$
4277: We shall solve this integro-differential equation by first
4278: turning it into a PDE using the Laplacian operator, and then
4279: turning it into a set of decoupled ODE's using the SO($3$)
4280: decomposition.
4281: As in the Linear Pressure model, the non locality of the
4282: projection operator can be removed by considering a differential
4283: version of the operator:
4284: $$
4285: \partial^2 \PO T^\alpha(\B r) = \partial^2 T^\alpha(\B r) -
4286: \partial^\alpha \partial_\mu T^\mu(\B r) \ .
4287: $$
4288: In stationary condition $\partial_t C^\alpha=0$, and therefore the
4289: differential form of the toy model is given by:
4290: \begin{eqnarray}
4291: && \partial^2 \PO
4292: \Big[K^{\mu\nu}(\B r)\partial_\mu\partial_\nu C^\alpha(\B r)\Big] =
4293: \partial^2 K^{\mu\nu}(\B r)
4294: \partial_\mu\partial_\nu C^\alpha(\B r) -
4295: \partial^\alpha \partial_\tau
4296: K^{\mu\nu}(\B r)\partial_\mu\partial_\nu C^\tau(\B r)\nonumber\\
4297: &&= \kappa \partial^2\partial^2 C^\alpha(\B r) +\partial^2 A^\alpha(\B r)
4298: \ ,
4299: \label{eq:diff-C} \\
4300: &&\partial_\alpha C^\alpha(\B r) = 0 \ \nonumber .
4301: \end{eqnarray}
4302: We have reached a linear PDE of order 4. This PDE will be solved
4303: by exploiting its symmetries, i.e., isotropy and parity
4304: conservation, as demonstrated in the next subsection.
4305:
4306: \Eq{eq:diff-C} and the incompressibility condition of
4307: $C^\alpha(\B r)$ are both isotropic and parity conserving.
4308: Therefore, if we expand $C^\alpha(\B r)$ in terms of spherical
4309: vectors with a definite behavior under rotations and under
4310: reflections, we would get a set of decoupled ODE's for their
4311: coefficients.
4312:
4313: For each sector $(j,m), j>0$ of SO($3$) we have three
4314: spherical vectors:
4315: \begin{eqnarray}
4316: B_{1jm}^\alpha(\hat{\B r}) &\equiv& r^{-j-1}r^\alpha \Phi_{j m}(\B r) \
4317: ,
4318: \nonumber \\
4319: B_{2jm}^\alpha(\hat{\B r}) &\equiv&
4320: r^{-j+1}\partial^\alpha \Phi_{j m}(\B r) \ ,
4321: \nonumber \\
4322: B_{3jm}^\alpha(\hat{\B r}) &\equiv& r^{-j}
4323: \epsilon^{\alpha\mu\nu}r_\mu\partial_\nu \Phi_{j m}(\B r) \ .
4324: \nonumber
4325: \end{eqnarray}
4326: Here $\Phi_{j m}(\B r)=r^j Y_{j m}(\hat{\B r})$, and see
4327: \cite{ara99b} for further details. The first two spherical vectors
4328: have a different parity than the third vector, hence the
4329: equations for their coefficients are decoupled from the equation
4330: for the third coefficient. In the following, we shall consider
4331: the equations for the first two coefficients only, as they have a
4332: richer structure and larger resemblance to the Linear Pressure
4333: model. Finally note that the isotropic sector, i.e., $j=0$, is
4334: identically zero. To see why, notice that in this special sector
4335: there is only one spherical vector, $B_{100}^\alpha(\hat{\B r})
4336: \equiv r^{-1}r^\alpha$. Hence the isotropic part of $C^\alpha(\B
4337: r)$ is given by $c(r)r^{-1}r^\alpha$, $c(r)$ being some scalar
4338: function of $r$. But then the incompressibility condition
4339: (\ref{eq:toy-incomp}) implies that $c(r) \sim r^{-2}$, which has
4340: a UV divergence. We therefore conclude that $c(r)=0$, and
4341: restrict the calculation to $j>0$.
4342:
4343: By expanding $C^\alpha(\B r)$ in terms of the spherical vectors
4344: $\B B_{1jm}, \B B_{2jm}$, we obtain a set of ODEs (decoupled in the
4345: $(j,m)$ labels) for the scalar functions that are the
4346: coefficients of these vectors in the expansion. The equations for
4347: these coefficients can thus be written in terms of matrices and
4348: column vectors. To simplify the calculations, we find the matrix
4349: forms of the Kraichnan operator and of the Laplacian of the
4350: projection operator separately, and only then combine the two
4351: results to one. \vskip 0.3 cm
4352: \subsubsection{The Matrix Form of the Operators and the Solution of the Toy
4353: Model}
4354: In this subsection we derive the matrix form of the Kraichnan operator
4355: and of the Laplacian of the Projection operator in each $j$ sector.
4356: To obtain the matrix of the Kraichnan operator in the basis of
4357: $\B B_{1jm}, \B B_{2jm}$, we expand $C^\alpha(\B r)$:
4358: $$
4359: C^\alpha(\B r) =
4360: c_1(r)B_{1jm}^\alpha(\hat{\B r}) + c_2(r)B_{2jm}^\alpha(\hat{\B r}) \ .
4361: $$ in appendix (\ref{app:lp1}) we show how to find the operator
4362: on $ C^\alpha(\B r) $ in a matrix form
4363: which results in the final equation for $c_1(r)$ and $c_2(r)$:
4364: \begin{eqnarray}
4365: && r^\epsilon \MM_4 \VecII{c^{(4)}_1}{c^{(4)}_2} +
4366: r^{\epsilon-1} \MM_3 \VecII{c^{(3)}_1}{c^{(3)}_2} +
4367: r^{\epsilon-2} \MM_2 \VecII{c^{(2)}_1}{c^{(2)}_2} \nonumber \\
4368: && \quad + r^{\epsilon-3} \MM_1 \VecII{c^{(1)}_1}{c^{(1)}_2} +
4369: r^{\epsilon-4} \MM_0 \VecII{c_1}{c_2} = \VecII{\rho_1}{\rho_2} \ .
4370: \label{eq:C-mat}
4371: \end{eqnarray}
4372: In addition also the incompressibility constraint $\partial_\alpha
4373: C^\alpha(\B r)=0$, can be expressed as a relation between
4374: $c_1(r)$ and $c_2(r)$:
4375: \begin{equation}
4376: \label{eq:incomp}
4377: c'_1 + 2\frac{c_1}{r} + j c'_2 - j(j-1)\frac{c_2}{r} = 0 \ .
4378: \end{equation}
4379: This constraint has to be taken into account when solving
4380: \Eq{eq:C-mat}.
4381: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4382: %\subsubsection{Solving the toy model}
4383: %\label{sec:solution}
4384: The solution of \Eq{eq:C-mat} is somewhat tricky due to the
4385: additional constraint (\ref{eq:incomp}). Seemingly the two
4386: unknowns $c_1(r), c_2(r)$ are over determined by the three
4387: equations (\ref{eq:C-mat}, \ref{eq:incomp}), yet this is not the
4388: case for the two equations (\ref{eq:C-mat}) are not independent.
4389: To see that this is the case and find the solution, it is
4390: advantageous to work in the new basis
4391: $$
4392: d_1 = c_1 + j c_2 \ , \qquad
4393: d_2 = -2c_1 + j(j-1)c_2 \ .$$
4394: In this basis the incompressibility constraint becomes very
4395: simple: $d_2 = rd_1' $,
4396: allowing us to express $d_2$ and its derivatives in terms of
4397: $d_1$. To do that in the framework of the matrix notation, we
4398: define the transformation matrix $\UM$:
4399: $$
4400: \UM \equiv \MatII{1}{j}{-2}{j(j-1)} \ ,
4401: \qquad
4402: \UM^{-1} = \frac{1}{j(j+1)}\MatII{j(j-1)}{-j}{2}{1} \ ,
4403: $$
4404: so that, $ \VecII{d_1}{d_2} = \UM \VecII{c_1}{c_2} \ .$
4405: The equations of $d_i(r)$ are the same as the equations for
4406: $c_i(r)$, with the matrices $\MM_i$ replaced by
4407: $
4408: \NM_i \equiv \UM \MM_i \UM^{-1} \ ,
4409: $
4410: and the sources $\rho_i$ replaced by
4411: $$
4412: \VecII{\rho^*_1}{\rho^*_2} = \UM \VecII{\rho_1}{\rho_2} \ .
4413: $$
4414: Notice that a divergence free forcing $A^\alpha(\B r)$ will cause
4415: $\rho^*_1(r), \rho^*_2(r)$ to be related to each other in the
4416: same way that $d_1(r), d_2(r)$ are related to each other, i.e.,
4417: $
4418: \rho^*_2 = r(\rho^*_1)' \ .
4419: $
4420: Next, we perform the following replacements:
4421: %\begin{eqnarray*}
4422: % d_2 &=& rd^{(1)}_1 \ , \\
4423: % d_2^{(1)} &=& rd^{(2)}_1 + d^{(1)}_1 \ , \\
4424: % d_2^{(2)} &=& rd^{(3)}_1 + 2d^{(2)}_1 \ , \\
4425: % d_2^{(3)} &=& rd^{(4)}_1 + 3d^{(3)}_1 \ , \\
4426: % d_2^{(4)} &=& rd^{(5)}_1 + 4d^{(3)}_1 \ .
4427: %\end{eqnarray*}
4428: $$
4429: d_2 = rd^{(1)}_1, \quad
4430: d_2^{(1)} = rd^{(2)}_1 + d^{(1)}_1, \quad
4431: d_2^{(2)} = rd^{(3)}_1 + 2d^{(2)}_1, $$
4432: $$ d_2^{(3)} = rd^{(4)}_1 + 3d^{(3)}_1, \quad
4433: d_2^{(4)} = rd^{(5)}_1 + 4d^{(3)}_1.
4434: $$
4435: We get an equation written entirely in terms of the function
4436: $d_1(r)$ and its derivatives:
4437: \be
4438: r^{\epsilon}(r V_5 d^{(5)}_1 + V_4 d^{(4)}_1 +
4439: r^{-1}V_3 d^{(3)}_1
4440: + r^{-2}V_2 d^{(2)}_1 + r^{-3}V_1 d^{(1)}_1 +
4441: r^{-4}V_0 d_1) =
4442: \VecII{\rho^*_1}{\rho^*_2}, \label{V5}
4443: \ee
4444: where $V_i$ are two dimensional vectors given by:
4445: \begin{eqnarray}
4446: V_5 \equiv \NM_4 \VecII{0}{1}, \nonumber \qquad
4447: V_4 \equiv \NM_4 \VecII{1}{4} + \NM_3\VecII{0}{1}, \qquad
4448: V_3 \equiv \NM_3 \VecII{1}{3} + \NM_2\VecII{0}{1}, \\
4449: V_2 \equiv \NM_2 \VecII{1}{2} + \NM_1\VecII{0}{1}, \qquad
4450: V_1 \equiv \NM_1 \VecII{1}{1} + \NM_0\VecII{0}{1}, \qquad
4451: V_0 \equiv \NM_0 \VecII{1}{0}. \nonumber
4452: \end{eqnarray}
4453: %\end{multicols}
4454: %\sepBR{8.6}{-0.3}{.5} \\
4455: Their explicit values can be found in \cite{ara01}.
4456: %\begin{eqnarray*}%
4457: % V_5 &=& D\VecII{0}{2} \ , \\
4458: % V_4 &=& D\VecII{2}{16+6\epsilon} \ , \\
4459: % V_3 &=& D\VecII{16+4\epsilon}{-4j^2 - 4j + 32 \epsilon
4460: % + 8 - \epsilon j^2 - \epsilon j + 6\epsilon^2} \ , \\
4461: % V_2 &=& D\VecII{-4j^2 - 4j + 20\epsilon + 24 - \epsilon j^2 -\epsilon j +
4462: %2\epsilon^2}
4463: % {-8\epsilon j^2 - 8\epsilon j - 4\epsilon - 48 +
4464: %22\epsilon^2
4465: % -2\epsilon^2 j^2 - 2\epsilon^2 j + 2\epsilon^3} \ , \\
4466: % V_1 &=& D\VecII{-(\epsilon +2)(-6\epsilon + \epsilon j^2 + \epsilon j +
4467: %4j^2 + 4j)}
4468: % {(\epsilon+2)(6\epsilon^2 - \epsilon^2j^2 - \epsilon^2 j-
4469: %\epsilon j^2 - 18\epsilon
4470: % - \epsilon j + j^4 + 11j^2 + 2j^3 + 10j)} \ , \\
4471: % V_0 &=& D\VecII{j(j-1)(j+2)(j+1)(\epsilon+2)}
4472: % {j(\epsilon+2)(\epsilon-4)(j-1)(j+2)(j+1)} \ .
4473: %\end{eqnarray*}
4474: %\rightline{\sepTL{8.7}{0.4}{0}}
4475: %\begin{multicols}{2}
4476: The Eq.~(\ref{V5}) are for a column vector, and can be regarded as
4477: two scalar differential equations that we refer to as the
4478: ``upper" and the ``lower". The upper ODE is of the fourth order,
4479: while the lower ODE is of fifth order. Non surprisingly, the lower
4480: equation is the first derivative of the upper equation, provided
4481: that $A^\alpha(\B r)$ is divergence free. Hence the two equations
4482: are dependent, and we restrict the attention to the upper
4483: equation. To simplify it, we divide both sides by $Dr^\epsilon$,
4484: replace $d_1(r)$ by $\psi(r)$ and define the RHS to be the
4485: function $S(r)$:
4486: \begin{equation}
4487: S(r)\equiv D^{-1}r^{-\epsilon}\rho^*_1(r) \ .
4488: \label{eq:S(r)}
4489: \end{equation}
4490: After doing so, we reach the following equation:
4491: \begin{equation}
4492: \psi^{(4)} + a_3\frac{\psi^{(3)}}{r} + a_2\frac{\psi^{(2)}}{r^2} +
4493: a_1 \frac{\psi^{(1)}}{r^3} + a_0 \frac{\psi}{r^4} =S(r) \ .
4494: \label{eq:psi}
4495: \end{equation}
4496: Its homogeneous solution is easily found once we substitute,
4497: $
4498: \psi(r) = \psi_0 r^\xi$.
4499: The scaling exponents are the roots of the polynomial,
4500: \begin{eqnarray}
4501: \label{def:Polynomial}
4502: P(\xi) &=& \xi(\xi-1)(\xi-2)(\xi-3)
4503: + a_3\xi(\xi-1)(\xi-2) + a_2\xi(\xi-1) + a_1\xi + a_0 \ .
4504: \nonumber
4505: \end{eqnarray}
4506: The polynomial roots are found to be real and non-degenerate. Two
4507: of them are positive while the other two are negative. They are
4508: given by:
4509: \begin{equation}
4510: \label{def:exponents}
4511: \xi_j(i) = -\frac{1}{2} - \frac{1}{2}\epsilon \pm \frac{1}{2}
4512: \sqrt{A(j,\epsilon) \pm \sqrt{B(j,\epsilon)}} \
4513: \qquad {\mbox i=1,2,3,4} ,
4514: % \xi_2 &=& -\frac{1}{2} - \frac{1}{2}\epsilon + \frac{1}{2}
4515: % \sqrt{A(j,\epsilon) - \sqrt{B(j,\epsilon)}} \ , \nonumber \\
4516: % \xi_3 &=& -\frac{1}{2} - \frac{1}{2}\epsilon - \frac{1}{2}
4517: % \sqrt{A(j,\epsilon) - \sqrt{B(j,\epsilon)}} \ , \nonumber \\
4518: % \xi_4 &=& -\frac{1}{2} - \frac{1}{2}\epsilon - \frac{1}{2}
4519: % \sqrt{A(j,\epsilon) + \sqrt{B(j,\epsilon)}} \ , \nonumber
4520: \end{equation}
4521: where,
4522: $$
4523: A(j,\epsilon) \equiv \epsilon^2 +\epsilon j^2 + \epsilon j - 2\epsilon +
4524: 5 + 4j
4525: + 4j^2 \ ,$$ and $$
4526: B(j,\epsilon) \equiv -8\epsilon^2j - 7\epsilon^2j^2 + 16\epsilon^2 +
4527: 2\epsilon^2j^3
4528: + \epsilon^2j^4 - 8\epsilon j^2 - 8\epsilon j - 32\epsilon + 16 + 64j
4529: +64j^2.
4530: $$
4531: In the limit $\epsilon \to 0$ the roots become, in decreasing order:
4532: $$
4533: \xi_j(1) = j+1, \qquad
4534: \xi_j(2) = j-1, \qquad
4535: \xi_j(3) = -j, \qquad
4536: \xi_j(4) = -j-2,
4537: $$
4538: Fig.~(\ref{fig:toy1})
4539: displays the first few exponents as a function of $\epsilon$.
4540: We note that the spectrum has no sign of saturation as $j$
4541: increases. Before we discuss the meaning of this observation we
4542: will make sure that these solutions are physically relevant and
4543: participate in the full (exact) solution including boundary
4544: conditions.
4545: %%%%%%%%%%%%%%%%%%%
4546: \begin{figure}
4547: %\epsfxsize=8cm
4548: \includegraphics[scale=0.65]{linearP1.eps}
4549: \caption{Scaling exponents of the first few $j$s as a function of
4550: $\epsilon$. Top panels show: set 1 (left); set 2 (right). Bottom panel:
4551: set 3 (left); set 4 (right)}\label{fig:toy1}
4552: \end{figure}
4553: The general solution of \Eq{eq:psi} is traditionally given as the
4554: sum of a special solution of the non-homogeneous equation plus a
4555: linear combination of the zero modes. However when attempting to
4556: match the solution to the boundary conditions it is convenient to
4557: represent it as:
4558: \begin{equation}
4559: \label{eq:psi-general-solution}
4560: \psi(r) = \sum_{i=1}^4
4561: \frac{r^{\xi_j(i)}}{\underbrace{ (\xi_j(i)-\xi_j(1)) \ldots
4562: (\xi_j(i)-\xi_j(4))}_{\scriptsize\mbox{all {\em different} roots}}}
4563: \int_{m_i}^r \!\! dx \, x^{3-\xi_j(i)}\, S(x) \ ,
4564: \end{equation}
4565: where the free parameters of the solution are the four constants
4566: $m_i$. Indeed a change in $m_i$ is equivalent to adding to the
4567: solution a term proportional to $ r^{\xi_j(i)}$. In the next subsection we
4568: find the values of $m_i$ to match the boundary conditions, and
4569: discuss the properties of the solution. \vskip 0.3 cm
4570: \subsubsection{Boundary Conditions and Inertial-range Behavior}
4571: From Eq.(\ref{eq:psi-general-solution}) it is clear that the only
4572: values of $m_i$ that guarantee that the solution remains finite
4573: as $r\to 0$ and that it decays as $r\to\infty$ are
4574: $m_1=m_2=+\infty$, $m_3=m_4 = 0$:
4575: \begin{eqnarray}
4576: \psi(r) =
4577: -\frac{r^{\xi_j(1)}}{(\xi_j(1)-\xi_j(2))(\xi_j(1)-\xi_j(3))(\xi_j(1)-\xi_j(4
4578: ))}
4579: \int_r^\infty \!\! dx \, x^{3-\xi_j(1)}\, S(x) \nonumber \\
4580: -\frac{r^{\xi_j(2)}}{(\xi_j(2)-\xi_j(1))(\xi_j(2)-\xi_j(3))(\xi_j(2)-\xi_j(4
4581: ))}
4582: \int_r^\infty \!\! dx \, x^{3-\xi_j(2) }\, S(x) \nonumber\\
4583: +\frac{r^{\xi_j(3)}}{(\xi_j(3)-\xi_j(1))(\xi_j(3)-\xi_j(2))(\xi_j(3)-\xi_j(4
4584: ))}
4585: \int_0^r \!\! dx \, x^{3-\xi_j(3) }\, S(x) \nonumber \\
4586: +\frac{r^{\xi_j(4)}}{(\xi_j(4)-\xi_j(1))(\xi_j(4)-\xi_j(2))(\xi_j(4)-\xi_j(3
4587: ))}
4588: \int_0^r \!\! dx \, x^{3-\xi_j(4)}\, S(x) \ .
4589: \label{eq:fullsol}
4590: \end{eqnarray}
4591: To understand the asymptotic of this solution we find from
4592: \Eq{eq:S(r)} that for $x \ll L$, $S(x)$ has a leading term which
4593: goes like $x^{j-1-\epsilon}$, whereas for for $x \gg L$, $S(x)$
4594: decays rapidly. It is now straightforward to prove that for $r
4595: \ll L$, the $\xi_j(3), \xi_j(4)$ terms scale like $r^{j+3-\epsilon}$,
4596: the $\xi_j(2)$ term scales like $r^{\xi_j(2)}$ and the $\xi_j(1)$
4597: term scales like $r^{\xi_j(1)}$ for values of $\epsilon$ for which
4598: $\xi_j(1) < j+3-\epsilon$ and like $r^{j+3-\epsilon}$ otherwise. In
4599: addition it is easy to see that for $r \gg L$, $\psi(r)$ exhibits
4600: an algebraic decay: the $\xi_j(1), \xi_j(2)$ terms decay rapidly due
4601: to the decay of $S(x)$ whereas the $\xi_j(3),\xi_j(4)$ terms decay
4602: algebraically like $r^{\xi_i}$ respectively. The asymptotic of
4603: the full solution are thus given by
4604: \begin{equation}
4605: \psi(r) \sim \left\{ \begin{array}{lcr}
4606: r^{\xi_j(2)} &,& r \ll L \\
4607: r^{\xi_j(3)} &,& r \gg L \end{array} \right. \ .
4608: \end{equation}
4609: The obvious conclusion is that there is no saturation in the
4610: anisotropic scaling exponents as $j$ increases. The lack of
4611: contradiction with the existence of an integral over all space
4612: has two aspects. The main one is simple and obvious. The
4613: Integro-differential equation (\ref{eq:C}) for $C^\alpha$ has a
4614: differential version (\ref{eq:diff-C}). Solving the differential
4615: version we are unaffected by any considerations of convergence of
4616: integrals and therefore the solution may contain exponents that
4617: increase with $j$ without limit. Nevertheless the full solution
4618: (\ref{eq:fullsol}) exhibits a cross over at $L$: it increases in
4619: the inertial range $r\ll L$ and decays for $r\gg L$. Thus
4620: plugging it back to the Integro-differential equation we are
4621: guaranteed that no divergence occurs.
4622:
4623: The question why the cross-over length $L$ does not spoil the
4624: scale invariance in the inertial range still remains. The answer
4625: is found in differential form of the equation of motion, given by
4626: \Eq{eq:diff-C}. From this equation we find that the integrand is
4627: a Green's function times a Laplacian of a tensor. By definition
4628: such an integral localizes, i.e. it is fully determined by the
4629: value of the tensor at the external vector $\B r$. In the
4630: language of Eq.(\ref{eq:ex}) $A(\B y)=\nabla^2 B(\B y)$!
4631:
4632: The second and less obvious aspect is that the window of locality
4633: widens up with $j$. This is due to the cancellations in the
4634: angular integration of the anisotropic solutions that are due to
4635: the orthogonality of the $Y_{j
4636: m}(\hat{\B r})$ and their generalizations $B^{\alpha}_{q j
4637: m}(\hat{\B r})$. To demonstrate this consider again the simple integral
4638: (\ref{eq:ex}), and assume that $C(\B y)$ belongs to $(j, m)$
4639: sector, i.e.
4640: $
4641: C(\B y) =a(y) Y_{j m}(\hat{\B y})
4642: $.
4643: For $y \gg r$, we may expand the Green function in $r/y$:
4644: \begin{eqnarray}
4645: G(\B r-\B y) &=& -\frac{1}{4\pi|\B r-\B y|} = -\frac{1}{4\pi
4646: y}\sum_{n=0}^\infty
4647: a_n \left[\left(\frac{r}{y}\right)^2
4648: - 2\frac{\B r \cdot \hat{\B y}}{y}\right]^n \ . \nonumber
4649: \end{eqnarray}
4650: Here $a_n$ are Taylor coefficients. Obviously the dangerous terms
4651: for the infrared convergence are those with low values of $n$.
4652: However all these terms will vanish for $n<j$ due to the
4653: angular integration against $Y_{j m}(\hat{\B y})$. The reason
4654: is that all these terms are of the form $r^{n_1} y^{n_2} (\B
4655: r\cdot \hat{\B y})^{n_3}$ with $n_3< j$. The angular part here
4656: has projections only $Y_{j' m'}$ with $j'\le k_3<j$. The
4657: first term to contribute comes when $n=j$, and is proportional
4658: to the amplitude integral $\int_r^\infty \!\! dy \, y^2 a_{j
4659: m}(y) y^{-j -1}$. For a power law $a_{j m}(y)\sim
4660: y^\lambda$ this implies locality for
4661: $
4662: \lambda< j-2,
4663: $,
4664: instead of $\lambda<-2$ in the isotropic sector. The lower bound
4665: of the window of locality is also extended, and a similar
4666: analysis for $y\ll r$ leads to $\lambda>-j-3$. For the toy
4667: model this translates to the window of locality
4668: $$
4669: -j-\epsilon < \xi_j(i) < j +1 -\epsilon \ .
4670: $$ From the previous analysis we find that the leading power law of
4671: the full solution in the inertial range is $r^{\xi_j(2)}$, which
4672: is inside this ``extended'' window of locality. Nevertheless, the
4673: subleading power $r^{\xi_j(1)}$ originating from the first term in
4674: \Eq{eq:fullsol} is above this window, and its presence in the
4675: solution can be explained only using the first mechanism.
4676:
4677: We will see when we turn back to the Linear Pressure model that
4678: both these mechanisms operate there as well, leading again to a
4679: lack of saturation in the exponents.
4680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4681: \subsubsection{Solving the Linear Pressure Model}
4682: \label{sec:zeromodes}
4683: We now return to the Linear Pressure model.
4684: The methods used to solve it follow very close those
4685: developed for the toy model and therefore will not be described
4686: in the full way.
4687: Contrary to the toy model where we can have the full solution
4688: in the present case we can solve only for the zero modes.
4689: These are scale invariant solutions which solve an
4690: equation containing an integral. Their exponent must therefore
4691: lie within the ``extended'' ($j$ dependent) window of
4692: locality. Finally one can argue that these zero modes are a part
4693: of the full solution that decays for $r \gg L$, and therefore
4694: solve the original equation as well.
4695: %%%%%%%%%%%%%%
4696: We start from Eqs (\ref{eq:dtC1}) and (\ref{eq:Tab1}). In the Appendix
4697: of (\cite{ara01}) \Eq{eq:Tab1}
4698: was brought to the form:
4699: \begin{equation}
4700: \label{eq:simpleT}
4701: T^{\alpha\beta}(\B r) = -\frac{1}{2}\PLO
4702: K^{\mu\nu}\partial_\mu\partial_\nu C^{\alpha\beta}(\B r) - \frac{1}{2}
4703: \frac{12\epsilon
4704: D}{(\epsilon-3)(\epsilon-5)}
4705: \int\!\! d\B y \, G(\B y) y^{\epsilon-2}
4706: \partial^2 C^{\alpha\beta}(\B r-\B y),
4707: \end{equation}
4708: which is true for every $\epsilon\neq 1$. The $\epsilon=1$ case will not be
4709: treated here explicitly. Nevertheless, in \cite{ara01} it was argued that
4710: that the results for $\epsilon=1$ can be deduced from the $\epsilon\neq 1$
4711: results by continuity.
4712:
4713: Looking at \Eq{eq:simpleT}, we note that when $\epsilon=2$, the
4714: integral on the RHS of the above equation trivializes to a local
4715: term $C^{\alpha\beta}(\B r)$. In this limiting case the model can
4716: be fully solved utilizing the same machinery used in the previous
4717: section. The solution can then be used to check the zero modes
4718: computed below for arbitrary values of $\epsilon$.
4719:
4720: To proceed, we substitute \Eq{eq:simpleT} into \Eq{eq:dtC1},
4721: noting that the projector $\PRO$ leaves the non-trivial integral
4722: in (\ref{eq:simpleT}) invariant since it is divergence-free in
4723: both indices. Setting $\partial_t C^{\alpha\beta}(\B r,t) = 0$ in
4724: the stationary case, we arrive to following equation
4725: \begin{eqnarray}
4726: \label{eq:fulleq}
4727: &&0 = -\Big[\PRO\PLO K^{\mu\nu}\partial_\mu\partial_\nu
4728: C^{\alpha\beta}\Big](\B r) \\&& -\frac{12\epsilon
4729: D}{(\epsilon-3)(\epsilon-5)}
4730: \int\!\! d\B y\, G(\B y) y^{\epsilon-2} \partial^2
4731: C^{\alpha\beta}(\B r-\B y) + 2\kappa\partial^2 C^{\alpha\beta}(\B
4732: r)+A^{\alpha\beta}(\B r) \ . \nonumber
4733: \end{eqnarray}
4734: As in the toy model, we apply two Laplacians to the above
4735: equation in order to get rid of the integrals of the projection
4736: operators, and obtain
4737: \begin{eqnarray}
4738: \label{eq:full-diff-eq}
4739: &&0 = -\partial^4\Big[\PRO\PLO K^{\mu\nu}\partial_\mu\partial_\nu
4740: C^{\alpha\beta}\Big](\B r)\\&&-\frac{12\epsilon
4741: D}{(\epsilon-3)(\epsilon-5)}
4742: \int\!\! d\B y\, G(\B y) y^{\epsilon-2} \partial^6
4743: C^{\alpha\beta}(\B r-\B y) + 2\kappa\partial^6 C^{\alpha\beta}(\B
4744: r)+\partial^4
4745: A^{\alpha\beta}(\B r) \ . \nonumber
4746: \end{eqnarray}
4747: Here and in the sequel, the operator $\partial^{2n}$ should be
4748: interpreted as $(\partial^2)^n$. We now seek the homogeneous
4749: stationary solutions of $C^{\alpha\beta}(\B r)$ in the inertial
4750: range (zero modes). These satisfy the equations obtained by
4751: neglecting the dissipation, and setting the forcing and time
4752: derivative to zero:
4753: \begin{eqnarray}
4754: \label{eq:zeroCab} && 0 = \partial^4
4755: K^{\mu\nu}\partial_\mu\partial_\nu C^{\alpha\beta}(\B r)
4756: + \partial^\alpha\partial^\beta\partial_\tau\partial_\sigma
4757: K^{\mu\nu}\partial_\mu\partial_\nu C^{\tau\sigma}(\B r)
4758: \\&&-\
4759: \partial^\alpha\partial_\tau\partial^2
4760: K^{\mu\nu}\partial_\mu\partial_\nu C^{\tau\beta}(\B r)
4761: - \partial^\beta\partial_\tau\partial^2
4762: K^{\mu\nu}\partial_\mu\partial_\nu C^{\alpha\tau}(\B r) \nonumber \\
4763: &&\ +\ \frac{12\epsilon D}{(\epsilon-3)(\epsilon-5)}
4764: \int\!\! d\B y \, G(\B y) y^{\epsilon-2}
4765: \partial^6 C^{\alpha\beta}(\B r-\B y) \ , \nonumber
4766: \end{eqnarray}
4767: Let us now define the RHS of the above equation as the ``zero
4768: modes operator'' $\OO(\epsilon)$, and write the zero modes equation
4769: compactly as
4770: $$
4771: 0 = \Big[ \OO(\epsilon) C^{\alpha\beta} \Big] (\B r) \ .
4772: $$
4773: The solutions of this problem is obtained as before
4774: by expanding $C^{\alpha\beta}(\B r)$ in a
4775: basis that diagonalizes $\OO(\epsilon)$. Full detail of this
4776: procedure are available in \cite{ara01}.
4777: We turn now to discuss the results.
4778: In Fig. (\ref{fig:toy2})
4779: \begin{figure}
4780: %\epsfxsize=8cm
4781: \epsfbox{linearP2.eps}
4782: \caption{Leading scaling exponents for the first
4783: few $j$s. The dashed line indicates the upper bound of the window of
4784: locality}
4785: \label{fig:toy2}
4786: \end{figure}
4787: we show the leading scaling exponents of the Linear Pressure
4788: model for $j=0,2,4,6,8,10$. The results are shown for
4789: From Fig. (\ref{fig:toy2}),
4790: we see that in the isotropic sector and in the
4791: $j=2$ sector, the leading exponent is $\xi_j^{(2)} = 0$,
4792: corresponding to the trivial $C^{\alpha\beta}(\B r) = const$
4793: solution. These zero modes will not contribute to the second
4794: order structure function, which is given by
4795: $$
4796: S^{\alpha\beta}(\B r) =
4797: 2\Big[C^{\alpha\beta}(\B r) - C^{\alpha\beta}(\B 0)\Big] \ ,
4798: $$
4799: and so we have to consider the zero mode with the consecutive
4800: exponent. In the isotropic sector this exponent is exactly
4801: $\xi_0^{(2)}=2-\epsilon$, as can be proven by passing to Fourier space. This
4802: special solution is a finger-print of the existence of a constant
4803: energy flux in this model.
4804: Returning to the main question of this subsection, we see that no
4805: saturation of the anisotropic exponents occurs since the leading
4806: exponent in every $j>2$ sector is $\xi_j^{(2)} \simeq
4807: j-2$. These exponents are within the window of locality of
4808: \Eq{eq:fulleq} which is given by $-j-3 < \xi_j^{(2)} < j - \epsilon$.
4809: However the next-to-leading exponents (which are the leading ones
4810: in the structure function for $j=0,2$), are already out of
4811: this window, and their relevance has to be discussed. in \cite{ara01} it was
4812: proposed
4813: that the same mechanism that works in the toy model
4814: also operates here, and that all these higher
4815: exponents can be found in the full solution. To understand this,
4816: let us write a model equation for the correlation function in the
4817: spirit of Eq.(\ref{eq:ex}):
4818: \begin{equation}
4819: \hat{\mathcal D}C(\B r) + \int d\B y\, K(\B r-\B y) C(\B y) = F(r) \ ,
4820: \label{eq:model}
4821: \end{equation}
4822: with $K$ being some kernel, and $\hat{\mathcal D}$ being some local
4823: differential operator. In view of \Eq{eq:fulleq}, the
4824: differential operator $\hat{\mathcal D}$ should be regarded as the
4825: Kraichnan operator, and the integral term should be taken for all
4826: integral terms in the equation, including integrals due to the
4827: projection operators. These integrals create a window of locality
4828: that we denote by $\lambda_{\mbox{\small low}} < \lambda <
4829: \lambda_{\mbox{\small hi}}$. Any pure scaling solution $C(\B r)
4830: \sim r^\lambda$ with $\lambda$ outside the window of locality
4831: will diverge and hence will not solve the homogeneous part of
4832: \Eq{eq:model}. Nevertheless, we will now demonstrate how this
4833: zero mode can be a part of a full solution without breaking scale
4834: invariance. For this we act with a Laplacian on both sides of
4835: \Eq{eq:model}, in order to get rid of the projection operators
4836: integrals. Of course, like in the Linear Pressure model, this
4837: will not eliminate all integral terms, and thus we can write the
4838: resultant equation as
4839: \begin{equation}
4840: \partial^2\hat{\mathcal D}C(\B r)
4841: + \int d\B y \, K(\B r-\B y)\partial^2 C(\B y) = \partial^2 F(r) \ .
4842: \label{eq:lap-model}
4843: \end{equation}
4844: The main assumption, which was proven analytically in the simple
4845: case of the toy model, is that the above equation has a solution
4846: which is finite for all $r$, and decays for $r \gg L$. Let us now
4847: consider the zero modes of \Eq{eq:lap-model}; their exponents
4848: have to be within the ``shifted'' window of locality
4849: $\lambda_{\mbox{\small low}}+2 < \lambda < \lambda_{\mbox{\small
4850: hi}}+2$. Suppose now that $r^\lambda$ with $\lambda_{\mbox{\small
4851: hi}} < \lambda < \lambda_{\mbox{\small hi}}+2$ is such a
4852: solution, which is therefore part of the full solution of
4853: \Eq{eq:lap-model}. We now claim that this solution also solves
4854: the original equation \Eq{eq:model}, hence allowing the existence
4855: of scaling exponents outside of its window of locality. To see
4856: that, we first notice that since the full solution decays for
4857: $r\gg L$, then all integrals in \Eq{eq:model} converge, and are
4858: therefore well defined. All that is left to show is that the
4859: equation is indeed solved by $C(\B r)$. But this is a trivial
4860: consequence of the uniqueness of the solution for Laplace
4861: equation with zero at infinity boundary conditions. Indeed, if we
4862: denote the integral term in \Eq{eq:model} by
4863: $$ I(\B r) = \int d\B y\, K(\B r-\B y) C(\B y) \ ,$$
4864: then from \Eq{eq:lap-model} we have
4865: $$
4866: \partial^2 I(\B r) = \partial^2 [F(r) - \hat{\mathcal D}C(\B r)] \ ,
4867: $$
4868: and since both $I(\B r)$ and $F(r) - \hat{\mathcal D}C(\B r)$ decay as
4869: $r\to\infty$, then they must be equal. Of course no breaking of
4870: scale invariance occurs because the equation is satisfied and
4871: $F(r) - \hat{\mathcal
4872: D}C(\B r)$ is a sum of an inhomogeneous solution and power laws.
4873:
4874: Returning to the Linear Pressure model, we have shown that not
4875: only the first, leading exponents in every sector are legitimate,
4876: but also the next few exponents. These exponents are inside the
4877: shifted window of locality of the ``Laplaced'' equation
4878: (\ref{eq:zeroCab}), which is given by $-j+1 < \lambda <
4879: j+4-\epsilon$.
4880: At this point, we may ask whether this is also the case for the
4881: other exponents, which are outside this shifted window of
4882: locality. In light of the above discussion, it is clear that all
4883: of them may be part of the full solution, for we can always
4884: differentiate \Eq{eq:fulleq} sufficient number of times, thus
4885: shifting the window of locality to include any of these
4886: exponents. However this procedure is unnecessary once we have
4887: written the prefactor $A(\lambda; j, \epsilon)$ as an infinite sum
4888: of poles in $\lambda$. In that case the equation is defined for
4889: all values of $\lambda$ except for a discrete set of poles,
4890: enabling us to look for exponents as high as we wish.
4891: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4892: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4893: \subsubsection{Summary and Conclusions}
4894: \label{sec:summary}
4895: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4896: The main question raised and answered in this subsection is whether the
4897: existence of the pressure terms necessarily leads to a saturation
4898: of the scaling exponents associated with the anisotropic sectors.
4899: Such terms involve integrals over all space, and seem to rule out
4900: the existence of an unbounded spectrum. We have discussed a
4901: mechanism that allows an unbounded spectrum without spoiling the
4902: convergence of the pressure integrals. The mechanism is
4903: demonstrated fully in the context of the simple toy model, and it is
4904: proposed that it also operates in the case of the Linear Pressure
4905: model. The mechanism is based on two fundamental observations.
4906: The first one is that the window of locality widens up linearly
4907: in $j$ due to the angular integration. The second, and more
4908: important, is that a scaling solution with an unbounded spectrum
4909: can exist {\em as a part of a full solution,
4910: which decays at infinity}. Indeed pure scaling
4911: solutions cannot solve
4912: themselves the zero modes equation if their scaling exponent is
4913: out of the window of locality. However the zero modes are always
4914: part of the full solution which decays to zero once $r \gg L$,
4915: and we have shown that if such a solution solves a differential
4916: version of the full equation, it must also solve the original
4917: equation. Therefore by differentiating the full equation
4918: sufficiently many times, we can always reach a differential
4919: equation with a window of locality as high as we wish. In that
4920: equation we can find zero mode solutions with arbitrarily high
4921: exponents (notice that in the toy model, it was sufficient to
4922: differentiate once to get rid of all integrals, thus obtaining an
4923: ``infinitely wide'' window of locality). But since these zero
4924: modes are part of a full solution that decays at infinity, then
4925: this solution is also valid for the original equation, hence
4926: showing that in the full solution there can be power laws with
4927: arbitrarily high exponents.
4928: Finally we want to comment about the relevance of these calculations
4929: to Navier-Stokes turbulence. If we substitute blindly $\epsilon=4/3$ in our
4930: results, we predict the exponents 2/3, 1.25226, 2.01922, 4.04843,
4931: 6.06860 and 8.08337 for $j=0,2,4,6,8$ and $10$ respectively. It would
4932: be tempting to propose that similar numbers may be expected for
4933: Navier-Stokes flows with weak anisotropy, and indeed for $j=0$ and $2$
4934: this is not too far from the truth. We return to this issue after
4935: analyzing the Navier-Stokes case in next section. The closeness of
4936: the Linear Pressure Model with Navier-Stokes equations has also
4937: been used in \cite{ben01a} to propose a closure for the non-linear
4938: turbulent problem.
4939: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4940: \subsection{A Closure Calculation of Anisotropic Exponents for
4941: Navier-Stokes Turbulence}
4942: In this subsection we start from the Navier-Stokes equations, and
4943: write down an approximate equation satisfied by the second order
4944: correlation function, in a closure approximation (renormalized
4945: perturbation theory in 1-loop order) \cite{yos01,lvo03}.
4946: This equation is nonlinear.
4947: For a weakly anisotropic system we follow \cite{lvo03} in linearizing the
4948: equation, to
4949: define a linear operator over the space of the anisotropic
4950: components of the second order correlation function. The solution
4951: is then a combination of forced solutions and ``zero modes" which
4952: are eigenfunctions of eigenvalue zero of the linear operator.
4953: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4954: \subsubsection{Model Equations for Weak Anisotropy in the Closure
4955: Approximation}
4956: \label{modeleqs}
4957: It is customary to discuss the closure equations in $\B k,t$
4958: representation. The Fourier transform of the velocity field $\B
4959: u(\B r,t)$ is defined by
4960: $$
4961: {\B u}({\B k},t)\equiv \int d{\B r}\, \exp[-i({\B r}\cdot{\B
4962: k} )] {\B u}({\B x},t).
4963: $$
4964: The Navier-Stokes equations for an incompressible fluid then read
4965: $$
4966: \Big[ {\partial \over \partial t}+\nu k^2\Big]u^\alpha (\B k,t) =
4967: \frac{i} {2}\Gamma ^{\alpha \beta \gamma }(\B k)\int {d^3 q d^3 p\over
4968: (2\pi)^3}
4969: \delta (\B k+\B q+\B p){u^*}^\beta (\B q,t)
4970: {u^*}^\gamma (\B p,t).$$
4971: The interaction amplitude $\Gamma ^{\alpha \beta \gamma }(\B k)$
4972: is defined by
4973: $ \Gamma ^{\alpha \beta \gamma }(\B k) =-\left[P^{\alpha\gamma}(\B
4974: k) k^\beta +P^{\alpha\beta}(\B k)k^\gamma \right] \ ,
4975: $ with the transverse projection operator $P^{\alpha\beta}$ defined
4976: as
4977: $ P^{\alpha\beta} \equiv \delta^{\alpha\beta} -\frac{k^\alpha
4978: k^\beta}{k^2}.$
4979: The statistical object that is the concern of this subsection is the
4980: second order (tensor) correlation function $\B F(\B k,t)$,
4981: $$
4982: (2\pi)^3 \FT^{\alpha\beta}(\B k,t)\delta (\B k-\B q) \equiv \langle
4983: u^\alpha (\B k,t){u^*}^\beta(\B q,t)\rangle \ .$$ In stationary conditions
4984: this object is time independent. Our aim
4985: is to find its $k$-dependence, especially in the anisotropic
4986: sectors.
4987:
4988: It is well known that there is no close-form theory for the
4989: second order simultaneous correlation function. We therefore need
4990: to resort to standard closure approximations that lead to model
4991: equations. Such a closure leads to
4992: approximate equations of motion of the form
4993: \begin{equation}
4994: \frac{\partial \FT^{\alpha\beta}(\B k,t)}{2\partial t} =
4995: I^{\alpha\beta} (\B k,t)-\nu k^2 \FT^{\alpha\beta}(\B k,t) \ ,
4996: \end{equation}
4997: where
4998: \begin{equation}
4999: I^{\alpha\beta}(\B k) = \int\frac {d^3q d^3p}{(2\pi)^3} \delta(\B
5000: k+\B p+\B q) \Phi^{\alpha\beta} (\B k,\B q,\B p) \ .
5001: \label{Integral}
5002: \end{equation}
5003: In this equation $
5004: \Phi^{\alpha\beta} (\B k,\B q,\B p)
5005: =\frac{1}{2}[\Psi^{\alpha\beta} (\B k,\B q,\B p)
5006: +\Psi^{\beta\alpha} (\B k,\B q,\B p) ]$,
5007: and
5008: \begin{eqnarray}
5009: &&\Psi^{\alpha\beta} (\B k,\B q,\B p) = \Theta(\B k,\B q,\B p)
5010: \Gamma^{\alpha\gamma\delta}(\B k)[\Gamma^{\delta\beta'\gamma'}(\B q)
5011: \FT^{\gamma\gamma'}(\B p) ^{\beta'\beta}(\B k) +\nonumber\\&&
5012: \Gamma^{\gamma\beta'\delta'}(\B p)
5013: \FT^{\delta\delta'}(\B q) A^{\beta'\beta}(\B k)
5014: + \Gamma^{\beta\delta'\gamma'}(\B k)
5015: \FT^{\delta\delta'}(\B q) A^{\gamma\gamma'}(\B p)] \ . \label{Phi}
5016: \end{eqnarray}
5017: In stationary conditions and for $k$ in the inertial range we need
5018: to solve the integral equation $I^{\alpha\beta}(\B k) = 0$.
5019:
5020: The process leading to these equations is long; one starts with
5021: the Dyson-Wyld perturbation theory, and truncates (without
5022: justification) at the first loop order. In addition one asserts
5023: that the time dependence of the response function and the
5024: correlation functions are the same. Finally one assumes that the
5025: time correlation functions decay in time in a prescribed manner.
5026: This is the origin of the ``triad interaction time" $\Theta(\Bk,\B q,\B p)$.
5027: If one assumes that all the correlation functions
5028: involved decay exponentially (i.e. like $\exp(-\gamma_{\B k}|t|)$,
5029: then
5030: \begin{equation}
5031: \Theta(\B k,\B q,\B p) =\frac{1}{\gamma_{\B k}+\gamma_{\B
5032: q}+\gamma_{\B p}} \ . \label{expdecay}
5033: \end{equation}
5034: For Gaussian decay, i.e. like $\exp[-(\gamma_{\B k} t)^2/2]$,
5035: \begin{equation}
5036: \Theta(\B k,\B q,\B p) =\frac{1}{\sqrt{\gamma^2_{\B
5037: k}+\gamma^2_{\B q}+\gamma^2_{\B p}}} \ . \label{gaussdecay}
5038: \end{equation}
5039: All these approximations are uncontrolled. Nevertheless this type
5040: of closure is known to give roughly correct estimates of scaling
5041: exponents and even of coefficients in the isotropic sector.
5042:
5043: Eq. (\ref{Integral}) poses a nonlinear integral equation which is
5044: closed once $\gamma_{\B k}$ is modeled. One may use the estimate
5045: $\gamma_{\B k}\sim k U_k$ where $U_k$ is the typical velocity
5046: amplitude on the inverse scale of $k$, which is evaluated as
5047: $U^2_k\sim k^3 \FT^{\alpha\alpha}(\B k)$.
5048: \begin{equation}
5049: \gamma_{\B k} = C_\gamma k^{5/2}\sqrt {\FT^{\alpha\alpha}(\B k)} \
5050: . \label{gamma}
5051: \end{equation}
5052: In isotropic turbulence Eqs. (\ref{Integral}) and (\ref{gamma})
5053: have an exact solution with K41 scaling exponents,
5054: \be
5055: \FT^{\alpha\beta}_0(\B k) = P^{\alpha\beta}(\B k) F(k), \quad
5056: F(k) = C\epsilon^{2/3} k^{-11/3}, \quad \gamma_k= \tilde
5057: C_\gamma \epsilon^{1/3} k^{2/3}. \label{j0}
5058: \ee
5059: Note that the scaling exponents in $\B k$-representation, denoted as $\tilde
5060: \zeta$, have a
5061: $d$-dependent difference from their numerical value in $\B r$-
5062: representation. In 3-dimensions $\tilde \zeta^{(2)}=\zeta^{(2)}+3$, and
5063: the exponent 2/3 turns to 11/3 in Eq.(\ref{j0}).
5064: For weak anisotropic turbulence Eq.(\ref{Integral}) will pose a
5065: {\em linear} problem for the anisotropic components which depends
5066: on this isotropic solution.
5067: %%%%%%%%%%%%%%%%%
5068: \subsubsection{Closure with Weak Anisotropy}
5069: In weakly anisotropic turbulence one has to consider a small anisotropic
5070: correction $f^{\alpha\beta} (\B k)$ to the fundamental isotropic
5071: background
5072: $$
5073: \FT^{\alpha\beta}(\B k)=\FT^{\alpha\beta}_0(\B k) + f^{\alpha\beta}
5074: (\B k).$$
5075: The first term vanishes with the solution (\ref{j0}). Linearizing
5076: the integral equation with respect to the anisotropic correction
5077: leads to:
5078: \begin{eqnarray}
5079: &&I^{\alpha\beta}(\B k)\!=\!\! \int\!\!\frac {d^3q d^3p}{(2\pi)^3}
5080: \delta(\B k+\B p+\B q)
5081: [S^{\alpha\beta\gamma\delta} (\B k,\B q,\B p) f^{\gamma\delta}(\B k)
5082: +2T^{\alpha\beta\gamma\delta} (\B k,\B q,\B p)
5083: f^{\gamma\delta}(\B q)]=0 \ , \nonumber\\
5084: &&S^{\alpha\beta\gamma\delta} (\B k,\B q,\B p)\equiv
5085: \frac{\delta\Phi^{\alpha\beta} (\B k,\B q,\B p)}{\delta
5086: \FT^{\gamma\delta}(\B k)}\ ,\quad
5087: T^{\alpha\beta\gamma\delta} (\B k,\B q,\B p)\equiv
5088: \frac{\delta\Phi^{\alpha\beta} (\B k,\B q,\B p)}{\delta
5089: \FT^{\gamma\delta}(\B q)} \ . \label{allj}
5090: \end{eqnarray}
5091: We reiterate that the functional derivatives in Eq.(\ref{allj})
5092: are calculated in the isotropic ensemble. In computing these
5093: derivatives we should account also for the implicit dependence
5094: of $\Theta(\B k,\B q,\B p)$ on the correlation function through
5095: Eq. (\ref{gamma}). We can rewrite Eq. (\ref{allj}) in a way that
5096: brings out explicitly the linear integral operator $\hat L$,
5097: \begin{equation}
5098: \hat L |\B f\rangle\equiv \int\!\!\frac {d^3q}{(2\pi)^3} {\mathcal
5099: L}^{\alpha\beta\gamma\delta}(\B k,\B q) f^{\gamma\delta}(\B q) = 0
5100: \ , \label{opereq}
5101: \end{equation}
5102: where the kernel of the operator is
5103: \begin{equation}
5104: {\mathcal L}^{\alpha\beta\gamma\delta}(\B k,\B q)\equiv \delta(\B
5105: k-\B q)\int\frac {d^3p}{(2\pi)^3}S^{\alpha\beta\gamma\delta} (\B
5106: k,\B p,-\B k-\B p) +2T^{\alpha\beta\gamma\delta}
5107: (\B k,\B q,-\B k-\B q) \ . \label{oper}
5108: \end{equation}
5109:
5110: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5111: \subsubsection{Symmetry Properties of the Linear Operator}
5112: The first observation to make is that the linear operator is
5113: invariant under all rotations. Accordingly we can block
5114: diagonalize it by expanding the anisotropic perturbation in the
5115: irreducible representation of the SO(3) symmetry group. These
5116: have principal indices $j$ with an integer $j$ going from 0
5117: to $\infty$. The zeroth component is the isotropic sector.
5118: Correspondingly the integral equation takes the form
5119: \begin{equation}
5120: I^{\alpha\beta}(\B k) = I^{\alpha\beta}_0(\B
5121: k)+\sum_{j=1}^\infty I^{\alpha\beta}_j(\B k) =0 \ .
5122: \label{intj}
5123: \end{equation}
5124: The block diagonalization implies that each $j$-block provides
5125: an independent set of equations (for every value of $\B k$):
5126: $
5127: I^{\alpha\beta}_j(\B k) =0$.
5128: The first term of (\ref{intj}) vanishes with the solution
5129: (\ref{j0}). For all higher values of $j$ we need to solve the
5130: corresponding equation
5131: \begin{equation}
5132: \hat L |\,\B f_j\rangle = 0 \ . \label{operell}
5133: \end{equation}
5134: We can block diagonalize further by exploiting additional
5135: symmetries of the linear operator. In all discussion we
5136: assume that the turbulent flow has zero helicity. Correspondingly
5137: all the correlation functions are invariant under the inversion
5138: of $\B k$.
5139: Consequently there are no odd $j$ components, and we can write
5140: $$
5141: f^{\alpha\beta} (\B k)= \sum_{j=2,4,...}^\infty
5142: f^{\alpha\beta}_j (\B k).$$
5143: We also note that in general $\B u(-\B k)=\B u^*(\B k)$.
5144: Accordingly, the correlation functions are real. From this fact
5145: and the definition it follows that the correlation functions are
5146: symmetric to index permutation,
5147: $ \FT^{\alpha\beta}_0(\B k) =\FT^{\beta\alpha}_0(\B k)$ and
5148: $f^{\alpha\beta}_j (\B k)=f^{\beta\alpha}_j (\B k)$.
5149: As a result the linear operator is invariant to permuting the
5150: first ($\alpha,\beta$) and separately the second
5151: ($\gamma,\delta$) pairs of indices. In addition, the operator is
5152: symmetric to $\B k\to -\B k$ together with $\B q\to -\B q$. This
5153: follows from the inversion symmetry and from the appearance
5154: of products of two interaction amplitudes (which are
5155: antisymmetric under the inversion of all wave-vectors by
5156: themselves).
5157: Finally, the kernel is a homogeneous function of the wavevectors,
5158: meaning that in every block we can expand in terms of basis
5159: functions that have a definite scaling behavior, being
5160: proportional to $k^{-\tilde\zeta}$.
5161: \subsubsection{SO(3) Decomposition} \label{SO3} As a result of the
5162: symmetry properties the operator $\hat L$ is block diagonalized
5163: by tensors that have the following properties:
5164: \begin{itemize}
5165: \item They belong to a definite sector $(j, m)$ of the SO($3$) group.
5166: \item They have a definite scaling behavior.
5167: \item They are either symmetric or antisymmetric under permutations of
5168: indices.
5169: \item They are either even or odd in $\B k$.
5170: \end{itemize}
5171: We have already explicitly presented the tensors involved
5172: for the case of passive vector advection. Here
5173: we only quote the final results translated into $\B k$ space.
5174: In every sector $(j, m)$ of
5175: the rotation group with $j > 1$, one can find 9 independent
5176: tensors $X^{\alpha\beta}(\B k)$ that scale like
5177: $k^{-x}$. They are given by $k^{-x} \tilde
5178: B_{j,j m}^{\alpha\beta}(\hat{\B k})$, where the index $j$ runs
5179: from 1 to 9, enumerating the different spherical tensors. The
5180: unit vector $\hat{\B k}\equiv \B k/k$. These nine tensors can be
5181: further subdivided into four subsets exactly like the real-space
5182: decomposition of
5183: Sect.~\ref{alternative}:
5184: \begin{itemize}
5185: \item \textbf{Subset I} of 4 symmetric tensors with $(-)^j$ parity.
5186: \item \textbf{Subset II} of 2 symmetric tensors with $(-)^{j+1}$ parity.
5187: \item \textbf{Subset III} of 2 antisymmetric tensors with $(-)^{j+1}$
5188: parity.
5189: \item \textbf{Subset IV} of 1 antisymmetric tensor with $(-)^j$ parity.
5190: \end{itemize}
5191: Due to the diagonalization of $\hat L$ by these subsets, the
5192: equation for the zero modes foliates, and we can compute the zero
5193: modes in each subset separately. In this subsection, we choose to
5194: focus on subset I, which has the richest structure. The four
5195: tensors in this subset are given here by
5196: \begin{eqnarray}
5197: \tilde B_{1,j m}^{\alpha\beta}(\hat{\B k}) &=&
5198: k^{-j-2}k^\alpha k^\beta \phi_{j m}(\B k) \nonumber , \\
5199: \tilde B_{2,j m}^{\alpha\beta}(\hat{\B k}) &=&
5200: k^{-j}[k^\alpha \partial^\beta + k^\beta\partial^\alpha]
5201: \phi_{j m}(\B k) \ , \nonumber \\
5202: \tilde B_{3,j m}^{\alpha\beta}(\hat{\B k}) &=&
5203: k^{-j}\delta^{\alpha\beta} \phi_{j m}(\B k) \ , \nonumber \\
5204: \tilde B_{4,j m}^{\alpha\beta}(\hat{\B k}) &=&
5205: k^{-j+2}\partial^\alpha \partial^\beta \phi_{j m}(\B k) \ ,
5206: \label{basis}
5207: \end{eqnarray}
5208: where $\phi_{j m}(\B k)$ are the standard spherical harmonics.
5209:
5210: The last property to employ is the incompressibility of the target
5211: function $f^{\alpha\beta} (\B k)$. Examining the basis
5212: (\ref{basis}) we note that we can find two linear combinations
5213: that are transverse to $\B k$ and two linear combinations that are
5214: longitudinal in $\B k$. We need only the former, which have the
5215: form
5216: \begin{eqnarray}
5217: \label{basis1}
5218: B_{1,j m}^{\alpha\beta}(\hat{\B k}) &=&
5219: k^{-j}P^{\alpha\beta}(\B k) \phi_{j m}(\B k) ,
5220: \\
5221: B_{2,j m}^{\alpha\beta}(\hat{\B k}) &=&
5222: k^{-j}[k^2 \partial^\alpha \partial^\beta -(j-1)( k^\beta
5223: \partial^\alpha
5224: +k^\alpha\partial^\beta) +j(j-1)
5225: \delta^{\alpha\beta} ]
5226: \phi_{j m}(\B k) \ .\nonumber
5227: \end{eqnarray}
5228: Using this basis we can now expand the target function as
5229: \begin{equation}
5230: \label{expand} f^{\alpha\beta}_j (\B k) = k^{-\tilde\zeta_j^{(2)}}
5231: \Big[ c_1 B_{1,j m}^{\alpha\beta}(\hat{\B k}) +
5232: c_2 B_{2,j m}^{\alpha\beta}(\hat{\B k})
5233: \Big] \ .
5234: \end{equation}
5235: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
5236: \subsubsection{Calculation of the Scaling Exponents}
5237: \label{calculation}
5238: Substituting Eq.(\ref{expand}) into Eq.(\ref{operell}) we find
5239: \begin{equation}
5240: \hat L q^{-\tilde\zeta_j^{(2)}}|\B B_{1,j m}\rangle c_1+\hat L
5241: q^{-\tilde\zeta_j^{(2)}}|\B B_{2,j m}\rangle c_2 =0 \ . \label{LonB}
5242: \end{equation}
5243: Projecting this equation on the two function of the basis
5244: (\ref{basis1}) we obtain for the matrix $L_{i,l}(j, \tilde\zeta_j^{(2)})
5245: \equiv \langle \B B_{i,j m}|\hat
5246: L q^{-\tilde\zeta_j^{(2)}}|
5247: \B B_{l,j m}\rangle$ the form:
5248: \begin{equation}
5249: \label{mateq}
5250: L_{i,l}(j, \tilde\zeta_j^{(2)}) = \int\!\!\frac {d^3q}{(2\pi)^3}\,d\hat {\B
5251: k}\,B^{\alpha\beta}_{i,j m}(\hat{\B k}) {\mathcal
5252: L}^{\alpha\beta\gamma\delta}(\B k,\B
5253: q)q^{-\tilde\zeta_j^{(2)}}B^{\gamma\delta}_{l,j m}(\hat{\B q}).
5254: \ee
5255: Here we have full integration with respect to $\B q$, but only
5256: angular integration with respect to $\B k$. Thus the matrix
5257: depends on $k$ as a power, but we are not interested in this
5258: dependence since we demand the solvability condition
5259: \begin{equation}
5260: \det L_{i,l}(j, \tilde\zeta_j^{(2)}) = 0 \ . \label{solvability}
5261: \end{equation}
5262: It is important to stress that in spite of the explicit $m$
5263: dependence of the basis functions, the matrix obtained in this way
5264: has no $m$ dependence. In the calculation below we can therefore
5265: put, without loss of generality, $m=0$. This is like having
5266: cylindrical symmetry with a symmetry axis in the direction of the
5267: unit vector $\hat{\B n}$. In this case we can write the matrix
5268: $\B B_{i, j}(\hat{\B k})$ (in the vector space $\alpha,\,
5269: \beta= x,\,y,\,z$) as
5270: \begin{equation}\label{B-as-operator}
5271: B^{\alpha\beta}_{i, j}(\hat{\B k}) = k^{-j} \hat {\C
5272: B}^{\alpha\beta}_{i, j,\B k} (k^{j} P_{j} (\hat{\B k}
5273: \cdot \hat{\B n}))\,,
5274: \end{equation}
5275: where $\hat {\C B}^{\alpha\beta}_{i, j,\B k}$ are matrix
5276: operators, acting on wave vector $\B k$:
5277: \begin{eqnarray}\label{def-B-operator}
5278: \hat {\C B}^{\alpha\beta}_{1, j,\B k} & \equiv &
5279: \delta^{\alpha\beta} - \frac{k^\alpha k^\beta}{k^2}, \\ \nonumber
5280: \hat {\C B}^{\alpha\beta}_{2, j,\B k} & \equiv & \frac{
5281: k^2\, \partial^2}{ \partial k^\alpha \partial k^\beta} - (j -
5282: 1) \Big ( \frac{ k^\alpha \partial}{\partial k^\beta}
5283: +\frac{ k^\beta \partial}{\partial k^\alpha }
5284: - j \, \delta^{\alpha\beta}\Big ) \, ,
5285: \end{eqnarray}
5286: and $P_j(x)$ denote $j$-th order Legendre polynomials.
5287: The technical details of the calculations
5288: were presented in \cite{lvo03}. Here we
5289: present and discuss the results.
5290: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5291: \subsubsection{Results and Concluding Remarks}
5292: \label{conclusions}
5293: The determinants $\det[L_{i,l}(j,\tilde\zeta^{(2)}_j]$ were computed as
5294: functions of the scaling exponents $\tilde \zeta^{(2)}_j$ in every
5295: $j$-sector
5296: separately, and the scaling exponent was determined from the zero
5297: crossing. The procedure is exemplified in Fig. \ref{Fig.1v} for
5298: the isotropic sector $j=0$. We expect for this sector
5299: $\tilde\zeta^{(2)}_0=11/3$, in accordance with $\zeta_0^{(2)}=2/3$. Indeed,
5300: for
5301: both decay models, i.e the exponential decay (\ref{expdecay}),
5302: shown in dark line, and the Gaussian decay (\ref{gaussdecay})
5303: shown in light line, the zero crossing occurs at the same point,
5304: which in the inset can be read as 3.6667.
5305: %%%%%%%%%%%%%%%%%%%
5306: \begin{figure}
5307: \includegraphics[scale=0.45]{j0.a.ps}
5308: \caption{determinant and zero crossing for the sector $j=0$. The scaling
5309: exponent
5310: computed from the zero crossing is $\zeta^{(2)}_0\approx
5311: 0.667$.} \label{Fig.1v}
5312: \end{figure}
5313: %%%%%%%%%%%%%
5314: For the higher $j$-sectors the agreement between the
5315: exponential and Gaussian models is not as perfect, indicating
5316: that the procedure is not exact. In Fig. \ref{Fig.2} we present
5317: the determinant and zero crossings for $j=2$.
5318: %%%%%%%%%%%%%%%%%%%
5319: \begin{figure}
5320: \includegraphics[scale=0.45]{j2.a.ps}
5321: \caption{determinant and zero crossing for the sector $j=2$. The scaling
5322: exponent
5323: computed from the zero crossing is $\zeta^{(2)}_2\approx
5324: 1.36-1.37$.} \label{Fig.2}
5325: \end{figure} From the inset we can read the exponents
5326: $\tilde\zeta_2^{(2)}=4.351$ and
5327: 4.366 for the exponential and Gaussian models respectively. This
5328: is in correspondence with $\zeta_2^{(2)}=1.351$ and $1.366$
5329: respectively. These numbers are in excellent correspondence with
5330: the experimental measurements reported in
5331: \cite{ara98,kur00a}, cf. the next
5332: Section.
5333: The results for $j=4$ are presented in Fig. \ref{Fig.3}. Here
5334: the zero crossing, as seen in the inset, yields very close
5335: results for $\tilde\zeta_4^{(2)}$ between the exponential and Gaussian decay
5336: models, i.e. $\tilde\zeta_4^{(2)}\approx 4.99$. Note that this result is
5337: very close to the boundary of locality as discussed in \cite{lvo03}.
5338: Nevertheless the zero crossing is still easily
5339: resolved by the numerics, with the prediction that
5340: $\zeta_4^{(2)}\approx 1.99$. The simulation estimate of this number
5341: in \cite{bif01a} was $1.7\pm 0.1$. We note that while the result
5342: $\zeta_4^{(2)}\approx 1.99$ is not within the error bars of the
5343: simulational estimate, it is very possible that the closeness of
5344: the exponent to the boundary of the window of locality gives rise
5345: to very slow convergence to asymptotic scaling. We therefore have
5346: to reserve judgment about the agreement with simulations until
5347: larger scaling ranges were available.
5348:
5349: Similar results are obtained for $j=6$, see Fig. \ref{Fig.4}.
5350: Also this case exhibits zero crossing close to the boundary of
5351: locality, with $\tilde\zeta_6^{(2)}\approx 6.98$. Again we find close
5352: correspondence between the exponential and Gaussian models. In
5353: terms of $\zeta^{(2)}$ this means $\zeta_6^{(2)}\approx 3.98$. This
5354: number appears higher than the simulational result from
5355: \cite{bif01a}, which estimated $\zeta_6^{(2)}\approx 3.3\pm 0.3$. We
5356: note however that for $j=6$ the log-log plots measured in DNS
5357: \cite{bif01a}
5358: possess a short scaling range.
5359: \begin{center}
5360: %%%%%%%%%%%%%%%%%%%
5361: \begin{figure}
5362: \includegraphics[scale=0.45]{j4.a.ps}
5363: \caption{determinant and zero crossing for the sector $j=4$. The scaling
5364: exponent
5365: computed from the zero crossing is $\zeta^{(2)}_4\approx 1.99$.}
5366: \label{Fig.3}
5367: \end{figure}
5368: %%%%%%%%%%%%%
5369: \end{center}
5370: %%%%%%%%%%%%%%%%%%%
5371: \begin{figure}
5372: \includegraphics[scale=0.45]{j6.a.ps}
5373: \caption{determinant and zero crossing for the sector $j=6$. The scaling
5374: exponent
5375: computed from the zero crossing is $\zeta^{(2)}_6\approx 3.98$.}
5376: \label{Fig.4}
5377: \end{figure}
5378: %%%%%%%%%%%%%
5379: Interestingly enough, the set of exponents $\zeta_j^{(2)}$=2/3,
5380: 1.36, 1.99 and 3.98 for $j=$0, 2, 4 and 6 respectively are in
5381: close agreement with the numbers obtained for the linear pressure
5382: model, $\xi_j^{(2)}$=2/3, 1.25226, 2.01922, 4.04843, for
5383: $j=0,2,4$ and 6 respectively. We reiterate at this point that
5384: the latter set is exact for the linear pressure model, whereas
5385: the former set is obtained within the closure approximation. In fact, the
5386: close correspondence
5387: is not so surprising since he linearization of Navier-Stokes equations for
5388: small anisotropy
5389: results in a linear operator which is very close to the one that exists
5390: naturally in the Linear
5391: Pressure Model.
5392: Numerical results \cite{bif01a,bif02,bif03a} obtained at moderate Re
5393: and with strong anisotropies
5394: show a small disagreement with the numbers calculated in the closure
5395: approximation.
5396: We do not expect a much more precise
5397: theoretical evaluation of these exponents before numerical and experimental
5398: data at higher Re are obtained and the
5399: intermittency problem in the isotropic sector is fully settled.
5400: %%%%%%%%%%%%%%%%%%%%
5401: %%%%%%%%%%%%%%%%%%%%
5402: \section{Analysis of Experimental Data}
5403: %%%%%%%%%%%%%%%%%%%%
5404: %%%%%%%%%%%%%%%%%%%%
5405: \label{chap:experiment}
5406: The major difficulty in applying the SO(3) decomposition to
5407: experimental data lies in the fact that one never has the whole
5408: field $\B u(\B x)$. We thus cannot project the statistical objects
5409: onto chosen basis functions $\B B_{qj m}$ and simply integrate out
5410: all other contributions. Rather, we need to extract the wanted information
5411: laboriously by fitting partially resolved data,
5412: or to measure quantities that do not have projections on the isotropic
5413: sector, to see right away the anisotropic contributions. We begin with the
5414: first
5415: approach.
5416: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5417: \subsection{Anisotropic Contribution to the Statistics of the
5418: Atmospheric Boundary Layer}
5419: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5420: \label{atmospheric}
5421: The atmospheric boundary layer offers a natural laboratory of
5422: turbulence that is unique in offering extremely high Re
5423: number. Students of turbulence interested in the scaling
5424: properties that are expected to be universal in the limit $Re\to
5425: \infty$ are thus attracted to atmospheric measurements. On the
5426: other hand the boundary layer suffers from strong inhomogeneity
5427: (explicit dependence of the mean velocity on the height) which
5428: leads to strong anisotropies such that the vertical and the
5429: horizontal directions are quite distinguishable. In addition, one
5430: may expect the boundary layer to exhibit large-scale quasi
5431: 2-dimensional eddys whose typical decay times and statistics may
5432: differ significantly from the generic 3-dimensional case. The aim
5433: of this section is to review systematic methods of data analysis
5434: that attempt to resolve such difficulties, leading to a useful
5435: extraction of the universal, 3-dimensional aspects of turbulence.
5436:
5437: Obviously, to isolate tensorial components belonging to other than
5438: isotropic sectors one needs to collect data from more than one vector
5439: component of the velocity field. Having {\em two} probes is actually
5440: sufficient to read surprisingly rich information about anisotropic
5441: turbulence. In the experiments discussed in this subsection two types
5442: of geometry were employed, one consisting of two probes at the same
5443: height above the ground and the other with the two probes separated
5444: vertically. In both cases the inter-probe separation is orthogonal to
5445: the mean wind.
5446: %%%%%%%%%%%%%%%%%%%%%%%%%%%
5447: \subsubsection{Experiments, Data Sets and the Extraction of
5448: Structure Functions}
5449: \label{expsetup}
5450: The results presented in this subsection are based on two experimental
5451: setups \cite{ara98,kur00,kur00a}, which are denoted
5452: throughout as {\rm I} and {\rm II}
5453: respectively. In both setups the data were acquired over the salt
5454: flats in Utah with a long fetch. In set {\rm I} the data were
5455: acquired simultaneously from two single hot wire probes at a
5456: height of 6 m above the ground, with a horizontal separation of
5457: 55 cm, nominally orthogonal to the mean wind. The Taylor
5458: microscale Reynolds number was about 10,000. Set {\rm II} was
5459: acquired from an array of three cross-wires, arranged {\em above}
5460: each other at heights 11 cm, 27 cm and 54 cm respectively. The
5461: Taylor microscale Reynolds numbers in this set were 900, 1400 and
5462: 2100 respectively.
5463: Table~1 lists a few relevant facts about the data records
5464: analyzed here. The various symbols have the following meanings:
5465: $\overline U$ = local mean velocity, $u^{\prime}$ =
5466: root-mean-square velocity, $\eb$ = energy
5467: dissipation rate obtained by the assumption of local isotropy and
5468: Taylor's hypothesis, $\eta$ and $\lambda$ are the Kolmogorov and
5469: Taylor length scales, respectively, the microscale Reynolds number
5470: $R_{\lambda} \equiv u^{\prime} \lambda/\nu$, and $f_s$ is the
5471: sampling frequency.
5472:
5473: For set I it is important to test whether the separation between the two
5474: probes is indeed orthogonal to the mean wind. (We do not need to
5475: worry about this point in set {\rm II}, since the probes are
5476: above each other). To do so one computes the cross-correlation
5477: function $\langle u_1(t+\tau)u_2(t)\rangle$. Here, $u_1$ and
5478: $u_2$ refer to velocity fluctuations in the direction of the mean
5479: wind, for probes 1 and 2 respectively. If the separation were
5480: precisely orthogonal to the mean wind, this quantity should be
5481: maximum for $\tau=0$. Instead, for set I, it was found that the maximum
5482: shifted slightly to $\tau=0.022$ s, implying that the separation
5483: was not precisely orthogonal to the mean wind. To correct for
5484: this effect, the data from the second probe were time-shifted by
5485: 0.022 s. This amounts to a change in the actual value of the
5486: orthogonal distance. The effective distance is
5487: $\Delta \approx 54$ cm (instead of the 55 cm that was set
5488: physically).
5489: %%%%%%%%%%%%%%%%%%%%%%%%%
5490: \begin{table}
5491: \begin{tabular} {|c|c|c|c|c|c|c|c|c|}
5492: \hline
5493: Height&$\overline U$ & $u^\prime$ &$10^2 \langle \eb
5494: \rangle $,& $\eta$ & $\lambda$ & $R_{\lambda}$ & $f_s,$ per & \#
5495: of \\meters& ms$^{-1}$ & ms $^{-1}$
5496: & m $^2$ s$^{-3}$ & mm & cm & & channel, Hz & samples\\
5497: \hline 6 & 4.1 & 1.08 & $1.1$ & 0.75 & 15 & 10,500 & 10,000 & $4 \times
5498: 10^7$\\
5499: \hline\hline 0.11 & 2.7
5500: & 0.47 & $6.6 $ & 0.47 & 2.8 & 900 & 5,000 & $ 8 \times 10^6$\\
5501: 0.27 &3.1 & 0.48 & 2.8& 0.6 & 4.4 & 1400& 5,000 &$8 \times 10^6$\\
5502: 0.54 &3.51 &0.5& 1.5& 0.7& 6.2&2100& 5,000& $8 \times
5503: 10^6$\\
5504: \hline
5505: \end{tabular}
5506: \caption{Data sets I (first line) and II (second-fourth lines).}
5507: \end{table}
5508: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5509: The coordinates were chosen such that the mean wind direction is along
5510: the 3-axis, the vertical is along the 1-axis and the third
5511: direction orthogonal to these is the 2-axis. We denote these
5512: directions by the three unit vectors $\hat {\B n}$, $\hat {\B m}$,
5513: and $\hat {\B p}$ respectively. The raw data available from set I
5514: is $u^{(3)}(t)$ measured at the positions of the two probes. In
5515: set II each probe reads a linear combination of $u^{(3)}(t)$ and
5516: $u^{(1)}(t)$ from which each component is extractable. From this
5517: raw data we would like to compute the scale-dependent structure
5518: functions, using the Taylor hypothesis to surrogate space for
5519: time. This needs a careful discussion.
5520: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5521: \subsubsection{Theoretical constructs: the Taylor Hypothesis, Inner
5522: and Outer Scales} \label{Taylor}
5523: Decades of research on the statistical aspects of thermodynamic
5524: turbulence are based on the Taylor Hypothesis \cite{tay38}, which
5525: asserts that the fluctuating velocity field measured by a given
5526: probe as a function of time, $\B u(t)$ is the same as the velocity
5527: $\B u(r/\overline{U})$ where $\overline{U}$ is the mean velocity
5528: and $r$ is the distance to a position ``upstream'' where the
5529: velocity is measured at $t=0$. The natural limitation on the
5530: Taylor hypothesis is provided by the typical decay time of
5531: fluctuations of scale $r$. Within a K41 scaling theory this time
5532: scale is the turn-over time $r/\sqrt{S(r)}$ where $S(r)\equiv
5533: S^{\alpha\alpha}(r)$. With this estimate the Taylor Hypothesis is
5534: expected to be valid when $\sqrt{S(r)}/\overline{U} \to 0$. Since
5535: $S(r)\to 0$ when $r\to 0$, the Taylor hypothesis becomes exact in
5536: this limit. We will use this to calibrate the units when we
5537: employ two different probes and read a distance from a
5538: combination of space and time intervals.
5539:
5540: The Taylor hypothesis has also been employed when the mean
5541: velocity vanishes, and instead of $\overline{U}$ one uses the
5542: root-mean-square $u'$. Ref.\cite{lpp99} has presented a
5543: detailed analysis of the consequences of the Taylor hypothesis on
5544: the basis of an exactly soluble model. In particular
5545: ways were proposed there to minimize the systematic errors introduced by
5546: the
5547: use of the Taylor hypothesis. In light of that analysis we will
5548: use here an ``effective" wind $U_{eff}$ which for surrogating the
5549: time data of a single probe is made of a combination of the mean
5550: wind $\overline{U}$ and the root-mean-square $u'$,
5551: \begin{equation} U_{eff}
5552: \equiv \sqrt{\overline{U}^2+(\tilde b u')^2} \ , \label{defUeff}
5553: \end{equation}
5554: where $\tilde b$ is a dimensionless parameter. Evidently, when we employ
5555: the Taylor hypothesis in log-log plots of structure functions
5556: using time series measured in a {\em single} probe, the value of
5557: the parameter $\tilde b$ is irrelevant, changing just the (arbitrary)
5558: units of length (i.e the arbitrary intercept). When we used data
5559: collected from two probes, we mix read distance and surrogated
5560: distance, and the parameter $\tilde b$ becomes a unit fixer. The
5561: numerical value of this parameter is found in \cite{lpp99} by
5562: the requirement that the surrogated and directly measured
5563: structure functions coincide in the limit $r\to 0$. When we do
5564: not have the necessary data we will use values of $\tilde b$ suggested
5565: by the exactly soluble model treated in \cite{lpp99}. The choice
5566: of these values can be justified a posteriori by the quality of the
5567: fit of to the predicted scaling functions.
5568:
5569: When we have two probes placed at different heights the mean
5570: velocity and $u'$ as measured by each probe do not coincide. In
5571: applying the Taylor hypothesis one needs to decide which value of
5572: $U_{eff}$ is most appropriate. This question has been addressed
5573: in detail in ref.\cite{lpp99}
5574: with the final conclusion that the
5575: choice depends on the velocity profile between the probe. In the
5576: case of {\em linear} shear the answer is the precise average
5577: between the two probes, \begin{equation} U_{eff} \equiv
5578: \sqrt{{\overline{U_1}^2+\overline{U_2}\over 2}^2+\tilde b{{u_1'}^2
5579: +{u_1'}^2\over2}} \ , \label{defUeff2}
5580: \end{equation}
5581: where the subscripts 1,2 refer to the two probes respectively.
5582: In all the subsequent expressions we will therefore denote separations by
5583: $r$, and invariably this will mean Taylor-surrogated time differences. The
5584: effective velocity will be (\ref{defUeff}) or (\ref{defUeff2}) depending
5585: on having probes at the same height or at different heights. The value of
5586: $\tilde b$ will be $\tilde b=3$ following ref.\cite{lpp99}. It can be shown
5587: that the
5588: computed scaling exponents are not sensitive to the changing $b$.
5589: (They change by a couple of percents upon changing $b$ by 30\%.)
5590: In seeking scaling behavior one needs to find the inner and outer
5591: scales. Below the inner scale all structure functions have an
5592: analytic dependence on the separation, $S(r)\sim r^2$, and above
5593: the outer scale the structure functions should tend to a constant
5594: value. We look at the longitudinal structure functions
5595: $$
5596: S^{33}(r) = \langle(u^3(x + r) - u^3(x))^2\rangle $$
5597: computed from a single probe in set I and from the probe at
5598: $0.54$m in set II , see Fig.\ref{susanpaper.scrange_long.eps}.
5599: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5600: \begin{figure}
5601: \includegraphics[scale=0.7]{susanpaper.scrange_long.eps}
5602: \caption{Raw log-log plot of the longitudinal component of the
5603: 2nd order structure function.} \label{susanpaper.scrange_long.eps}
5604: \end{figure}
5605: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5606: We simultaneously consider the transverse structure function
5607: $$
5608: S^{11}(r) = \langle(u^1(x + r) - u^1(x))^2\rangle $$
5609: computed from the probe at $0.54$m in set II, see Fig.
5610: \ref{susanpaper.scrange_trans.eps}.
5611: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5612: \begin{figure}
5613: \includegraphics[scale=0.7]{susanpaper.scrange_trans.eps}
5614: \caption{Raw log-log plot of the transverse component of the 2nd
5615: order structure function.} \label{susanpaper.scrange_trans.eps}
5616: \end{figure}
5617: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5618: The spatial scales are computed using the local mean wind in both
5619: cases since the scaling exponent for the
5620: single-probe structure function are not expected
5621: to be affected by the choice of
5622: convection velocity. This choice does determine the value of $r$
5623: corresponding to a particular time scale however. One may expect
5624: that any correction to the numerical value of $r$ is small for a
5625: different choice of convection velocity, and not crucial for the
5626: qualitative statements that follow. In
5627: Fig.~\ref{susanpaper.scrange_long.eps} we clearly see the $r^2$
5628: behavior characterizing the transition from the dissipative to
5629: the inertial range. As is usual, this behavior persists about a
5630: half-decade above the ``nominal" Kolmogorov length scale $\eta$.
5631: There is a region of cross-over and then the isotropic scaling
5632: $\sim r^{0.68}$ expected for small scales in the inertial range
5633: begins. We thus have no difficulty at all in identifying the
5634: inner scale, it is simply revealed as a natural crossover length
5635: in this highly resolved data.
5636: \noindent
5637: We understand by now that we cannot expect to be
5638: able to fit with this single exponent for larger scales and must
5639: include scaling contributions due to anisotropy. We expect that
5640: the contributions due to anisotropy will account for scaling
5641: behavior up to the outer scale of a 3-dimensional flow patterns.
5642: The question therefore is how to identify what this large scale
5643: is.
5644: One approach would be to simply use the scale where the structure
5645: function tends to a constant, which corresponds to the scale
5646: across which the velocity signal has decorrelated. It becomes
5647: immediately apparent that this is not a reasonable estimate of
5648: the relevant large scale. Fig.~\ref{susanpaper.scrange_long.eps}
5649: shows that the structure function stays correlated up to scales
5650: that are at least an order of magnitude larger than the height at
5651: which the measurement is made. On the other hand, if we look at
5652: the transverse structure function computed from the probe at
5653: $0.54$m, Fig.~\ref{susanpaper.scrange_trans.eps} we see that it
5654: ceases to exhibit scaling behavior at a scale that is of the
5655: order of the height of the probe.
5656: \noindent
5657: It appears that we are observing extremely flat eddys that are
5658: correlated over very long distances in the horizontal direction
5659: but have a comparatively small correlation lengths in the
5660: direction perpendicular to the boundary. Since we know that the
5661: presence of the boundary must limit the size of the largest
5662: 3-dimensional structures, the height of the probe should be
5663: something of an upper bound on the largest 3-dimensional flow
5664: patterns that we can detect in experiments. Thus we arrive at a
5665: qualitative understanding of the kind of flow that is observed in
5666: these atmospheric measurements. The size of the largest
5667: 3-dimensional structures is determined by the decorrelation
5668: length of the transverse structure function. This is because the
5669: transverse components of the velocity are unaffected by the
5670: extended, persistent, 2-dimensional eddys that govern the
5671: behavior of the longitudinal components. The theory of scaling
5672: behavior in 3-dimensional turbulence can usefully be applied to
5673: only those flow patterns that are truly 3-dimensional. The
5674: extended flat eddys must be described in terms of a separate
5675: theory, including maybe notions of 2-dimensional turbulence which
5676: has very different scaling properties \cite{kra67}. Such considerations
5677: are outside the scope of this review.
5678: Rather, in the following
5679: analysis, the outer-scale was chosen to be of the order of the
5680: decorrelation length of the {\em transverse} structure function
5681: (where available) or of the height of the probe. We will see
5682: below that up to a factor of 2 these are the same; taking $L$ to
5683: be as twice the height of the probe is consistent with all
5684: data. We use this estimate in the study of both transverse and
5685: longitudinal objects.
5686: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5687: \subsubsection{Extracting the universal exponents of higher $j$
5688: sectors}
5689: \label{extract}
5690: %\label{j2com}
5691: We consider the second order structure function
5692: \begin{equation}\label{Sab} S^{\alpha\beta}({\bf r}) =
5693: \la (u^\alpha({\bf x} + {\bf r}) - u^\alpha({\bf x})) (u^\beta({\bf
5694: x} + {\bf r}) - u^\beta({\bf x})) \ra. \end{equation} The lowest
5695: order anisotropic contribution to the symmetric (in indices),
5696: even parity (in ${\bf r}$ due to homogeneity), second-order
5697: structure function is the $j=2$ component of the SO(3) symmetry
5698: group. Ref.~\cite{ara98} presents a derivation of the $m=0$
5699: axisymmetric (invariant under rotation about the 3-axis) part of
5700: the $j=2$ contribution to this structure function in homogeneous
5701: turbulence. The derivation of the full $j=2$ contribution to
5702: the symmetric, even parity structure function appears in Appendix
5703: \ref{app:fullj2}.
5704: Fig.~\ref{Fig.1}~b shows the
5705: fit to the structure function computed from a single probe in set
5706: I
5707: \begin{equation}\label{tta0}
5708: S^{33}(r,\theta = 0) = \langle (u_1^{(3)}(x+r) -
5709: u_1^{(3)}(x))^2\rangle, \end{equation} where the subscript $1$
5710: denotes one of the two probes, with just the $j=0$ contribution.
5711: The best-fit exponent for the range $0<r/\Delta<4.5$ is
5712: $\zeta^{(2)}_0=0.68\pm0.01$ (Fig.~\ref{Fig.1}~a). Above this range, was
5713: impossible to obtain a good fit to the data with just the
5714: isotropic exponent and Fig.~\ref{Fig.1}~b shows the peel-off from
5715: isotropic behavior above $r/\Delta = 4$.
5716: \begin{center}
5717: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5718: \begin{figure}
5719: \includegraphics[scale=0.7]{susanpaper.Fig1.eps}
5720: \caption{The single-probe structure
5721: function computed from data set I. (a) shows the $\chi^2$
5722: minimization by the best-fit value of the exponent in the
5723: isotropic sector $\zeta_0^{(2)}\approx 0.68$ for the single-probe
5724: structure function in the range $0 < r/\Delta <4.5$. (b) shows
5725: the fit using the best value of $\zeta_0^{(2)}$ obtained in (a),
5726: indicating the peel-off from isotropic behavior at the end of
5727: the fitted range.}\label{Fig.1}
5728: \end{figure}
5729: \end{center}
5730: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5731: To find the $j=2$ anisotropic exponent one needs to use data taken
5732: from the two probes. To clarify the procedure we show in
5733: Fig.\ref{Figexpsetup} the geometry of set I.
5734: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5735: \begin{figure}
5736: \includegraphics[scale=0.3]{susanpaper-setup.eps}
5737: \caption{Diagrammatic illustration of the experimental set-up.
5738: Shown is the positioning of the probes with respect to the mean
5739: wind and how the Taylor hypothesis is employed}
5740: \label{Figexpsetup} \end{figure}
5741: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5742: What was computed is actually
5743: $$
5744: S^{33}(r,\theta)=\langle [u^{(3)}_1( U_{eff} t + U_{eff}t_{\tilde
5745: r})- u^{(3)}_2( U_{eff}t)]^2\rangle $$ Here
5746: $\theta=\arctan(\Delta/ U_{eff}t_{\tilde r})$, $t_{\tilde
5747: r}=\tilde r/ U_{eff}$, and $r=\sqrt{\Delta^2+(\bar
5748: U_{eff}t_{\tilde r})^2}$. $U_{eff}$ was as in Eq.(\ref{defUeff})
5749: with $b=3$. We will refer from now on to such quantities as
5750: \begin{equation}\label{ttadep}
5751: S^{33}(r,\theta) = \langle (u_1^{(3)}(x+r) -
5752: u_2^{(3)}(x))^2\rangle \ . \end{equation}
5753:
5754: Next, one may fix the scaling exponent of the isotropic sector as
5755: $0.68$ and find the $j=2$ anisotropic exponent that results from
5756: fitting to the full $j=2$ tensor contribution. Finally, one needs
5757: to fit the objects
5758: in Eqs.~(\ref{tta0}) and (\ref{ttadep}) to the sum of the $j=0$ (with
5759: scaling exponent $\zeta^{(2)}_0 = 0.68$) and the $j=2$ contributions
5760: (see Appendix \ref{app:fullj2})
5761: \begin{eqnarray}
5762: &&S^{33}(r,\theta)=S^{33}_{j=0}(r,\theta)+ S^{33}_{j=2}(r,\theta)
5763: =c_0\left({r\over \Delta}\right)^{\zeta^{(2)}_0} \Big[ 2
5764: +\zeta^{(2)}_0-\zeta^{(2)}_0 \cos^2\theta\Big] \nonumber\\
5765: &+&a\left({r\over \Delta}\right)^{\zeta_2^{(2)}} \Big[
5766: (\zeta_2^{(2)}+2)^2 -\zeta_2^{(2)}
5767: (3\zeta_2^{(2)}+2)\cos^2\theta+2\zeta_2^{(2)}(\zeta_2^{(2)}-2)\cos^4\theta\B
5768: ig] \nonumber\\
5769: &+&b\left({r\over \Delta}\right)^{\zeta_2^{(2)}} \Big[
5770: (\zeta_2^{(2)}+2) (\zeta_2^{(2)}+3)-
5771: \zeta_2^{(2)}(3\zeta_2^{(2)}+4)\cos^2\theta+(2\zeta_2^{( 2)}+1)
5772: (\zeta_2^{(2)}-2)\cos^4\theta\Big] \nonumber \\ &+&a_{9,2,1} \left({r\over
5773: \Delta}\right)^{\zeta_2^{(2)}}
5774: \Big[-2\zeta_2^{(2)} (\zeta_2^{(2)}+2) \sin\theta\cos\theta
5775: + 2\zeta_2^{(2)}(\zeta_2^{(2)}-2)\cos^3\theta\sin\theta \Big]\nonumber\\
5776: &+&a_{9,2,2} \left({r\over
5777: \Delta}\right)^{\zeta_2^{(2)}}\Big[-2\zeta_2^{(2)}
5778: (\zeta_2^{(2)}-2) \cos^2\theta\sin^2\theta\Big]
5779: \nonumber\\
5780: &+&a_{1,2,2}\left({r\over \Delta}\right)^{\zeta_2^{(2)}}
5781: \Big[-2\zeta_2^{(2)} (\zeta_2^{(2)}-2)\sin^2\theta\Big].
5782: \label{fulltens}
5783: \end{eqnarray}
5784: The above fit was performed using values of $\zeta_2^{(2)}$ ranging
5785: from $0.5$ to $3$. The best value of this exponent is the one that
5786: minimizes the $\chi^2$ for the fits. From Fig.\ref{chj2} one may read
5787: the best value to to be $1.38\pm0.15$. The fits with this choice of
5788: exponent are displayed in Fig.\ref{sfj2full}. The corresponding
5789: values of the 5 fitted coefficients can be found in the paper
5790: \cite{ara98}. The range of scales that are fitted to this expression
5791: is $1 < r/\Delta < 25$. We thus conclude that that structure
5792: functions which is symmetric in $\B r$ exhibits scaling behavior
5793: over the whole scaling range, but this important fact is missed if one
5794: does not consider a superposition of the $j=0$ and $j=2$
5795: contributions.
5796: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5797: \begin{figure}
5798: \includegraphics[scale=0.7]{susanpaper.Fig2.eps}
5799: \caption{The $\chi^2$ minimization
5800: by the best-fit value of the exponent in the $j=2$ anisotropic
5801: sector from the fit to both the $\theta=0$ and the
5802: $\theta$-dependent structure function in the range $0 < r/\Delta
5803: <25$. }\label{chj2} \end{figure}
5804: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5805: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5806: \begin{figure}
5807: \includegraphics[scale=0.7]{susanpaper.Fig3.eps}
5808: \caption{The structure functions
5809: computed from data set I and fit with the $j=0$ and full $j=2$
5810: tensor contributions using the best fit values of exponents
5811: $\zeta^{(2)}_0=0.68$ and $\zeta_2^{(2)}=1.38$ in the range $0 <
5812: r/\Delta <25$. Panel (a) shows the fit to the single-probe
5813: ($\theta =0$) structure function and panel (b) shows the fit to
5814: the $\theta$-dependent structure function.}\label{sfj2full}
5815: \end{figure}
5816: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5817: Finally, let us note that the value of the exponent is
5818: perfectly in agreement with the analysis of numerical simulations
5819: \cite{ara99} in which one can comfortably integrate the
5820: structure function against the basis functions, eliminating all
5821: contributions except $j=2$ (see next section).
5822: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5823: %\begin{table}
5824: %\begin{tabular} {|c|c|c|c|c|c|c|c|}
5825: %$\zeta^{(2)}_0$ & $\zeta_2^{(2)}$ & $c_0 \times 10^3 $ & $a \times
5826: %10^3$ & $b
5827: %\times 10^3 $ & $a_{9,2,1} \times 10^3 $ & $a_{9,2,2} \times 10^3 $ &
5828: %$a_{1,2,2} \times 10^3$\\
5829: %\hline 0.68 & 1.38 $\pm 0.10$ & 7 $\pm .5$ & -3.2 $\pm$ 0.3 & 2.6
5830: %$\pm $0.3 & -0.14$ \pm$ 0.02 & -5.6 $\pm $ 0.7& -4 $\pm $ 0.5
5831: %\end{tabular}
5832: %\caption{The scaling exponents and the 6 coefficients in units of
5833: %(m/sec)$^2$ as determined from the nonlinear fit of Eq. 7 to data
5834: %set I.}
5835: %\end{table}
5836: %%%%%%%%%%%%%%%%%%%%%%%%%%%
5837: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5838: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5839: \subsubsection{Extracting the j=1 component}
5840: \label{j1com}
5841: In homogeneous flows, it follows from symmetry and parity that the often
5842: computed and
5843: widely analyzed structure
5844: function as defined in Eq. (\ref{Sab}) does not possess any contribution
5845: from the
5846: $j=1$ sector. The lowest order anisotropic contribution belongs to the
5847: $j=2$ sector. In order to isolate the scaling
5848: behavior of the $j=1$ contribution in atmospheric shear flows
5849: we must either explicitly construct a new tensor object which will
5850: allow for such a contribution, or see if it can be extracted
5851: from the structure function itself computed in the case of {\em
5852: inhomogeneity}. We have pursued both avenues. In the former, we
5853: construct the tensor
5854: \begin{equation}\label{Tab}
5855: T^{\alpha\beta}({\B r}) = \langle u^\alpha({\B x} + {\B r}) -
5856: u^\alpha({\B x}))(u^\beta({\B x} + {\B r}) + u^\beta({\B
5857: x}))\rangle. \end{equation} It is easily seen that the function
5858: vanishes both in the case of $\alpha=\beta$ and when ${\B r}$ is
5859: in the direction of homogeneity. From data set II we can
5860: calculate this function for non-homogeneous (in the shear
5861: direction) scale-separations. In general, this will exhibit mixed
5862: parity and symmetry and therefore, to minimize as far as possible
5863: the final number of fitting parameters we look at only the
5864: antisymmetric contribution. We derive the tensor contributions in
5865: the $j=1$ sector for the antisymmetric case in Appendix \ref{app:fullj1}
5866: and
5867: use this to fit for the unknown $j=1$ exponent.
5868: Below we describe the results of this analysis. Next, we computed
5869: the $\theta$-dependent structure function from set II. We expect
5870: that this could exhibit the $j=1$ component as inhomogeneity does
5871: not allow us to apply incompressibility in the different symmetry
5872: and parity sectors to eliminate this contribution as in the case
5873: of the homogeneous structure function. This structure function is
5874: symmetric but of mixed parity. We derive the tensor contributions
5875: in the $j=1$ sector for the symmetric case in Appendix \ref{app:fullj1}
5876: and use this to fit for the $j=1$ exponent.
5877: \subsubsection{Antisymmetric Contribution}
5878: We consider the tensor object in Eq.~(\ref{Tab}). In order to have
5879: as few parameters as possible in the fitting procedure, we take
5880: the antisymmetric part
5881: $$
5882: {\widetilde T}^{\alpha\beta}({\B r}) = {T^{\alpha\beta}({\B r})
5883: - T^{\beta\alpha}({\B r}) \over 2} = \langle u^\alpha({\B
5884: x})u^\beta({\B x} + {\B r})\rangle - \langle u^\beta({\B
5885: x})u^\alpha({\B x} + {\B r})\rangle
5886: $$ which will only have contributions from the antisymmetric
5887: $j=1$
5888: basis tensors. An additional useful property of this object is
5889: that it does not have any contribution from the isotropic
5890: helicity-free $j=0$ sector due to its antisymmetry. This allows
5891: us to isolate the $j=1$ contribution and determine its scaling
5892: exponent $\zeta_1^{(2)}$ starting from the smallest scales
5893: available. Using data from the probes at $0.27$m (probe 1) and at
5894: $0.11$m (probe 2) we calculate
5895: $$
5896: {\widetilde T}^{31}({\B r}) = \langle u_2^{(3)}(\B
5897: x)u_1^{(1)}(\B x + \B r)\rangle - \langle u_1^{(3)}(\B x + \B
5898: r)u_2^{(1)}(\B x)\rangle \ ,$$ where again
5899: super-scripts denote the velocity component and sub-scripts
5900: denote the probe at which this component is measured. We want to
5901: fit this object to the tensor form derived in Appendix (\ref{app:fullj1}),
5902: namely:
5903: $$ {\widetilde
5904: T}^{31}(r,\theta,\phi=0) = - a_{3,1,0}r^{\zeta_1^{(2)}}\sin\theta
5905: + a_{2,1,1}r^{\zeta_1^{(2)}} +
5906: a_{3,1,-1}r^{\zeta_1^{(2)}}\cos\theta.
5907: $$
5908: Fig.~\ref{chj1T} gives the $\chi^2$ minimization of the fit as a
5909: function of $\zeta_1^{(2)}$ and we use the best value of $1 \pm
5910: 0.15$ for the final fit. This is shown in the left panel.
5911: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5912: \begin{figure*}
5913: \includegraphics[scale=0.55]{susanpaper.Fig4.eps}
5914: \includegraphics[scale=0.8]{susanpaper.Fig5.eps}
5915: \caption{Right: the $\chi^2$ minimization
5916: by the best-fit value of the exponent $\zeta_1^{(2)}$ of the
5917: $j=1$ anisotropic sector from the fit to $\theta$-dependent
5918: ${\widetilde T}^{31}(r,\theta)$ function in the range $0 <
5919: r/\Delta < 2.2$. Left: The fitted ${\widetilde
5920: T}^{31}(r,\theta)$ function. The dots indicate the data and the
5921: line is the fit. }\label{chj1T}
5922: \end{figure*}
5923: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5924: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5925: %\begin{figure}
5926: %\epsfxsize=8truecm \epsfysize=6truecm
5927: %\epsfbox{susanpaper.Fig5.eps}
5928: %\caption{The fitted ${\widetilde
5929: %T}^{31}(r,\theta)$ function. The dots indicate the data and the
5930: %line is the fit.} \label{Tfit}
5931: %\end{figure}
5932: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5933: The fit in Fig.~\ref{chj1T} peels off at the end of the fitted
5934: range.
5935: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5936: %\begin{table}
5937: %\begin{tabular}{|c|c|c|c|}
5938: %\hl
5939: %$\zeta_1^{(2)}$ & $a_{3,1,0} $ & $a_{2,1,1}$ & $a_{3,1,-1}$
5940: %\\ \hline $1 \pm 0.15 $ & $ 0.0116 \pm 0.0013 $ & $ 0.0124 \pm 0.0013 $ &
5941: %$-0.0062 \pm 0.0008$\\
5942: %\end{tabular}
5943: %\caption{The values of the exponents and coefficients (in units
5944: %of (m/sec)$^2$) obtained from the fit to the function
5945: %${\widetilde T}^{31}(r,\theta)$.}
5946: %\end{table}
5947: The maximum range over which one can fit is of the order of the
5948: height of the probes and again, this is consistent with the
5949: considerations presented above.
5950: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5951: \subsubsection{Symmetric Contribution}
5952: Finally, we compute the structure function Eq.~(\ref{ttadep}) where
5953: the subscripts denote probe 1 at $0.27$m and probe 2 at $0.11$cm.
5954: As discussed in Appendix (\ref{app:fullj1}),
5955: since the scale separation
5956: has an inhomogeneous component, we expect a contribution from the
5957: $j=1$ anisotropic sector and we would like to extract what the
5958: scaling exponent in this sector is. Note that the $j=0$ sector
5959: contributes {\em two} independent tensor forms with coefficients
5960: we will denote by $c_1$ and $c_2$, since incompressibility does
5961: not provide a constraint to relate them. This fact combined with
5962: Eq.~(\ref{b15}) gives us the the tensor form to which we must fit our
5963: function
5964: \begin{eqnarray}\label{sj1tens}
5965: &&S^{33}(r,\theta) = c_1r^{\zeta_0^{(2)}} + c_2r^{\zeta_0^{(2)}}\cos^2\theta
5966: +a_{1,1,0}r^{\zeta_1^{(2)}}\cos\theta
5967: +a_{7,1,0}r^{\zeta_1^{(2)}}2\cos\theta \\&&
5968: +a_{9,1,0}r^{\zeta_1^{(2)}} \cos^3\theta +
5969: a_{8,1,1}r^{\zeta_1^{(2)}} (-2\cos\theta\sin\theta)
5970: + a_{1,1,-1}r^{\zeta_1^{(2)}} \sin\theta +
5971: a_{9,1,-1} r^{\zeta_1^{(2)}}\cos^2\theta\sin\theta \nonumber \end{eqnarray}
5972: We fix the exponent $\zeta^{(2)}_0$ to be $0.68$ and perform fits with
5973: varying values of $\zeta_1^{(2)}$ for 8 unknown coefficients. The
5974: best value of $\zeta_1^{(2)}$ is obtained for the range
5975: $0<r/\Delta<4.2$ and is $1.05 \pm 0.15$ as is shown in
5976: Fig.~\ref{chj1S6}. In the left panel we show the fit to the
5977: data using this value of the exponent.
5978: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5979: %\begin{table}
5980: %\begin{tabular}{|c|c|c|c|c|c|c|c|c|}
5981: %$\zeta_2$ & $\zeta_1^{(2)}$ & $c_1 \times 10^2 $ & $c_2 \times
5982: %10^2 $ & $(a_{1,1,0}+2a_{7,1,0}) $ & $a_{9,1,0} $ & $a_{8,1,1} $
5983: %& $a_{1,1,-1} $ &
5984: %$a_{9,1,-1} $ \\
5985: %& & & & $\times 10^2$ & $\times 10^2$& $\times 10^2$&$\times
5986: %10^2$ &$\times 10^2$ \\ \hline 0.68 & $1.05 \pm 0.15$ & $-41.9
5987: %\pm 5.8$ & $21.8 \pm 3.2$ &
5988: %$11.6 \pm 1.3 $ & $-7.4 \pm 0.8$ & $8.7 \pm 1.4$ & $ 55.6 \pm 6.9 $ & $
5989: -8.2
5990: %\pm 0.98 $ \\
5991: %\end{tabular}
5992: %\caption{The values of the exponents and coefficients (in units
5993: %of (m/sec)$^2$) obtained from the fit to the inhomogeneous
5994: %structure function.}
5995: %\end{table}
5996: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5997: The fit peels off at the end of the fitted range at the scale on
5998: the order of twice the height of the probe, consistent with the
5999: earlier discussion. There does not exist a well-defined
6000: $\zeta_1^{(2)}$ as given by the standard $\chi^2$ minimization
6001: procedure for ranges smaller or larger than that fitted for in
6002: Fig.~\ref{chj1S6}. The quality of the fit is
6003: good although, as was expected from the large number of
6004: parameters in the fitting function Eq.~(\ref{sj1tens}), $\chi^2$ as
6005: a function of the $\zeta_1^{(2)}$ is not as smooth as for all
6006: previous fits and its minimum is a relatively weak one.
6007: Therefore, we present this result mainly as it provides support
6008: to that of the antisymmetric case in the previous section.
6009: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6010: \begin{figure*}
6011: \includegraphics[scale=0.6]{susanpaper.Fig6.eps}
6012: \includegraphics[scale=0.55]{susanpaper.Fig7.eps}
6013: \caption{Right: the $\chi^2$ minimization
6014: by the best-fit value of the exponent $\zeta_1^{(2)}$ of the
6015: $j=1$ anisotropic sector from the fit to $\theta$-dependent
6016: inhomogeneous structure function in the range $0 < r/\Delta <
6017: 2.2$. Left: The fit to the
6018: inhomogeneous structure function computed as in Eq.(\ref{sj1tens}). The dots
6019: indicate the data and the line is the fit.}\label{chj1S6}
6020: \end{figure*}
6021: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6023: %\begin{figure}
6024: %\epsfxsize=8truecm \epsfysize=6truecm
6025: %\epsfbox{susanpaper.Fig7.eps}
6026: %\caption{The fit to the
6027: %inhomogeneous structure function computed as in Eq.(\ref{sj1tens}). The
6028: dots
6029: %indicate the data and the line is the fit.} \label{Sfit7}
6030: %\end{figure}
6031: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6032: \subsubsection{Summary and conclusions}
6033: In summary, we considered the 2nd order tensor structure
6034: functions of velocity differences in the atmospheric boundary
6035: layers. The following conclusions appear important:
6036: \begin{enumerate}
6037: \item The atmospheric boundary layer exhibits 3-dimensional statistical
6038: turbulence
6039: intermingles with activities whose statistics are quite
6040: different. The latter are eddys with quasi-two dimensional
6041: nature, correlated for hundreds of meters, having little to do
6042: with the three-dimensional fluctuations discussed above. We found
6043: that the
6044:
6045: \item We found that the ``outer scale of turbulence" as measured by the
6046: three-dimensional
6047: statistics is of the order of twice the height of the probe.
6048:
6049: \item The inner scale is the the usual dissipative crossover, which is
6050: clearly seen
6051: as the scale connecting two different slopes in log-log plots.
6052:
6053: \item Between the inner and the outer scales
6054: the sum of the components up to $j=2$
6055: appears
6056: to offer
6057: an excellent representation of the structure function.
6058: \item The scaling exponents $\zeta_j^{(2)}$ are measured as $0.68\pm 0.01,~
6059: 1\pm 0.15, ~1.38\pm0.10$ for $j=0,1,2$ respectively.
6060: \end{enumerate}
6061:
6062: We note that as far as the low order $j$ sectors are concerned, the picture
6063: that emerges for Navier-Stokes turbulence is not different from the linear
6064: advection
6065: problems that were treated in the previous section. If the trends seen here
6066: continue for higher
6067: $j$ values, we can rationalize the apparent tendency toward isotropy with
6068: decreasing scales. If indeed every anisotropic contribution
6069: introduced by the large scale forcing (or boundary conditions)
6070: decays as $(r/L)^{\zeta_j^{(2)}}$ with increasing $\zeta_j^{(2)}$
6071: as a function of $j$, then obviously when $r/L\to 0$ only the
6072: isotropic contribution survives. This is a pleasing notion that
6073: justifies the modeling of turbulence as isotropic at the small
6074: scales.
6075: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6076: \subsection{Homogeneous Shear}
6077: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6078: In this subsection we discuss recent experiments in which anisotropy is
6079: created without
6080: inhomogeneity \cite{she00,she02,she02a}; such experiments are
6081: particularly appealing for our purposes. {\it Homogeneous}-shear
6082: flow can be realized in a wind tunnel by
6083: using a variable
6084: solidity screen followed by flow straighteners.
6085: %\begin{center}
6086: %\begin{figure}[h]
6087: %\scalebox{.9}{\includegraphics[scale=0.6]{shear_omo_lab.ps}}
6088: %\caption{FIGURE of experimental apparatus, warhaft}
6089: %\label{fig:homoshear}
6090: %\end{figure}
6091: %\end{center}
6092: Such a set-up results in a shear flow that remains approximately
6093: constant for the length of the tunnel. To produce high Reynolds
6094: numbers one places an active grid before the shear generating screen
6095: \cite{she00}. In this way $Re_{\lambda}$ can be as high as 1000.\\ To
6096: assess directly the effects of anisotropy it is useful to measure
6097: statistical objects that vanish identically in the isotropic sector. A
6098: possible choice is the set of skewness and hyper-skewness
6099: \cite{fer00,sch00,sch01} as explained in Sect.~(\ref{sec:persis}).
6100: Other purely anisotropic inertial range observables can be defined by
6101: mixing longitudinal and transversal increments with an odd number of
6102: transversal components: \be S^{(p,2q+1)}(\Br) = \left< \delta
6103: u_\ell^{p}(\Br) \delta u_t^{2q+1}(\Br) \right>.
6104: \label{eq:sn_aniso} \ee
6105: Systematic measurements of these anisotropic mixed correlation functions was
6106: reported
6107: in \cite{she02,she02a,kur00}.
6108: From the experimental data it is not possible to exactly disentangle
6109: different anisotropic projections in different sectors. This is because
6110: SO(3) projection requires the knowledge of the whole velocity field
6111: in a 3d sub-volume, something clearly out of reach in any experimental
6112: apparatus. The simplest
6113: working hypotheses one can make is that, due to the hierarchical
6114: organization of anisotropic scaling exponents, the statistical
6115: behavior of quantities as (\ref{eq:sn_aniso}) is dominated, at
6116: scales small enough, by the leading $j=2$ sector. In other words,
6117: the experimental measurements of the scaling properties of
6118: (\ref{eq:sn_aniso}) is the best estimate of the
6119: exponent $\zeta_{2}^{(n=p+2q+1)}$.
6120: %\begin{center}
6121: %\begin{figure}
6122: %\scalebox{.9}{
6123: %\includegraphics[scale=0.6]{warhaftsfaniso.ps}}
6124: %\caption{FIGURE from warhaft. Luca}
6125: %\label{fig:warhaft}
6126: %\end{figure}
6127: %\end{center}
6128: In \cite{she02,she02a} the plots of purely anisotropic quantities like
6129: (\ref{eq:sn_aniso}) up to order $n=8$ with $n=p+2q+1$ were shown. The
6130: data clearly shows that these purely anisotropic structure functions
6131: have quite good power-laws behavior with exponents that are
6132: sub-leading with respect to the exponents of the isotropic structure
6133: functions of the same order, $n$. For example $ S^{(1,3)}(r) \sim
6134: r^{1.56}$ while the fourth-order longitudinal structure function in
6135: isotropic ensembles is known to scale as $ S^{(4,0)}(r) \sim
6136: r^{1.27}$. Similar qualitative and quantitative results were obtained
6137: by analyzing data from an atmospheric boundary layer in \cite{kur00}
6138: and in the boundary layer close to a wall \cite{iacob04}. In the
6139: latter two works, a phenomenological fitting procedure to the large
6140: scale behavior allowed the authors to find a power law for the
6141: anisotropic structure functions which pertain to a much larger range
6142: of scales. We draw the reader's attention to the discrepancy in the
6143: best fit for the scaling exponents founds for $ S^{(1,3)}(r)$ and
6144: $S^{(3,1)}(r)$ in \cite{she02,she02a}. Similar discrepancies are also
6145: reported for higher order structure functions. In our view, this
6146: cannot be taken as evidence that there is a $q$-dependence of the
6147: scaling exponents of the SO(3) projections. First, the anisotropic
6148: exponents are relatively inaccurate due to statistical errors; the
6149: amplitudes of the anisotropic fluctuations are relatively small.
6150: Second, as already said, the experimental data cannot disentangle
6151: exactly the contribution of the $j=2$ sector. Therefore, it may well
6152: be that contributions from the $j=4$ (and higher) sectors affect
6153: differently the correlation functions with different tensorial
6154: structure. Similarly, other experimental investigation focused on the
6155: SO(3) decomposition \cite{sta02,sta03,wander} have found results
6156: depending on the geometric set-up of the analyzing probes. The
6157: experimental analysis of anisotropic turbulence via the SO(3)
6158: decomposition is at its infancy; more refined experimental techniques
6159: are needed before a firm conclusion can be reached on these issues.
6160:
6161: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6162: \subsubsection{Explanation of Persistence of Anisotropies}
6163: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6164: \label{pers_ani}
6165: As discussed in Sect.~(\ref{sec:persis}) there are
6166: numerical and experimental evidences of the persistence of small-scale
6167: anisotropic fluctuations in various instances
6168: \cite{gar98,she00,kur00,pum96,bif01}.
6169: The issue has many important consequences. We would like to
6170: refer to the
6171: violation of the {\it return-to-isotropy} in different
6172: meanings \cite{bif01}. A {\it strong} violation would be implied if the
6173: following set of inequalities between different anisotropic exponents of the
6174: same correlation function were broken:
6175: \begin{equation}
6176: \zeta^{(n)}_0 < \zeta^{(n)}_1 < \cdots < \zeta^{(n)}_j \ ,
6177: \label{eq:SnDecomp}
6178: \end{equation}
6179: i.e. if one, or more, anisotropic sector becomes leading with respect
6180: to the isotropic one. This would destroy the phenomenology of turbulence as
6181: developed
6182: since Kolmogorov's theory in 1941. Turbulence would become more and
6183: more anisotropic at smaller and smaller scales. As a result,
6184: strong non-universalities in small-scales statistics would show up
6185: depending on which anisotropic sector is switched on/off by the
6186: large-scale forcing. A {\it strong} violation of the {\it
6187: return-to-isotropy} postulate has never been observed in
6188: Navier-Stokes turbulence. On the other hand, when the hierarchy
6189: (\ref{eq:SnDecomp}) holds, any dimensionless anisotropic observables
6190: made of ratios between anisotropic and isotropic projections of the
6191: {\it same} correlation function vanishes in the small-scales limit.
6192: For example, focusing on the decomposition of longitudinal structure
6193: functions
6194: (\ref{eq:fundamental}) we may write:
6195: \be
6196: \lim_{r\rightarrow 0}
6197: \frac{S_{jm}^{(n)}(r)}{(S_{00}^{(n)}(r))}\;\sim
6198: r^{\zeta^{(n)}_j-\zeta^{(n)}_0}
6199: \rightarrow 0.
6200: \label{eq:ret-b}
6201: \ee
6202: A new phenomenon occurs when anisotropic fluctuations are assessed
6203: by using dimensionless observables made of {\it different}
6204: correlation functions.
6205: For instance, by
6206: using again the SO($3$) decomposition of longitudinal structure
6207: function (\ref{eq:fundamental})
6208: one may build up anisotropic observables defined as:
6209: \be R_{jm}^{(n)}(r) =
6210: \frac{S_{jm}^{(n)}(r)}{(S_{00}^{(2)}(r))^{n/2}}\;\sim r^{\chi_j^{(n)}}
6211: \;\;\;
6212: \mbox{ with } \;\; {\chi_j^{(n)} = \zeta_j^{(n)}-{ n \over 2} \zeta_0^{(2)}}
6213: \ .
6214: \label{eq:ret-a}
6215: \ee
6216: This is the $n$th order moment of the velocity
6217: probability density function, normalized by its isotropic
6218: second order moments. The quantities defined in (\ref{eq:ret-a}) must
6219: be exactly zero in isotopic ensembles, and should
6220: go to zero as power laws, $R_{jm}^{(n)}(r) \sim r^{j/3}$,
6221: in an anisotropic ensemble in which
6222: the dimensional scaling (\ref{lumleygen}) is satisfied. On the
6223: other hand, results from experiments and numerics show
6224: a much slower decay, and, in some cases, no decay at all
6225: \cite{she00,bif01}. We refer to this phenomenon as {\it weak} violation
6226: of the {\it return-to-isotropy}. Such a weak violation is not in
6227: contradiction
6228: with the inequalities (\ref{eq:SnDecomp}); there the
6229: relative importance of anisotropic fluctuations with respect to
6230: isotropic fluctuation of the {\it same} correlation function are
6231: implied. The violation of the dimensional recovery-of-isotropy is
6232: simply due to the existence of anomalous scaling in the anisotropic
6233: sectors. Indeed, in this case, the exponents, $\chi_j^{(n)}$, governing
6234: the LHS of (\ref{eq:ret-a}) can assume values much smaller than the
6235: dimensional estimate (including negative values!).
6236: This is exactly what is observed in the experiments and numerics. From
6237: Table~3 one realizes that due to the
6238: presence of anomalous scaling in the anisotropic sectors we have a
6239: slow recovery-of-isotropy,
6240: in agreement with what was explained before.
6241:
6242: The anisotropic observables built in terms of the generalized flatness
6243: or skewness discussed in section (\ref{sec:persis}) are nothing but
6244: Eq.~(\ref{eq:ret-a}) evaluated at the dissipative length scale,
6245: $r=\eta$. Therefore, the ``persistence-of-anisotropies" discussed in
6246: \cite{pum96,she00} can be
6247: explained invoking the very same reasoning.
6248: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6249: \subsubsection{Summary of experimental results: universality of the
6250: anisotropic sectors} \label{expuniver}
6251: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6252:
6253: Comparing the results obtained in
6254: \cite{she02,she02a,kur00} the following picture emerges.
6255: First, all the correlation functions up to $n=10$, show {\it
6256: anomalous} scaling behavior, where
6257: anomalous is meant with respect to the dimensional
6258: Lumley-like prediction discussed in Sect.~(\ref{sec:dim_ana}).
6259: Second, the values of scaling exponents extracted from the two
6260: different experiments \cite{she02a} and \cite{kur00} are in good
6261: qualitative agreement
6262: (see Table~2).
6263: \begin{table}[ht]
6264: \label{tab:scaling_exp}
6265: \begin{center}
6266: \begin{tabular}{|c|c|c|c|c|c|c|c|c|c|}
6267: \hline
6268: $(p,2q+1)$ & (1,1) & (1,3) & (3,1) & (5,1) & (3,3) & (1,5) & (7,1) &
6269: (5,3) & (3,5) \\
6270: \hline
6271: WS & 1.05 & 1.56 & 1.42 & 2.02 & 1.89 & 1.71 & 2.33 &
6272: 2.22 & 1.99 \\
6273: KS & 1.22,1.12 & 1.58,1.54 & - & - & 2.14,2.00& - & - &
6274: - & - \\
6275: %ISO & 0.67 & 1.26 & 1.26
6276: % & 1.75 & 1.75 & 1.75 & 2.15 & 2.15 & 2.15 \\
6277: \hline
6278: \end{tabular}
6279: \caption{Measured scaling exponents for $S^{(p,2q+1)}(\B r)$
6280: of various orders in two experiments. WS corresponds
6281: to \cite{she02a} and KS to \cite{kur00}}
6282: \end{center}
6283: \end{table}
6284: This is an important first confirmation of the {\it universality} of
6285: scaling exponents in the $j=2$ anisotropic sector. Finally, there
6286: exists a clear hierarchy between isotropic and anisotropic scaling
6287: exponents, the latter being always larger for any given order, $n$ of
6288: the correlation function. This hierarchical organization is the
6289: necessary and sufficient requirement for the {\it
6290: return-to-isotropy} to hold, i.e. the small scales statistics
6291: of any correlation function is dominated by the isotropic
6292: fluctuations. Nevertheless the gap between isotropic and anisotropic
6293: exponents, $\zeta_0^{(n)} - \zeta_{2}^{(n)}$, tends to shrink when $n$
6294: increases, implying that anisotropic contributions may exhibit
6295: important sub-leading effects also at very high Re.
6296: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6297: %%
6298: %
6299: % Chapter 5 - Numerical Analysis
6300: %
6301: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6302: %%
6303: \section{Analysis of DNS Data}
6304: \label{chap:numerics}
6305: Direct numerical simulations of turbulence are natural grounds
6306: where the utility of the SO($3$) decomposition can be exploited to its
6307: maximum benefit. The reason
6308: for this is that numerical simulations, in contrast to current
6309: experiments, provide access to the full velocity field at all points
6310: of the turbulent domain. Therefore, the full SO($3$) decomposition can
6311: be realized, without the constraints of best-fits to partial data. Given a
6312: tensor structure function $\B S^{(n)}(\Br)$, cf. Eq.(\ref{Sn}),
6313: we can integrate it against the spherical tensors,
6314: $\B B^{(n)}_{q j m}(\hat{\Br})$ [e.g., (\ref{eq:so3sn})],
6315: on a sphere of radius $r$. These integrations yield the
6316: projection of the structure function on the different sectors of the
6317: SO($3$) group, by virtue of the orthogonality of the basis tensors. On the
6318: other hand DNS suffer from limited Reynolds numbers;
6319: consequently they have relatively short inertial ranges.
6320:
6321: Prior to the introduction of the SO(3) decomposition the numerical
6322: investigations of anisotropic flows were focused on either single
6323: point or two-points correlations, limited, often, to the analysis of
6324: the Fourier transforms in wavevector space. The most recent,
6325: highly-resolved, numerical investigation of this kind was reported in
6326: \cite{ish02}; there the full tensorial properties of the Fourier
6327: transform of the two-point velocity correlation, $Q^{\alpha
6328: \beta}(\Bk) \EqDef \int d\Br e^{i
6329: \Bk \Br} \la u^\alpha(\Bx+\Br)u^\beta(\Bx) \ra $, were calculated in a {\it
6330: homogeneous} shear \cite{hin75,rog82}. The main result is a
6331: confirmation of Lumley's prediction for the scaling exponent of the purely
6332: anisotropic
6333: co-spectrum:
6334: \begin{equation}
6335: E^{\alpha \beta}(k) \sim k^{-7/3}\;\;\; \mbox
6336: {where}\;\;\; E^{\alpha \beta}(k) = \int_{{k \over 2}<|\Bp|<2k} d\Bp
6337: Q^{\alpha \beta}(\Bp) \ , \label{cospec}
6338: \end{equation}
6339: where $\alpha \neq \beta$ to eliminate the
6340: isotropic contribution. Only recently DNS were performed to probe
6341: the anisotropic component in a systematic way by
6342: exploiting the SO($3$) decomposition
6343: \cite{ara99,bif01a,bif02a,bif02,bif03a}.
6344: Here we review the main findings, showing that
6345: \begin{enumerate}
6346: \item The scaling laws (log-log plots) at moderate Reynolds numbers are
6347: significantly improved by projecting the raw correlation functions onto
6348: each $j$-sectors. The improvement is particular noticeable whenever strong
6349: anisotropies are present in
6350: the system, as in the case of channel flows \cite{ara99,bif02a};
6351: \item Anisotropic sectors with $j \ge 2$ (inaccessible in present
6352: experimental data)
6353: possess good scaling laws \cite{bif01a,bif02};
6354: \item The scaling exponents are discrete and increasing as a
6355: function of $j$.
6356: \item The exponents are {\it anomalous}; i.e. they differ from the
6357: dimensional
6358: prediction (\ref{lumleygen}).
6359: \item There exists
6360: preliminary evidence that also for $j > 2$, the anomalous exponents
6361: are {\it universal}, i.e. the scaling properties are independent of the
6362: external forcing mechanism \cite{bif03a}.
6363: \end{enumerate}
6364: DNS were performed both in wall-bounded flows and in homogeneous (but
6365: anisotropic)
6366: turbulence. In wall-bounded flows the anisotropies are accompanied by {\it
6367: inhomogeneous} effects. The presence of such effects may
6368: spoil the very meaning of scaling, and the SO($3$) decomposition
6369: should be supplemented by some tool to project on the homogeneous
6370: components. Otherwise, the SO(3) decomposition must be used
6371: carefully, and locally, only in those regions of the tested flow
6372: where inhomogeneous effects are confined mostly to
6373: large-scales \cite{ara99,bif02a}. In the second part of
6374: this section, we discuss numerical experiments
6375: built such as to have a perfectly {\it homogeneous} and {\it
6376: anisotropic} statistics at all scales. One such example is {\it
6377: homogeneous shear flows} \cite{sch00,cas00}. More recently,
6378: other homogeneous anisotropic flows have been invented and simulated,
6379: in particular the random-Kolmogorov-flow
6380: \cite{bif01a,bif02,bif03b} and a convective cell with an imposed linear
6381: mean profile of
6382: temperature \cite{bif03a}.
6383: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6384: \subsection{Anisotropic and inhomogeneous statistics: Channel flows}
6385: \label{sec:num_channel}
6386: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6387: In this section, we discuss the analysis of a DNS of a channel flow
6388: using the SO($3$) decomposition. The coordinates are chosen
6389: such that $\hat{x}$, $\hat{y}$ and $\hat{z}$ are the stream-wise,
6390: span-wise,
6391: and wall-normal direction respectively. The simulation was done on a
6392: grid with $256$ points in the stream-wise direction and $128\times128$
6393: points in the two other directions. The boundary conditions were
6394: periodic in the span-wise and stream-wise directions and no-slip on the
6395: walls. The Reynolds-number based on the Taylor micro-scale was quite
6396: moderate,
6397: $R_{\lambda}\approx 70$ at the center of the channel $(z=64)$. The
6398: simulation was fully symmetric with respect to the central plane. For
6399: more details on the averaged quantities and on the numerical code, see
6400: refs.~\cite{ara99,ama97,ama97a}.
6401:
6402: The analysis focused on longitudinal second-, forth- and sixth-order
6403: structure functions:
6404: $$
6405: S^{(n)}(\Br^c,\Br) \equiv \left<\big[\delta
6406: u_\ell(\Br^c,\Br)\big]^n\right>, \quad \delta u_\ell(\Br^c,\Br) \equiv
6407: \hat{\Br}\cdot \big[\Bu(\Br^c+\Br,t)-\Bu(\Br^c-\Br,t) \big] \ .
6408: $$
6409: The $\Br^c$ coordinate specifies the location of the measurement
6410: (i.e., the center of mass of the two measurement points), and $2\Br$
6411: is the separation vector. Previous analysis of the same data-base
6412: \cite{ama97} as well as of other DNS \cite{bor96} and
6413: experimental data \cite{sad94,gau98} in anisotropic flows
6414: found that the scaling exponents of energy spectra, energy co-spectra
6415: and of longitudinal structure functions exhibit strong dependence on
6416: the position $r_c$. For example, in \cite{tos99}
6417: the authors studied the longitudinal structure functions at fixed
6418: distances from the walls: $$S^{(n)}(r,z)\equiv \la
6419: (u_x(x+r,y,z)-u_x(x,y,z))^n\ra_z$$ where $\la \cdots \ra_z$ denotes a
6420: spatial average on a plane at a fixed height $z$, $1<z<64$. For this
6421: set of observables they found that: (i) These structure functions did
6422: not exhibit clear scaling behavior as a function of the distance
6423: $r$. Consequently, one needed to resort to Extended-Self-Similarity
6424: (ESS) \cite{ben93b} in order to extract a set of {\rm relative}
6425: scaling exponents $ \hat{\zeta}^{(n)}(z) \equiv
6426: \zeta^{(n)}(z)/\zeta^{(3)}(z)$;
6427: (ii) the relative exponents, $ \hat{\zeta}^{(n)}(z)$ depended strongly on
6428: the height $z$. Moreover, only at the center of the channel and very
6429: close to the walls the error bars on the relative scaling exponents
6430: extracted by using ESS were small enough to claim the very existence
6431: of scaling behavior in any sense. Similarly, an experimental analysis
6432: of a turbulent flow behind a cylinder \cite{gau98} showed a
6433: strong dependence of the relative scaling exponents on the position
6434: behind the cylinder for not too big distances from the obstacle,
6435: i.e. where anisotropic effects may still be relevant in a wide range
6436: of scales. In the following we present an interpretation of the
6437: variations in the scaling exponents observed in non-isotropic and
6438: non-homogeneous flows upon changing the position in which the analysis
6439: is performed. In particular, we will show that decomposing the
6440: statistical objects into their different $(j,m)$ sectors rationalizes
6441: the findings, i.e. scaling exponents in given $(j,m)$ sector appear
6442: quite independent of the spatial location; only the {\em amplitudes}
6443: of the SO(3) decomposition depend strongly on the spatial location.
6444: The analysis showed three major results. The first was the vast
6445: improvement in scaling behavior of the structure functions as a result
6446: of the decomposition. A typical example is found in
6447: Fig.(\ref{fig:channel.fig1} )
6448: \begin{figure}[!h]
6449: \includegraphics[scale=0.9]{fig1.so3.eps}
6450: \caption{Log-log plot of the isotropic sector of the 4th order structure
6451: function $S^{(4)}_{0,0}(r)$, vs. $r$ at the center of the channel
6452: $r^c_z=64$ (+). The data represented by ($\times$) correspond to the
6453: raw longitudinal structure function, $S^{(4)}(r^c_z=64,r\hat x)$
6454: averaged over the central plane only. The dashed line corresponds to
6455: the intermittent isotropic high-Reynolds numbers exponents
6456: $\zeta^{(4)}_0 = 1.28$.}
6457: \label{fig:channel.fig1}
6458: \end{figure}
6459: where the raw fourth-order
6460: structure function, evaluated on the central plane, is compared to its
6461: $j=0$ component. Without the SO($3$) decomposition there is no
6462: scaling behavior at all and one needs ESS
6463: to estimate the scaling exponents. On the other hand, the
6464: $j=0$ component of the structure function shows a clear scaling
6465: behavior with the expected exponent, $\zeta^{(4)}_{0} = 1.28$. This
6466: strengthens the foliation hypothesis, according to which, the raw
6467: structure function is a superposition of power laws from different
6468: sectors of the SO($3$) group. Such a sum looses its scale invariance
6469: once the weights of the different exponents are of the same order and
6470: the inertial range is small. In such cases, one needs the SO($3$)
6471: decomposition to isolate the different sectors and retain scale
6472: invariance.
6473: \begin{figure}[!h]
6474: \includegraphics[scale=0.9]{fig3.so3.eps}
6475: \caption{
6476: Logarithmic local slopes
6477: of the ESS plot of raw structure function,$\frac{d log(S^{(4)}(r,r_c)}{d
6478: log(S^{(2)}(r,r_c)}$,
6479: of order 4 versus raw structure function of order 2
6480: at $r_c^z=64$ ($\times$), at $r_c^z=32$ ($\star$) and of the $j=0$
6481: projection, $\frac{d log(S^{(4)}_{00}(r,r_c)}{d log(S^{(2)}_{00}(r,r_c)}$,
6482: centered at $r_c^z=64$ ($+$), and at $r_c^z=32 $($\Box$).
6483: Also two horizontal lines
6484: corresponding to the high-Reynolds number limit, $1.82$, and to the
6485: K41 non-intermittent value, $2$, are shown.}
6486: \label{fig:channel.fig3}
6487: \end{figure}
6488:
6489: A second prominent result is the apparent universality of the
6490: isotropic exponents. To show this in \cite{ara99} the local slopes,
6491: $\frac{d log(S^{(4)}_{00}(r)}{d log(S^{(2)}_{00}(r)}$,
6492: of the ESS curves of the isotropic forth-order structure-function versus
6493: the
6494: isotropic second-order structure function were calculated at varying
6495: the
6496: distance from the wall.
6497: Despite their different locations, all curves show the same ESS slope
6498: 1.82, which is the expected (anomalous) value. In
6499: Fig.(\ref{fig:channel.fig3}) one picture is presented for the
6500: logarithmic local slopes at two different distances from the
6501: channel boundary. To appreciate the improvements in scaling
6502: and universality, also the slopes of the ESS on the raw structure
6503: functions are presented
6504: Finally, the analysis provided another evidence that the $j=2$
6505: scaling exponent of the second order structure function is about
6506: $4/3$, which is the dimensional theoretical prediction given in
6507: (\ref{eq:Lumley})
6508: (see also \cite{lum67,yak94,fal95,gro94,bif02}).
6509: Considering the relatively low
6510: Reynolds-number and the fact that the prefactors $a_{j,m}$ in the
6511: SO($3$) decomposition (\ref{eq:SnDecomp}) are non-universal, together
6512: with the experimental result reported in
6513: \cite{ara98,kur00,she02,she02a}, these findings give
6514: strong support to the view that the scaling exponent in the $j=2$
6515: sector is universal.
6516: Before concluding this section we cite that
6517: SO(3) and SO(2) decomposition have also been exploited in the analysis
6518: of channel flow data to highlights the importance of structures as streaks
6519: and
6520: hairpin filaments typical of many wall bounded flow \cite{bif02a}.
6521: Preliminary investigation of the importance of SO(3) decomposition
6522: to evaluate the performance of sub grid models used in Large Eddy
6523: Simulations
6524: \cite{mene00}
6525: have also been reported in \cite{bif03b}.
6526: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6527: \subsection{Anisotropic-homogeneous flows}
6528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6529: \label{sec:anis-hom}
6530: Direct numerical simulations offer the unique opportunity to
6531: study the physics of anisotropy in {\it ideal} situations, that is in
6532: perfectly homogeneous flows. Recently, considerable effort has been spent
6533: on simulating a Random-Kolmogorov-Flow (RKF)
6534: \cite{bif01a,bif02,bif03b}. The RKF is
6535: fully periodic, incompressible and with anisotropic large-scale energy
6536: injection. A convenient choice for the forcing is ${\f}=(0,0,f_z(x))$
6537: with $f_z(x)=F_1 \cos[2\pi x/L_x +\phi_1(t)] + F_2
6538: \cos[4 \pi x/ L_x +\phi_2(t)]$, with constant amplitudes $F_{1,2}$ and
6539: independent, uniformly distributed, $\delta$-correlated in time and with
6540: random
6541: phases $\phi_{1,2}(t)$. The random phases
6542: lead to a homogeneous statistics. To give a first validation of the
6543: statistical properties of the RKF flow we plot in
6544: fig.~\ref{fig:RKF.spectra}
6545: %%%%%%%%%%%%%%%%%%%%%%%
6546: \begin{figure}[h]
6547: \includegraphics[scale=0.9]{RKF.spectra.tex.eps}
6548: \caption{Log-log plot of instantaneous energy spectrum in the
6549: isotropic sector $E(k)$ (top). The straight line is the reference isotropic
6550: $k^{-5/3}$ power law. Instantaneous co-spectrum $E_{yz}(k_z)$
6551: (bottom). Here the straight line gives the reference $k_z^{-7/3}$
6552: anisotropic Lumley prediction. The two spectra have been shifted along
6553: the vertical direction for the sake of presentation.}
6554: \label{fig:RKF.spectra}
6555: \end{figure}
6556: %%%%%%%%%%%%%%%%%%%%%%%%%%
6557: the instantaneous energy spectrum,
6558: $$E(k)=\int_{|\Bq|=k} \la \Bu(\Bq)\cdot \Bu^*(\Bq) \ra d\Bq.$$ It
6559: exhibits a scaling law in close agreement with the K41 isotropic behavior
6560: $k^{-5/3}$. Also purely anisotropic quantities as the co-spectra
6561: (\ref{cospec}),
6562: show a good agreement with the Lumley $k^{-7/3}$.
6563: DNS of the RKF were reported in \cite{bif01a,bif02,bif03b}. The
6564: resolution was $256^3$ reaching $Re_{\lambda}\sim 100$,
6565: collecting up to $70$ eddy
6566: turn over times. A long time average is necessary because of the
6567: formation of persistent
6568: large-scale structures inducing strong oscillations of the mean
6569: energy evolution. This is typical to many strongly anisotropic flows.
6570: The viscous term was replaced by a
6571: second-order hyper-viscosity, $-\nu \Delta^2{\Bu}$.
6572: Thanks to both the high degree of homogeneity and to
6573: the high number of independent samples, a quantitative analysis of
6574: scaling laws of longitudinal structure functions up to the anisotropic
6575: sector $j=6$ and up to order $n=6$ was possible. In other words, the
6576: longitudinal structure functions could be decomposed according to
6577: $$
6578: S^{(n)}(\Br) = \sum_{j=0}^{6}\sum_{m=-j}^{j} S^{(n)}_{jm}(r)
6579: Y_{jm}(\unitr ) \quad \rm{for }~n\le 6 .
6580: $$
6581: In fig.~\ref{fig:RKF.fig1} we present the results for the
6582: isotropic sector. Here we compare the raw structure
6583: functions in the three directions with the projection
6584: $S^{(2)}_{00}(r)$, and their logarithmic local slopes (inset).
6585: \begin{figure}[!h]
6586: \includegraphics[scale=0.9]{RKF.fig1.eps}
6587: \caption{ Analysis on
6588: the real space. Log-log plot of $S^{(2)}_{00}(r)$ versus $r$ (top
6589: curve), and of the
6590: three undecomposed longitudinal structure functions in the three
6591: directions $x,y,z$ (three bottom curves). The
6592: straight line gives the best fit slope $\zeta_{0}^{(2)}=0.7$. Inset:
6593: logarithmic
6594: local slopes of the same curves in the main body of the figure (same
6595: symbols). Notice that only the projected curve shows a nice plateau }
6596: \label{fig:RKF.fig1}
6597: \end{figure}
6598: Only for the projected correlation it is possible to measure (with
6599: $5\%$ of accuracy) the scaling exponent by a direct log-log fit versus
6600: the scale separation. The best fit gives $\zeta_{0}^{(2)} = 0.70 \pm
6601: 0.03$. On the contrary, the undecomposed structure functions are
6602: overwhelmed by the anisotropic effects present at all scales, and the
6603: scaling law is completely spoiled. We stress the accuracy
6604: of these results; already at these modest Reynolds numbers it is
6605: possible to ascertain the isotropic scaling laws if the anisotropic
6606: fluctuations are disentangled properly.
6607:
6608: In figure \ref{fig:RKF.fig3} there is an overview for the second order
6609: structure functions in all the sectors, isotropic and anisotropic, for which
6610: the signal-to-noise ratio is high enough to ensure statistically
6611: stable results. Sectors with odd $j$ are absent due to the
6612: parity symmetry of the longitudinal structure function. We conclude from
6613: figure \ref{fig:RKF.fig3} a clear foliation in terms of the $j$
6614: index\,: sectors with the same $j$ but different $m$ exhibit very
6615: close scaling exponents.
6616: In Table~3 the measured exponents are compiled,
6617: showing the best power law fits for
6618: structure functions of orders $n=2,4,6$.
6619: \begin{figure}[!h]
6620: \includegraphics[scale=0.9]{RKF.fig3.eps}
6621: \caption{Log-log plot of the second order structure function in all
6622: sectors with a strong signal. Symbols refer to sectors (j,m) as follows:
6623: $(0,0)$, ($+$); $(2,2)$,
6624: ($\times$); $(4,0)$, ($\Box$); $(4,2)$, ($\star$); $(6,0)$, ($\circ$);
6625: $(6,2)$, ($\blacksquare$). The statistical and numerical noise affecting
6626: the SO(3) projection is estimated as the threshold where the $j=6$
6627: sector starts to deviate from the monotonic decreasing behavior,
6628: i.e. ${\mathcal O}(10^{-3})$}
6629: \label{fig:RKF.fig3}
6630: \end{figure}
6631: We stress again the discreteness and monotonicity of the scaling
6632: exponents as assumed in Eq.~(\ref{eq:SnDecomp});
6633: there is no saturation of the exponents as a function of
6634: $j$. Second, the measured exponents in the sectors $j=4$
6635: and $j=6$ are anomalous, i.e. they differ from the dimensional
6636: estimate given in \ref{lumleygen}.
6637: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6638: \begin{table}[!h]
6639: \label{tab:compile}
6640: \begin{center}
6641: \begin{tabular}{|c|c|c|c|c|}
6642: \hline
6643: $n$ & $j=0$ & $j=2$ & $j=4$ & $j=6 $\,\\
6644: \hline
6645: & $\zeta_{0}^{(n)}$ --- $n/3 $ \,& $\zeta_{2}^{(n)}$ ---
6646: $ (n+2)/3 $\, & $\zeta_{4}^{(n)}$ --- $ (n+4)/3 $\, &
6647: $\zeta_{6}^{(n)}$ --- $ (n+6)/3 $\,\\
6648: \hline
6649: 2 & 0.70 (2) --- 0.66 & 1.1 (1) --- 1.33 & 1.65 (5) --- 2.00 & 3.2 (2)
6650: --- 2.66 \\
6651: 4 & 1.28 (4) --- 1.33 & 1.6 (1) --- 2.00 & 2.25 (10) --- 2.66
6652: & 3.1 (2) --- 3.33 \\
6653: 6 & 1.81 (6) --- 2.00 & 2.1 (1) --- 2.33 & 2.50
6654: (10) --- 3.33 & 3.3 (2) --- 4.00 \\
6655: \hline \end{tabular}
6656: \caption{Summary of the numerical and experimental
6657: findings for the scaling exponents in the isotropic and anisotropic
6658: sectors. The values for the anisotropic sector $j=2$ are taken from
6659: the experiments {\protect \cite{kur00,she02a}}. For the values
6660: extracted from the numerical simulation (columns $j=0,4,6$), error
6661: bars are estimated from the oscillation of the local slopes {\protect
6662: \cite{bif01a,bif03b}}. For the experimental data the error is
6663: given as the mismatch between the two experiments. For all sectors we
6664: also give the dimensional estimate $\zeta_{j}^{(n)} = (n+j)/3$
6665: {\protect \cite{bif02}}.}
6666: \end{center}
6667: \end{table}
6668: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% tabella %%%%%%%%%%%%%%%%%
6669: Unfortunately, from the RKF data it was not possible to obtain clean
6670: results for the $j=2$ sector. This is because of the presence of
6671: an annoying change of sign in the projections $S^{(n)}_{2m}(r)$ for
6672: any $m$ (and any order $n$). Still, the overall consistency of the
6673: foliation and hierarchical organization of scaling exponents can be
6674: checked by collecting the scaling exponents in the $j=2$ sector from
6675: the two sets of experiments
6676: \cite{kur00,she02a} previously discussed.
6677: In Fig.~\ref{fig:RKF+EXP} we show both numerical data and the
6678: experimental values as extracted from
6679: \cite{kur00,she02a}.
6680: The resulting picture
6681: is fully coherent\,: experimental data coming from the $j=2$ sector
6682: fit well in the global trend. As one can see from table~3
6683: all the anisotropic sectors show {\it anomalous} scaling laws.
6684: \begin{figure}[!h]
6685: \includegraphics[scale=0.9]{RKF.fig4.eps}
6686: \caption{Scaling exponents, $\zeta_{j}^{(n)}$, of
6687: structure functions of order $n=2,4,6$ for isotropic and anisotropic
6688: sectors. From the DNS of RKF we have\,: isotropic sector, $j=0$
6689: ($+$); anisotropic sectors, $j=4$ ($\times$) and $j=6$
6690: ($\star$). From the experimental data {\protect \cite{kur00,she02a}},
6691: we have $j=2$ ($\Box$). For an estimate
6692: of error bars see Table 3.
6693: %%%%% The straight lines
6694: %correspond to the dimensional prediction (\ref{lumleygen})
6695: }
6696: \label{fig:RKF+EXP}
6697: \end{figure}
6698: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6699: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6700: \subsubsection{Universality of Anisotropic Fluctuations}
6701: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6702: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6703: The
6704: third numerical experiment that we discuss here is devoted to study {\it
6705: universal}
6706: properties of anisotropic scaling. We have already commented that
6707: there is a nice qualitative and quantitative agreement between the
6708: values extracted for the $j=2$ sector (the only one available from
6709: experimental data) from different experiments. In order to check
6710: whether this universality holds also in higher anisotropic sectors one
6711: has to rely on DNS. In \cite{bif03a} a first direct comparison
6712: between anisotropic scaling of longitudinal structure functions from
6713: two different homogeneous systems, the RKF and a {\it homogeneous \rb}
6714: convective flow was reported.
6715:
6716: A Homogeneous \rb system is a convective cell with fixed
6717: linear mean temperature profile along the vertical direction. The
6718: flow is obtained by decomposing the temperature field as the sum of a
6719: linear profile plus a fluctuating part, $T(x,y,z;t)= T'(x,y,z;t) + (\Delta
6720: T/2 - z \Delta T/H) $, where $H$ is the cell height and $\Delta T$
6721: the background temperature difference. The evolution of the system can
6722: be described by a modified version \cite{gro02} of the Boussinesq system
6723: \cite{kad01}:
6724: \bea
6725: \partial_t \Bu &+& \lp\Bu\cdot\nabla\rp \Bu =
6726: -\nabla p + \nu \nabla^2 \Bu + \alpha g T' {\unitz} \nonumber \\
6727: \partial_t T' &+& \lp\Bu\cdot\nabla\rp T' =
6728: \kappa \nabla^2 T' - {{\Delta T} \over H} v_z. \nonumber
6729: \eea
6730: %MODIFICATO QUI
6731: where $\alpha$ is the thermal expansion constant,
6732: $\nu$ and $\kappa$ the kinematic
6733: viscosity and the thermal diffusivity coefficients,
6734: and $g$ is the acceleration due to gravity.
6735: In \cite{bif03a}
6736: fully periodic boundary conditions were used for the velocity field,
6737: $\Bu$, and temperature, $T'$, fields.
6738:
6739: Anisotropic effects in the \rb system were analyzed in \cite{bif03a}
6740: starting from the stationary equation for the second order
6741: velocity structure functions; the extension of K\'arm\'an-Howarth
6742: equation in the presence of a buoyancy term \cite{yak92}. The result
6743: is, neglecting for simplicity tensorial symbols:
6744: \bea
6745: \label{general2}
6746: \la \delta u(\r)^3 \ra &\sim& \ \ \eb \ r \ + \ \alpha g\unitz r\
6747: \cdot \ \la \delta T(\r)\ \delta u(\r)\ra\\ _{j = 0,1,\dots} &\ & \ \
6748: _{j=0} \quad \ \ _{j=1 \quad \otimes \quad j=1,2,\dots} \nonumber \eea
6749: where $\bar \epsilon$ denotes the energy dissipation, $\la
6750: \delta u(\r)^3 \ra$ and $\la \delta T(\r)\ \delta u(\r)\ra$, the
6751: general third-order velocity correlation and temperature-velocity
6752: correlation, respectively. In Eq.~(\ref{general2}) for each term the
6753: value of its total angular momentum, $j$, is indicated. Notice that
6754: the
6755: energy dissipation term in (\ref{general2}) has a non-vanishing limit,
6756: for high Re, only in the isotropic sector, $j=0$. On the
6757: other hand, the buoyancy coupling, $\alpha g\unitz$, brings only
6758: angular momentum $j=1$. Due to the usual rule of composition of angular
6759: momenta we have that the buoyancy term, $\alpha g \unitz \cdot \la
6760: \delta T(\r)\ \delta u(\r)\ra$, has a {\it total} angular momentum
6761: given by: $j_{tot}= 1 \otimes j = \{j-1,j,j+1\}$. Using the
6762: angular momenta summation rule for $j$, one can decompose the previous
6763: equation obtaining the following dimensional matching, in the
6764: isotropic sector: $$
6765: \la \delta u(r)^3 \ra_{j=0} \sim \epsilon\ r +
6766: \alpha g \hat{{\bf z}} r \la \delta u(r) \ \delta
6767: T(r)\ra_{j=1} + \dots $$
6768: and in the anisotropic sectors, $j >0$:
6769: \be
6770: \la \delta u(r)^3 \ra_{j} \sim \ \ \alpha g\ \hat{{\bf z}}
6771: r\ \la \delta u(r) \ \delta T(r) \ra_{(j-1)} + \dots
6772: \label{aniso}
6773: \ee
6774: where sub-dominant contributions coming from the $j$ and $j+1$ sectors
6775: of $\la \delta v(r) \ \delta T(r) \ra$ are neglected.
6776:
6777: In the isotropic sector the buoyancy term is sub-dominant with respect
6778: to the dissipation term at scales smaller than the Bolgiano length,
6779: $L_B=(\bar \epsilon)^{5/4}N^{-3/4}(\alpha g)^{-3/2}$ where $N$ is the rate
6780: of temperature dissipation.
6781: This is the case for the numerical simulation presented in
6782: \cite{bif03a}, where velocity fluctuations are closer to the typical
6783: Kolmogorov scaling, $\delta u(r) \sim r^{1/3}$, rather than to the
6784: Bolgiano-Obukhov scaling \cite{my75},
6785: $\delta u(r) \sim r^{3/5}$.\\ Regarding the
6786: anisotropic sectors, Eq.(\ref{aniso})
6787: is the {\it dimensional prediction} for the system,
6788: consistent with the anisotropic properties of the buoyancy term,
6789: sector by sector. \\ In \cite{bif03a} the SO(3)
6790: decomposition was applied in this system
6791: to velocity structure functions (\ref{eq:fundamental}) and to objects
6792: $$G^{(q,1)}(\r) =
6793: \la\left[\lp\Bu\lp\r\rp-\Bu\lp 0 \rp\rp\cdot \unitr\right]^{q}\lp T\lp\r
6794: \rp-T\lp 0 \rp\rp \ra\, = \sum_{jm} G^{(q,1)}_{jm}(r)$$
6795:
6796: The dimensional matching of Eq.~(\ref{aniso}) can be extended to any
6797: order, giving:
6798: $$
6799: S^{(p)}_{jm}(r) \sim r { G}^{(p-2,1)}_{j-1,m}(r).
6800: $$
6801: Denoting with $\chi_j^{(q,1)}$ the anisotropic scaling exponents of
6802: the buoyancy terms, ${G}^{(q,1)}_{jm}(r) \sim r^{\chi_j^{(q,1)}}$
6803: we
6804: get the dimensional estimate:
6805: \be
6806: \zeta_j^{(p)} = 1 + \chi_{j-1}^{(p-2,1)}, \quad \mbox{ (dimensional
6807: prediction)}
6808: \label{dim_exp}
6809: \ee
6810: In \cite{bif03a} it was shown that
6811: this dimensional prediction is not obeyed; the exponents $\zeta_j^{(n)}$
6812: appear to be systematically smaller than the prediction (\ref{dim_exp}).
6813: Interestingly, enough the log-log plots computed in the sectors
6814: $j=4,6$ show a good
6815: qualitative agreement with those calculated in the RKF
6816: of ref \cite{bif01a} as can be seen in
6817: fig.~\ref{fig:RKF+HRB.1} where we compare the
6818: projection on the $j=4$ sector of structure
6819: functions of different orders.
6820: \begin{figure}[h]
6821: \includegraphics[scale=0.9]{HRB+RKF.eps}
6822: \caption{ Log-log plot of compensated anisotropic $j=4,m=0$ projections
6823: $S^{(p)}_{4,0}(r)/r^{\zeta_4^{(p)}}$ {\it vs} $r$, for both HRB and RKF
6824: flows. Top curves refer to $p=2$: the best fit exponents which
6825: compensate HRB and RKF curves are $\zeta_4^{(2)}=1.7$ and
6826: $\zeta_4^{(2)}=1.66$, respectively. Curves in the middle refer to the
6827: same quantities but for $p=4$: compensation has been obtained with
6828: $\zeta_4^{(4)}=2.05$ for HRB, and $\zeta_4^{(4)}=2.2$ for RKF. Bottom
6829: curves
6830: refer to $p=6$: here $\zeta_4^{(6)}=2.3$ for HRB, and $\zeta_4^{(6)}=2.5$
6831: for
6832: RKF. Notice that the curves of the two flows are compensated with
6833: very similar values of the exponents (within $10\%$). Inset: the
6834: same but for $j=4,m=2$, compensation has been done with the same
6835: values used for $j=4,m=0$, to show the independence of the scaling
6836: exponents from the choice of the reference axis labeled with $m$ .
6837: %Log-log plot of $S^{(p)}_{4,0}(r)$
6838: %{\it vs} $r$ for both the Homogeneous \rb and the Random Kolmogorov Flow.
6839: % The curves refer to the orders $p=3,4,5,6$ (top to bottom).
6840: % Inset: same for $j=6$.
6841: }
6842: \label{fig:RKF+HRB.1}
6843: \end{figure}
6844: Similar results are obtained for $j=6$
6845: sector. These preliminary findings, if confirmed by other
6846: independent measurements, would support {\it universality}
6847: for anisotropic scaling exponents in three dimensional turbulence.
6848: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6849: \subsection{Scaling of Longitudinal and Transversal structure functions}
6850: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6851: As discussed in
6852: Sect.~\ref{sec:long-tran}
6853: there exists experimental and numerical data suggesting that
6854: {\it longitudinal} and {\it transversal } structure functions
6855: in supposedly isotropic flows show
6856: different scaling exponents \cite{she02,che97,dhr97,got02,wat99}.
6857: One needs to clearly
6858: distinguish between experimental and numerical data. The former can
6859: never be considered fully isotropic; the best one can do is to try to
6860: perform a multi-fit procedure to clean out sub-leading anisotropic
6861: contributions as already explained in details in Sect.
6862: \ref{chap:experiment}. This fitting procedure is, of
6863: course, affected by experimental errors which cannot be eliminated.
6864: Therefore it
6865: is quite dangerous to make any firm conclusion about
6866: supposed different scaling exponents of longitudinal and transversal
6867: isotropic scaling on the basis of only experimental data .
6868: Numerical data are not much safer. Here anisotropy can be much better
6869: controlled.
6870: With isotropic forcing the only source of
6871: anisotropy is the
6872: 3-dimensional grid whose effect is usually too small
6873: to explain
6874: possible discrepancies between longitudinal and transversal
6875: scalings. Indeed some state-of-the-art isotropic
6876: DNS indicate the possibility
6877: of different scaling exponents both for inertial range structure functions
6878: \cite{got02} and for coarse grained energy and enstrophy measures
6879: \cite{che97,che97b}. In Table~4 we
6880: summarize the best-fit values of the scaling exponents measured in
6881: \cite{got02}.
6882: %\begin{figure}[h]
6883: %\scalebox{1.9}{
6884: %\includegraphics[scale=0.6]{ls.long.toshi.ps}}
6885: %\caption{FIGURE of the longitudinal local scaling
6886: %exponents, from gotoh}
6887: %\label{fig:toshi}
6888: %\end{figure}
6889: %\begin{figure}[h]
6890: %\scalebox{1.9}{
6891: %\includegraphics[scale=0.6]{ls.tra.toshi.ps}}
6892: %\caption{FIGURE of the transversal local scaling
6893: %exponents, from gotoh}
6894: %\label{fig:toshib}
6895: %\end{figure}
6896: The small scale fluctuations were probed \cite{che97b}
6897: by comparing the
6898: scaling of the coarse grained energy dissipation
6899: over a box of size $r$, $\tilde \epsilon (r)$ (Eq.~\ref{eq:coarse}),
6900: and of the coarse grained enstrophy dissipation:
6901: $\omega(r,\Bx) = {1\over r^3}
6902: \int_{|y| < r} d\By ~\omega(\Bx +\By)$
6903: where $\omega(\Bx)$ is the local enstrophy dissipation. Different
6904: scaling exponents were measured for the averaged quantities, $\la
6905: (\tilde \epsilon)^p(r)\ra$, $\la\omega^p(r)\ra$. Being scalar
6906: quantities one expects that in isotropic ensembles they would not have
6907: different exponents. From the theoretical point of view, different scaling
6908: exponents of
6909: longitudinal and transverse structure functions in isotropic ensembles
6910: are unlikely. In the language of the SO(3) decomposition it amounts
6911: to the scaling exponents depending on the $q$-index which labels
6912: different basis functions with the same rotation properties. In the
6913: exactly solvable models examined before, this had never happened. In
6914: general one would need a different symmetry to lift the degeneracy of
6915: different $q$ dependent basis functions. At this point this problem
6916: remains somehow unsettled. New numerical tests on larger grids and/or
6917: with a better resolved viscous behavior are needed before a firm
6918: statement can be made.
6919: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6920: \begin{table}[h]
6921: \begin{tabular}{|c|c|c|c|c|c|}
6922: \hline
6923: n&2&4&6&8&10 \\
6924: \hline
6925: $\zeta^{(n)}_0$ &0.701 (14)& 1.29 (3)& 1.77 (4)& 2.17 (7)& 2.53 (9) \\
6926: \hline
6927: $\zeta^{(n)}_0$ & 0.709 (13) &1.27 (2)& 1.67 (4) & 1.93 (9) & 2.08 (18)\\
6928: \hline
6929: \end{tabular}
6930: \label{table:gotoh}
6931: \caption{Measured values of
6932: the longitudinal (first raw) and transverse (second raw)
6933: scaling exponents
6934: at $Re_{\lambda}=460$ taken from \cite{got02}. One should note that
6935: the scaling range displayed by the scaling plots in \cite{got02} are
6936: relatively short, indicating that finite Re effects may still
6937: be rather important}
6938: \end{table}
6939: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6940: \subsection{Anisotropies in decaying turbulence}
6941: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6942: Decaying turbulence has attracted the attention of various communities
6943: and is often considered in experimental, numerical and theoretical
6944: investigations \cite{bat53,fri95,my75}. It is in fact quite common
6945: that even experiments aimed at studying stationary properties of
6946: turbulence involve processes of decay. Important examples are provided
6947: by a turbulent flow behind a grid (see \cite{skr00} and references therein)
6948: or the turbulent flow created
6949: at the sudden stop of a grid periodically oscillating within a bounded
6950: box \cite{des94}. In the former case, turbulence is slowly decaying
6951: going farther and farther away from the grid and its characteristic
6952: scale becomes larger and larger (see \cite{skr00} for a thorough
6953: experimental investigation). Whenever there is sufficient separation
6954: between the grid-size $L_{in}$ and the scale of the tunnel or the tank
6955: $L_{0} \gg L_{in}$, a series of interesting phenomenological
6956: predictions can be derived. For example, the decay of the two-point
6957: velocity correlation function, for both isotropic and anisotropic
6958: flows, can be obtained under the so-called self-preservation
6959: hypothesis (see \cite{my75} chapter XVI). That posits the existence of
6960: rescaling functions allowing to relate correlation functions at
6961: different spatial and temporal scales. By inserting the
6962: rescaling function
6963: into the equations of motion, asymptotic results can be obtained both
6964: for the final viscosity-dominated regime and for the intermediate
6965: asymptotic when nonlinear effects still play an important role.
6966:
6967: Here, we review some recent attempts to investigate the decay of
6968: three-dimensional homogeneous and anisotropic turbulence by direct
6969: numerical simulations of the Navier-Stokes equations in a periodic box
6970: \cite{bif03c} for both short and large times. The initial
6971: conditions are taken from the stationary ensemble of the Random
6972: Kolmogorov Flow discussed in the previous subsection. Here the
6973: correlation length-scale of the initial velocity field $L_{in}$ is of
6974: the order of the size of the box $L_{0}\approx L_{in}$.
6975:
6976: On the one hand, one is interested in the
6977: long time decay regime where the typical interesting questions are: (i) how
6978: do global quantities, such as single-point velocity and vorticity
6979: correlations, decay? (ii) What is the effect of the outer boundary on
6980: the decay laws~? (iii) Do those quantities keep track of the initial
6981: anisotropy~? (iv) As for the statistics of velocity differences within
6982: the inertial range of scales, is there a recovery of isotropy at large
6983: times~? (v) If so, do strong fluctuations get isotropic at a
6984: faster/slower rate with respect to those of average intensity~? (vi)
6985: Do anisotropic -and isotropic- fluctuations decay self-similarly~?
6986: (vii)If not, do strong fluctuations decay slower or faster than
6987: typical ones~? \
6988: On the other hand, the
6989: interest in the early stages of the decay is led by a hope of
6990: establishing a link between the small-scale velocity statistics in
6991: this phase and in forced turbulence. If such links existed, they would
6992: shed additional light on the universality of forced
6993: turbulence.
6994: As turbulence decays, the effective Re
6995: decreases, while the viscous characteristic scale and time
6996: increase. In \cite{bif03c} an offline analysis at fixed
6997: multiples $\{0,1,10,10^{2},10^{3},10^4,10^5,10^6\}\,\tau_0$ of the
6998: initial large-scale eddy turnover time $\tau_0 = L_{0}/u_{rms}^{t=0}$
6999: was performed. \\ A first hint on the restoration of isotropy at large
7000: times can be extracted from the analysis of single point quantities as
7001: $$ E_{il} = {\overline{u_i(\Bx, t)u_l(\Bx, t)}},\quad
7002: \Omega_{il} = {\overline{\omega_i(\Bx, t) \omega_l(\Bx, t)}}. $$
7003: Here with $\overline{\cdots}$ we denote the average over spatial
7004: coordinates only, whereas $\la \cdots \ra$ indicates the
7005: average over both initial conditions and space. The symmetric
7006: matrices $E_{il}(t)$ and $\Omega_{il}(t)$ can be diagonalized at
7007: each time-step and the eigenvalues $E_1(t),E_2(t),E_3(t)$ and
7008: $\Omega_1(t),\Omega_2(t),\Omega_3(t)$ can be extracted. The typical
7009: decay of $E_{i}(t)$ and $\Omega_{i}(t)$ for $i=1,\dots,3$ is
7010: shown in Fig.~\ref{fig:2}.
7011: \begin{figure}[h]
7012: %\epsfxsize=8.0truecm
7013: %\epsfysize=6.0 cm
7014: \epsfbox{DEC.fig2.eps}
7015: \vspace{4pt}
7016: \caption{Log-log plot of the eigenvalues of energy and vorticity matrices
7017: {\it vs.} time, expressed in $\tau_0$ unit.}
7018: \label{fig:2}
7019: \end{figure}
7020: During the self-similar stage, $t\in
7021: [10,10^6]$, the energy eigenvalues fall off as $E_{\{1,2,3\}}\sim
7022: t^{-2}$, as expected for the decay in a bounded domain
7023: \cite{skr00,bor95}. The enstrophy eigenvalues,
7024: $\Omega_{\{1,2,3\}}$ decay as $t^{-12/5}$ as predicted from a
7025: simple dimensional argument \cite{bif03c}.
7026: To focus on the process of recovery of isotropy
7027: in terms of global quantities one may track the behavior of
7028: two sets of observables
7029: $$ \Delta_{il} E(t) =\frac{\la E_i(t)-E_l(t) \ra}{\la E_i(t) +
7030: E_l(t)}
7031: \ra\,,\quad
7032: \Delta_{il} {\Omega}(t) = \frac{ \la \Omega_i(t)-\Omega_l(t) \ra}
7033: {\la \Omega_i(t) + \Omega_l(t)\ra } \,, $$
7034: which vanish for isotropic statistics. Their rate of decay is therefore
7035: a direct measurement of the return to isotropy. The energy matrix $E_{il}$
7036: is particularly sensitive to the large scales while small-scale
7037: fluctuations are sampled by $\Omega_{il}$. As seen from
7038: Fig.~\ref{fig:3},
7039: \begin{figure}[h]
7040: %\epsfxsize=8.0truecm
7041: %\epsfysize=6.0 cm
7042: \epsfbox{DEC.fig3.eps}
7043: \caption{Log-log plot of the anisotropy content at the large scales
7044: ($\Delta_{12}E(t)$, $\Delta_{13}E(t)$, top curves) and the small scales
7045: ($\Delta_{12}\Omega(t)$, $\Delta_{13}\Omega(t)$, bottom curves) as a
7046: function
7047: of time, expressed in $\tau_0$ unit. The large-scale (small-scale)
7048: anisotropy
7049: content is defined as the mismatch between the eigenvalues of the
7050: single-point
7051: velocity (vorticity) correlation.}
7052: \label{fig:3}
7053: \end{figure}
7054: both large and small scales begin to isotropize
7055: after roughly one eddy turnover time and become fully isotropic
7056: (within statistical fluctuations) after $100$ eddy turnover
7057: times. However, small scales show an overall degree of anisotropy much
7058: smaller than the large scales.
7059:
7060: %---------------------------------------------------------
7061: %The observables which characterize the decay of small-scale velocity
7062: %fluctuations are the longitudinal structure functions
7063: %\be
7064: %S^{(n)}\lp\r,t\rp = \la\left[\lp \v\lp\x+ \r,t\rp-\v\lp\x,t\rp\rp
7065: % \cdot\hat\r\right]^n \ra\,.
7066: %\label{eq:longitudinal}
7067: %\ee
7068: % Another method, equivalent to the SO(3) decomposition
7069: %of longitudinal structure functions, to look at
7070: %the anisotropic properties of the velocity field is to
7071: %focus directly on the PDF of the longitudinal velocity
7072: %differences. In this case, denoting by $\cP(\Delta,\r;t)$ the
7073: %probability that the longitudinal incremement $\dv \equiv
7074: %\lp\v(\r,t)-\v(0,t)
7075: %\rp \cdot \unitr$ be equal to $\Delta$, we may project $\cP(\Delta,\r;t)$
7076: %on
7077: %the SO(3) basis:
7078: %\be \cP(\Delta,\r;t) =\sum_{j=0}^{\infty}\sum_{m=-j}^{j}
7079: %\cP_{jm}\lp r,\Delta;t\rp Y_{jm}(\unitr).
7080: %\label{so3_pdf}
7081: %\ee
7082: %The projection $\cP_{jm}\lp r,\Delta;t\rp$ plays the role of an
7083: %``effective PDF'' for each single SO(3) sector. Indeed, the
7084: %projections of the longitudinal structure function on any sector
7085: %$(j,m)$ can be resphconstructed from the corresponding $\cP_{jm}\lp
7086: %r,\Delta;t\rp$ by averaging over all possible $\Delta$'s:
7087: %\be
7088: %S^{(n)}_{jm}\lp r,t\rp =
7089: %\int d\Delta \Delta^n \cP_{jm}\lp r,\Delta;t\rp\,.
7090: %\label{pdf_jm}
7091: %\ee
7092: %That establishes the equivalence between the decompositions
7093: %(\ref{so3_sf}) and (\ref{so3_pdf}).\\ The main points broached here
7094: %are about the long-time properties of the SO(3) projections,
7095: %$S^{(n)}_{jm}\lp r,t\rp$.
7096: Concerning small scales properties, in \cite{bif03c}
7097: a simple anisotropic generalization of the
7098: self-preservation hypothesis (see, e.g.~Ref.~\cite{fri95}) was proposed:
7099: $$
7100: S^{(n)}_{jm}\lp r,t\rp =
7101: V^{(n)}_{jm}\lp t \rp f^{(n)}_{jm} \lp r/L_{jm}(t) \rp .
7102: $$
7103: Here with $V^{(n)}_{jm}\lp t \rp$ we take explicitly into account
7104: the fact that large-scale velocity properties may depend in a nontrivial
7105: way
7106: on both $(j,m)$ and the order $n$. Furthermore, $L_{jm}(t)$ accounts
7107: for the possibility that the characteristic length scale depend on the
7108: SO(3) sector. In analogy with the observations made in the
7109: stationary case \cite{gar98,ara98,ara99,kur00,bif01a,bif02,she02,bif01}
7110: a scaling law was postulated:
7111: \be S^{(n)}_{jm}\lp r,t \rp \sim a^{(n)}_{jm}(t) \lp \frac{r}
7112: {L_{jm}(t)}\rp^{\zeta_{j}^{(n)}}\,.
7113: \label{anyt}
7114: \ee
7115: The time behavior is encoded in both the decay of the overall
7116: intensity, accounted by the prefactors $a^{(n)}_{jm}(t)$, and the
7117: variation of the integral scales $L_{jm}(t)$. The representation
7118: Eq.~(\ref{anyt}) is the simplest one fitting the initial time statistics
7119: for $t=0$ and agreeing with the evolution given by the self
7120: preservation hypothesis in the isotropic case. The power law behavior
7121: for $f^{(n)}_{jm} \lp r/L_{jm}(t) \rp$ can be expected only in a
7122: time-dependent inertial range of scales $ \eta(t) \ll r \ll L(t)$. As
7123: for the exponents appearing in (\ref{anyt}), their values are
7124: expectedly the same as in the stationary case. Concerning the time
7125: evolution, it seems difficult to disentangle the dependence due to the
7126: decay of $a^{(n)}_{jm}(t)$ from the one due to the growth of the
7127: integral scale $L_{jm}(t)$. The existence of a running reference
7128: scale, $L_{jm}(t)$ introduces some non-trivial relations between the
7129: spatial anomalous scaling and the decaying time properties, and those
7130: relations might be subject to experimental verification. In the case
7131: discussed in \cite{bif03c}, the fact that the initial condition
7132: has a characteristic length-scale comparable with the box size,
7133: simplifies the matter. Indeed we expect that $L_{jm}(t) \approx
7134: L_{0}$, and the decay is due only to the fall-off of
7135: $a^{(n)}_{jm}(t)$. An obvious shortcoming is that the width
7136: of the inertial range $L_{0}/\eta(t)$ shrinks monotonically in time,
7137: thereby limiting the possibility of precise quantitative statements.
7138: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7139: \subsubsection{Long time decay}
7140: A quantitative way to define the temporal rate of recovery of isotropy
7141: at a fixed scale in the inertial range is given by the
7142: dimensionless ratio:
7143: \be \Pi^{(n)}_{jm}\lp r,t \rp \equiv
7144: \frac{S^{(n)}_{jm}\lp r,t \rp}{S^{(n)}_{0,0}\lp r,t \rp} \sim
7145: t^{-\Xi_{j}^{(n)}}\,.
7146: \label{ratio_n}
7147: \ee
7148: In Fig.~\ref{fig:9},
7149: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
7150: \begin{figure}[h]
7151: %\epsfxsize=8.0truecm
7152: %\epsfysize=6.0 cm
7153: \epsfbox{DEC.fig9.eps}
7154: \caption{Hierarchical organization of anisotropic fluctuations at long
7155: times. Log-log plot of the anisotropic projections normalized by the
7156: corresponding isotropic projection (see text), at two fixed scales $r=80$
7157: and
7158: $r=40$ (inset) for $n=2, 4, 6$ in the anisotropic sector $j=4,m=0$.
7159: Symbols read as follows\,: $\Pi^{(2)}_{40}$ (full box);
7160: $\Pi^{(4)}_{40}$ (star);
7161: $\Pi^{(6)}_{40}$ (empty box). The straight line
7162: is $t^{-\Xi}$ with $\Xi \sim 0.3$.
7163: Same symbols in the inset.}
7164: \label{fig:9}
7165: \end{figure}
7166: %%%%%%%%%%%%%%%%%
7167: we plot $\Pi^{(n)}_{jm}\lp r,t \rp$ at $r=80$
7168: for structure functions of order $n=2,4,6$ and for the most intense
7169: anisotropic sector, $(j,m)=(4,0)$. All anisotropic sectors,
7170: for all measured structure functions, decay faster than the isotropic
7171: one. The measured slope in the decay is about $\Xi_j^{(n)} \sim 0.3$
7172: for all $n$, within the statistical errors.
7173: Note that these results agree with the simple picture that the
7174: time-dependence in (\ref{anyt}) is entirely carried by the prefactors
7175: $a_{jm}^{(n)}(t)$ and the value of the integral scales $L_{jm}(t)$ is
7176: saturated at the size of the box. Indeed, by assuming that large-scale
7177: fluctuations are almost Gaussian we have that the leading
7178: time-dependence of $a_{jm}^{(2n)}$ is given by $a_{jm}^{(2)}
7179: a_{00}^{(2n-2)}$. For the isotropic sector, $a_{00}^{(2n)} \sim
7180: (a_{00}^{(2)})^n$, and plugging that in (\ref{ratio_n}), one get:
7181: $\Pi^{(n)}_{jm}\lp r,t \rp \sim a^{(2)}_{jm}(t)/a^{(2)}_{00}(t) \sim
7182: t^{-\Xi}$ with $\Xi \sim 0.3 (\pm 0.1)$ independent of $n$. The
7183: quality of data is insufficient to detect possible residual effects
7184: due to $L_{jm}(t)$, which would make $\Xi_j^{(n)}$ depend on $n$ and
7185: $j$ because of spatial intermittency.\\
7186:
7187: Let us denote with $\cP(\Delta,\r;t)$ the probability to observe a
7188: given longitudinal fluctuation, $\delta u_{\ell}(\B r,t)=\Delta$
7189: in the direction $\r$ at a given time,
7190: $t$. For any given fixed value $\Delta$ and for any given time, $t$,
7191: we can project $\cP(\Delta,\r;t)$ on the SO(3) basis functions:
7192: \be
7193: \cP(\Delta,\r;t) = \sum_{j=0}^{\infty}\sum_{m=-j}^{j}
7194: \cP_{jm}\lp r,\Delta;t\rp Y_{jm}(\unitr),
7195: \label{so3_pdf}
7196: \ee
7197: where now the projection, $\cP_{jm}\lp r,\Delta;t\rp$ play the role
7198: of an {\it effective PDF} for each SO(3) sector. The projection
7199: of any longitudinal structure function,
7200: $ S^{(n)}(\r,t) $ on any sector, $(j,m)$ can be reconstructed from the
7201: corresponding
7202: projection of the PDF on the same sector,
7203: $\cP_{jm}\lp r,\Delta;t\rp$, by averaging over all possible $\Delta$:
7204: $$
7205: S^{(n)}_{jm}\lp r,t\rp = \int d\Delta \Delta^n \cP_{jm}\lp r,\Delta;t\rp
7206: $$
7207: which establish the link between decomposition (\ref{eq:fundamental}) and
7208: (\ref{so3_pdf}).\\
7209:
7210: The interesting fact that the
7211: decay properties of the anisotropic sectors are almost independent of
7212: $n$ indicates that a non-trivial time
7213: dependence in the shape of the PDF's $\cP_{jm}\lp r ,\Delta;t\rp$ for
7214: $j>0$ must be expected. The most accurate way to probe the rescaling
7215: properties of $\cP_{jm}\lp r,\Delta;t\rp$ in time is to compute the
7216: generalized flatness:
7217: $$
7218: K_{jm}^{(n)}(r,t)\equiv \frac{S^{(n)}_{jm}\lp r,t \rp}{\lp
7219: S^{(2)}_{jm}\lp r,t \rp \rp^{\frac{n}{2}}} \sim t^{\alpha_{j}^{(n)}}
7220: $$
7221: Were the PDF projection in the $(j,m)$ sector self-similar for $t \gg
7222: \tau_0$, then $K_{jm}^{(n)}(r,t)$ would tend to constant values. This
7223: is not the case for anisotropic fluctuations, as it is shown in
7224: Fig. (\ref{fig:10}).
7225: \begin{figure}[h]
7226: %\epsfxsize=7.9truecm
7227: %\epsfysize=6.0 cm
7228: \epsfbox{DEC.fig10.eps}
7229: \caption{Log-log plot of the generalized flatness, $K_{jm}^{(n)}(r,t)$
7230: of order $n=4,6$ for both the isotropic (two bottom curves), and
7231: the anisotropic sector $(j=4,m=0)$ (two top curves) at $r=80$, and as
7232: a function of time. In the inset we plot the same quantities, in the same
7233: order, at a different inertial range scale, $r=40$.}
7234: \label{fig:10}
7235: \end{figure}
7236: The curves $K_{jm}^{(n)}(r,t)$ are collected for
7237: two fixed inertial range separations, $r=80$ and $r=40$ (inset), for
7238: two different orders, $n=4,6$ and for both the isotropic and one of
7239: the most intense anisotropic sectors $(j=4,m=0)$ . The isotropic
7240: flatness tends toward a constant value for large $t$. Conversely, its
7241: anisotropic counterparts are monotonically increasing with $t$,
7242: indicating a tendency for the anisotropic fluctuations to become more
7243: and more intermittent as time elapses. Also the behavior in
7244: Fig.~\ref{fig:10} is in qualitative agreement with the observation
7245: previously made that all the time dependence can be accounted for by the
7246: prefactors $a_{jm}^{(n)}(t)$. Indeed, assuming that the length scales
7247: $L_{jm}(t)$ have saturated and that the large scale PDF is close to
7248: Gaussian, it is easy to work out the prediction $K_{jm}^{(n)}(r,t)
7249: \sim t^{-\Xi(1-n/2)}$, i.e. $\alpha_{j}^{(n)}=\Xi(n/2-1)$.
7250: %------------------------------------------------------------------
7251: %-------------------------------------------------------------
7252: We conclude this
7253: section by a brief summary of the results. It was found that
7254: isotropic fluctuations persist longer than anisotropic ones,
7255: i.e. there is a time-recovery, albeit slower than predicted by dimensional
7256: arguments, of isotropy during the decay process. It was also found
7257: that isotropic fluctuations decay in an almost self-similar way while
7258: the anisotropic ones become more and more intermittent. Qualitatively,
7259: velocity configurations get more isotropic but anisotropic
7260: fluctuations become, in relative terms, more ``spiky'' than the
7261: isotropic ones as time elapses.
7262: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7263: \subsubsection{Short-time decay}
7264: \label{shorttimes}
7265: It is interesting to notice that it is possible to relate the
7266: small-scale universal properties of forced turbulent statistics to
7267: those of short-time decay for an ensemble of initial configurations
7268: \cite{bif03c}. As already remarked, one cannot expect an
7269: universal behavior for all statistical observables, as the very
7270: existence of anomalous scaling is the signature of the memory of the
7271: boundaries and/or the external forcing throughout all the scales.
7272: Indeed, the main message we want to convey here is that only the
7273: scaling exponents of both isotropic and anisotropic small-scale
7274: fluctuations are universal, at least for forcing concentrated at large
7275: scales. The prefactors are not expected to be so. There is therefore
7276: no reason to expect that quantities such as the skewness, the kurtosis
7277: and in fact the whole PDF of velocity increments or gradients be
7278: universal. \\
7279: This is the same situation that we discussed in great details in previous
7280: sections
7281: for the passive transport
7282: of scalar and vector fields.
7283: However, carrying over the analytical
7284: knowledge developed for linear hydrodynamic problems involve some
7285: nontrivial, yet missing, steps. For the Navier-Stokes dynamics, linear
7286: equations of motion appear when we consider
7287: the whole set of correlation functions as discussed in
7288: Sect.~\ref{sec:hierachy}.
7289: These equations
7290: can be rewritten in a schematic form:
7291: \be
7292: \partial_t C^{(n)} = \Gamma^{(n+1)}C^{(n+1)} +\nu D^{(n)} C^{(n)} + F^{(n)},
7293: \label{compact}
7294: \ee
7295: where $\Gamma^{(n+1)}$ is the integro-differential linear operator
7296: coming from the inertial and pressure terms, $C^{(n)}$ is a
7297: shorthand notation for a generic $(n)$-th order correlator and
7298: $D^{(n)}$ is the linear
7299: operator describing dissipative effects. Finally, $F^{(n)}$ is the
7300: correlator involving increments of the large-scale forcing $\f$ and of
7301: the velocity field. The balance between inertial and injection
7302: terms cannot lead to anomalous scaling. A natural possibility is that
7303: a mechanism similar to the one identified in linear transport problems
7304: be at work in the Navier-Stokes case as well. The anomalous
7305: contributions to the correlators would then be associated with
7306: statistically stationary solutions of the unforced equations
7307: (\ref{compact}). The scaling exponents would {\it a fortiori} be
7308: independent of the forcing and thus universal. As for the prefactors,
7309: the anomalous scaling exponents are positive and thus the anomalous
7310: contributions grow at infinity. They should then be matched at the
7311: large scales with the contributions coming from the forcing to ensure
7312: that the resulting combination vanish at infinity, as required for
7313: correlation functions. The aim here is not to prove the previous
7314: points but rather to test whether they fail: the
7315: Navier-Stokes equations, being integro-differential and non-local,
7316: might directly couple inertial and injection scales and
7317: spoil the argument. This effect might be particularly relevant for
7318: anisotropic fluctuations where infrared divergences may appear in the
7319: pressure integrals (see Sect.~\ref{chap:linearP}).
7320: In order to investigate the previous point, we performed two sets of
7321: numerical experiments in decay.
7322:
7323: The first set, A, is of the same kind
7324: as in the previous section, i.e. we integrated the unforced
7325: Navier-Stokes equations
7326: with initial conditions picked from an ensemble obtained from a forced
7327: anisotropic stationary run. Statistical observables are measured as
7328: an {\it ensemble} average over the different initial conditions.
7329: The ensemble at the initial time of
7330: the decay process therefore coincides with the stationary
7331: state in forced runs. If correlation functions are indeed dominated
7332: at small scales by statistically stationary solutions of the unforced
7333: equations then the field should not decay. Specifically, the field
7334: should not vary for times smaller than the large-scale eddy turnover
7335: time $\tau_0$. Those are the times when the effects
7336: of the forcing terms start to be felt. Note that this should hold
7337: at all scales, including the small ones whose turnover times are much
7338: faster than $\tau_0$.
7339:
7340: The second set of numerical simulations (set B) takes the same initial
7341: conditions but for the random scrambling of the phases\,:
7342: $\Bu_i(\k)
7343: \rightarrow P_{il}(\k) \B u_l(\k) \exp(i \theta_l(\k)) $, with
7344: $\theta_l(\k)$ i.i.d. random variables. In this way, the spectrum and
7345: its scaling exponent are preserved but the wrong organization of the
7346: phases is expected to spoil the statistical stationarity of the
7347: initial ensemble. As a consequence, two different decays are expected
7348: for the two sets of initial conditions. In particular, contrary to
7349: set A, set B should vary at small scales on times of the order of the
7350: eddy turnover times $\tau_r \sim r^{2/3}$. This is exactly what has
7351: been found in the numerical simulations for both isotopic and
7352: anisotropic statistics as can be seen for the anisotropic case in
7353: Fig.~\ref{fig:12},
7354: %%%%%%%%%%%%%%%%%%%%%%%%%%%
7355: \begin{figure}[h]
7356: %\epsfxsize=8.3truecm
7357: %\epsfysize=6.0 cm
7358: \epsfbox{DEC.fig12.eps}
7359: \caption{
7360: Top: Temporal decay of the second-order anisotropic structure
7361: function $S^{(2)}_{40}(r,t)$, rescaled by its value at $t=0$.
7362: Here $r=30$, inside the inertial
7363: range. The two curves refer to the time evolution of the structure
7364: function starting from the forced-stationary velocity
7365: fields (squares, set A) and from the randomly dephased velocity
7366: fields (circles, set B). Time is normalized by the integral eddy turnover
7367: time.
7368: Notice that for set B we observe changes on a time
7369: scale faster than the integral eddy turnover time. That is to be
7370: contrasted with the case A, where structure functions are strictly constant
7371: in time up to an integral eddy turnover
7372: time. Bottom: The same curves but for the fourth-order structure function.}
7373: \label{fig:12}
7374: \end{figure}
7375: %%%%%%%%%%%%%%%%%%%%%
7376: where the temporal behavior of
7377: longitudinal structure functions of order 2 and 4 is shown. The
7378: scaling exponents of the contributions responsible for the observed behavior
7379: at
7380: small scales are thus forcing independent.\\
7381: To conclude, the data presented here support the conclusion that
7382: nonlocal effects peculiar to the Navier-Stokes dynamics do not spoil
7383: arguments on universality based on analogies with passive turbulent
7384: transport. The picture of the anomalous contributions to the
7385: correlation functions having universal scaling exponents and
7386: non-universal prefactors follows.
7387:
7388: \newpage
7389: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7390: %%
7391: %
7392: % Chapter 9 - Concluding Discussion
7393: %
7394: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7395: %%
7396:
7397: \section{Concluding Discussion, }
7398: \label{chap:conclusions}
7399:
7400: In this review we presented a mathematical framework in which
7401: anisotropy in turbulence can be studied, and we have tested its
7402: utility in the context of experimental analysis, numerical simulations
7403: and analytical models. The basic idea is to express the various
7404: statistical quantities of turbulence (e.g., structure functions,
7405: correlation functions) in terms of their projections on the different
7406: sectors of the SO($3$) group.
7407:
7408: The utility of the SO($3$) decomposition should be assessed in two
7409: main aspects. The first aspect is its functionality as a tool for
7410: characterizing anisotropy, whereas the second, and deeper aspect, is its
7411: physical relevance and the theoretical and analytical advantages that
7412: are gained by using it.
7413:
7414: As a tool for describing anisotropy, the SO($3$) decomposition is
7415: probably the most natural and general method. It is of \emph{high
7416: resolution} - it subdivides the observed anisotropy into different
7417: sectors - the $(j,m)$ sectors. The weights of the various sectors
7418: give us a fine resolution of the anisotropy structure. Instead of
7419: having one measure for anisotropy (e.g., the overall percentage of
7420: anisotropy), we have an infinite set of numbers that compose a
7421: detailed profile of the anisotropy structure.
7422:
7423:
7424: The SO($3$) decomposition is also very \emph{general}. It is
7425: applicable to any physical observable that has a well-defined
7426: transformation under rotations. These can be, for example, correlation
7427: functions, structure functions or Green's functions (response
7428: functions). The observables themselves may depend on any number of
7429: space coordinates or even be space independent. They may also be
7430: scalars, vectors or tensors. Any such quantity can be presented as
7431: a sum of parts that belong to the different $(j,m)$ sectors of the
7432: rotation group. Additionally, since the SO($3$) decomposition is
7433: invariant to isotropic operations, it is invariant to the most common
7434: operations that we use. For example, to obtain the $n$th order
7435: longitudinal structure function we can take the full $n$th order
7436: structure function (which is a tensor) and contract it with $n$ unit
7437: vectors in the direction of the separation distance. Since this
7438: operation is linear and isotropic, it will preserve the $(j,m)$
7439: sectors of the full structure function. That is, the $(j,m)$ sector
7440: of the full structure function will be transformed into the same
7441: $(j,m)$ sector of the longitudinal structure functions. The same
7442: thing happens for operations such as differentiations (for example
7443: when we look at moments of the gradient fields), space averaging, time
7444: averaging, coordinate fusion etc... From the pure theoretical point of view
7445: the first and most obvious advantage of using the
7446: SO($3$) decomposition is its elegance and the overwhelming
7447: simplification that it offers in analytical calculations.
7448: The SO($3$) decomposition may also have a deeper physical
7449: justification if it produces universal quantities such as distinct
7450: scaling behavior in the anisotropic sectors. There are several
7451: different evidences that suggest that this is indeed the
7452: case. Experimental results clearly show that a better scaling is
7453: achieved if we take higher $j$ components into account.
7454: Additionally, different experimental setups
7455: seem to lead to the same numerical values of the anisotropic exponent.
7456: This is
7457: a strong support for the hypothesis that the
7458: anisotropic sectors of the structure functions have universal
7459: exponents.
7460: Another support for the idea that the SO($3$) decomposition exposes
7461: universal quantities, comes from numerical simulations.
7462: In DNS, the SO($3$) decomposition
7463: can be performed directly (since the velocity field is accessible in
7464: every point in space and time) which makes the results much less
7465: ambiguous. We clearly see that even in the very moderate Reynolds
7466: numbers of the simulations, a scaling behavior is detected once we use
7467: the SO($3$) decomposition. In some cases, without the SO($3$)
7468: decomposition no scaling behavior is seen at all. Furthermore, the
7469: resulting exponents in the isotropic sector are remarkably similar to
7470: the experimental values which are measured at very high Reynolds
7471: numbers. This is a strong indication that at least the isotropic
7472: sector has a universal profile, and therefore by disentangling it from
7473: the anisotropic sectors we get universal results. In other sectors of
7474: the rotation group, the scaling behavior is not as good, and in some
7475: sectors there is no scaling at all. However, in those sectors where
7476: scaling was detected, the scaling exponents seem to agree with the
7477: theoretical and experimental predictions.
7478: Sectors with same $j$ and different $m$s had
7479: the same scaling exponents (when scale-invariance was observed). And
7480: finally, all exponents increased as a function of $n$ (order of the
7481: structure-function) as well as $j$. It is still not clear whether
7482: the ``bad'', non scale-invariant behavior that was detected in some
7483: sectors, is a result of the poor Reynolds numbers of the simulations,
7484: or is a genuine effect that tells us that the foliation picture is
7485: incomplete. A further research with higher resolution is probably
7486: needed to settle this issue.
7487: % From the theoretical point of view, there is
7488: %little hope that the
7489: %universality of the anisotropic sectors in Navier-Stokes turbulence
7490: %stems from exact foliation. Exact foliation occurs in
7491: %linear models.
7492: In the Navier-Stokes case
7493: one can prove a ``weak foliation''. Weak
7494: foliation is an approximate foliation that happens in the case of weak
7495: anisotropy, when we linearize the anisotropic part of the theory
7496: around its isotropic part.
7497: In such case, the linearized anisotropic part of the
7498: theory is subject to a linear and isotropic equation whose kernel
7499: contains the isotropic solution, and hence foliation occurs. This is a
7500: very robust approximation since it holds for virtually any nonlinear
7501: and isotropic theory in the case of weak anisotropy.
7502: Additionally, we know that as the Reynolds number
7503: increases, the statistics becomes more and more isotropic and
7504: therefore the linear approximation becomes better and better.
7505:
7506: The SO($3$) decomposition has a physical relevance also in the
7507: presence of strong anisotropy. If we merely consider the direct
7508: application of the Navier-Stokes and continuity equations on the
7509: various structure functions, we get a family of linear and isotropic
7510: constraints. These constraints are valid independently of the
7511: amount of anisotropy in the system. Their linearity and isotropy lead
7512: to foliation and hence one can discuss them in every sector
7513: independently. In some sectors, they are sufficient to determine the
7514: full solution, whereas in others they can reveal some general
7515: properties of the solution. For example, the isotropic sector of the
7516: third-order structure function is completely determined by the
7517: different constraints, and is given by the well-known 4/5 law of
7518: Kolmogorov. Notice that because of foliation, this is true also in the
7519: presence of anisotropy, which means that the 4/5 law holds also in an
7520: anisotropic turbulence. In the $j=2$ sectors, on the other hand,
7521: the third-order structure-function is given by two undetermined scalar
7522: functions, whereas in the $j=4,6,\ldots$ it is given by
7523: three. Another example is the $j=1$ sectors in the second-order
7524: correlation function, which
7525: must all vanish.
7526:
7527: To conclude, the framework of the SO($3$) decomposition provides an
7528: elegant and efficient way to describe anisotropy in turbulence. It
7529: also greatly simplifies many analytical calculations that involve
7530: anisotropic quantities, mainly through the mechanism of
7531: foliation. This mechanism is present in simplified models of
7532: turbulence, and may also be valid approximately in Navier-Stokes
7533: turbulence. It predicts that the anisotropic sectors of the statistics
7534: have universal properties such as scaling exponents. Further research
7535: is needed to accurately measure these anisotropic exponents in
7536: experiments as well as in numerical simulations.
7537:
7538: A quantitative computation of the anisotropic exponents in Navier-Stokes
7539: turbulence from first principles may very well be an illusive goal.
7540: Nevertheless, we believe that the general principles of the physics
7541: behind these exponents and anisotropy in general can be
7542: understood.
7543: % The best tool for that was, and still is, the theoretical
7544: %study of simplified models.
7545: %%%%%%%%%%%%%%%%%%%%%%%%%%
7546: \section{Acknowledgements}
7547: We benefited from many discussion with many colleagues, without which we
7548: could not reach the picture presented in this review. We thank in particular
7549: Itai Arad, Roberto Benzi, Guido Boffetta, Carlo Casciola, Antonio Celani,
7550: Isabelle Daumont, Siegfried Grossmann, Boris Jacob, Susan
7551: Kurien, Alessandra Lanotte, Detlef Lohse, Victor S. Lvov, Irene
7552: Mazzitelli,
7553: Evgeny Podivilov, K.R. Sreenivasan, Federico Toschi and Massimo Vergassola.
7554: Much
7555: of the work reviewed here was
7556: supported in part by the European Commission through a TMR grant (No.
7557: HPRN-CT 2000-00162),
7558: the Italian Ministry of Education (MURST), the Instituto Nazionale di
7559: Fisica
7560: della Materia, the German Israeli
7561: Foundation, the Minerva
7562: Foundation, Munich, Germany, the Israel Science Foundation, and the US
7563: Israel Bi-National
7564: Science Foundation.
7565: \appendix
7566: \section{The General Form of the 2nd Rank Tensor}
7567: \label{sec:general-form}
7568: In this appendix we discuss the general structure of the second rank
7569: correlation functions
7570: \begin{equation}
7571: F^{\alpha \beta }({\bf r})\equiv\langle u^{\alpha }({\bf x}+{\bf r})u^{\beta
7572: }({\bf
7573: x})\rangle \ , \label{cf}
7574: \end{equation}
7575: In (\ref{cf}) {\em homogeneity} of the flow is assumed, but not isotropy.
7576: Note that this object is more general than the structure function $S^{\alpha
7577: \beta }$ in being nonsymmetric in the indices, and having no definite
7578: parity. We wish to find the basis functions $B_{q,jm}^{\alpha \beta }(
7579: {\bf \hat{r}})$, with which we can represent $F^{\alpha \beta }({\bf r})$ in
7580: the form:
7581: \begin{equation}
7582: F^{\alpha \beta }({\bf r})=\sum_{q,jm}a_{q,jm}(r)B_{q,jm}^{\alpha \beta
7583: }({\bf
7584: \hat{r}}) \label{trial-tensor}
7585: \end{equation}
7586: and derive some constraints among the functions $\ a_{q,jm}(r)$ that result
7587: from incompressibility. We shall see, that due to the isotropy of the
7588: incompressibility conditions, the constraints are among $a_{q,jm}(r)$ with
7589: the {\em same} $j,m$ only.
7590:
7591: We begin by analyzing the incompressibility condition: An incompressible
7592: flow with constant density is characterized by the relation:
7593: $
7594: \partial _{\alpha }u^{\alpha }({\bf x},t)=0
7595: $
7596: as a result, one immediately gets the following constraints on $F^{\alpha
7597: \beta }({\bf r})$:
7598: $$
7599: \partial _{\alpha }F^{\alpha \beta }({\bf r}) =0, \qquad
7600: \partial _{\beta }F^{\alpha \beta }({\bf r}) =0.
7601: $$
7602: Plugging the trial tensor (\ref{trial-tensor}) into the last two equations
7603: we obtain 2 equations connecting the different $a_{q,jm}$:
7604: \be
7605: \partial _{\alpha }\sum_{q,jm}a_{q,jm}(r)B_{q,jm}^{\alpha \beta }({\bf
7606: \hat{r}}
7607: ) =0 \label{incomp-1} \qquad
7608: \partial _{\beta }\sum_{q,jm}a_{q,jm}(r)B_{q,jm}^{\alpha \beta }({\bf
7609: \hat{r}})
7610: =0
7611: \ee
7612: We first notice that the differentiation action is isotropic. As a result,
7613: if $T^{\alpha \beta }\left( {\bf r}\right) $ is some arbitrary tensor with a
7614: definite $j,m$ transformation properties, then the tensor $\partial _{\alpha
7615: }T^{\alpha \beta }({\bf r})$ will have {\em the same} $j,m$ transformation
7616: properties. Components with different $j,m$ are linearly independent.
7617: Therefore equations (\ref{incomp-1}) should hold for each $j,m$ separately.
7618:
7619: Next, we observe that (\ref{incomp-1}) are invariant under the
7620: transformation $
7621: F^{\alpha \beta }\longrightarrow F^{\beta \alpha }$. As a result, the
7622: symmetric and anti-symmetric parts of $F^{\alpha \beta }$ should satisfy
7623: (\ref{incomp-1}) independently. To see that, let us write $F^{\alpha \beta
7624: }$
7625: as a sum of a symmetric term and an anti-symmetric term: $F^{\alpha \beta
7626: }=F_{S}^{\alpha \beta }+F_{A}^{\alpha \beta }$, we then get:
7627: \begin{eqnarray*}
7628: \partial _{\alpha }F^{\alpha \beta } &=&\partial _{\alpha }F_{S}^{\alpha
7629: \beta }+\partial _{\alpha }F_{A}^{\alpha \beta }=\partial _{\alpha
7630: }F_{S}^{\beta \alpha }-\partial _{\alpha }F_{A}^{\beta \alpha }=0 \\
7631: \partial _{\beta }F^{\alpha \beta } &=&\partial _{\beta }F_{S}^{\alpha \beta
7632: }+\partial _{\beta }F_{A}^{\alpha \beta }=0
7633: \end{eqnarray*}
7634: from which we conclude:$
7635: \partial _{\alpha }F_{S}^{\alpha \beta }=\partial _{\alpha }F_{A}^{\alpha
7636: \beta }=0 .
7637: $
7638: Finally, (\ref{incomp-1}) is invariant under the transformation $F^{\alpha
7639: \beta }({\bf r)}\longrightarrow F^{\alpha \beta }(-{\bf r)}$ and as a result
7640: the odd parity and the even parity parts of $F^{\alpha \beta }$ should
7641: fulfill (\ref{incomp-1}) independently. We conclude that a necessary and
7642: sufficient condition for (\ref{incomp-1}) to hold is that it holds
7643: separately for parts with definite $j,m$, definite symmetry in the $\alpha
7644: ,\beta $ indices and a definite parity in ${\bf r}$:
7645: \[
7646: \partial _{\alpha }\sum_{q}a_{q,jm}({ r})B_{q,jm}^{\alpha \beta }({\bf
7647: \hat{r}})=0\mbox{ \
7648: \begin{tabular}{l}
7649: ~~{\small summation is over} \\
7650: ~~$B_{q,jm}^{\alpha \beta }$ {\small with definite symmetries}
7651: \end{tabular}
7652: }
7653: \]
7654: where the summation is over $q$ such that $B_{q,jm}^{\alpha \beta }$ has a
7655: definite indices symmetry and a definite parity.
7656:
7657:
7658: According to (\ref{eq:second-rank-tensors}) we can write these
7659: $B_{q,jm}^{\alpha \beta }$ as:
7660: \begin{enumerate}
7661: \item $(-)^{j}$ parity, symmetric tensors:
7662:
7663:
7664: \begin{itemize}
7665: \item $B_{1,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j}\delta ^{\alpha
7666: \beta }\Phi _{jm}(${\bf $r$}$)$,
7667: \item $B_{7,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j}\left[ r^{\alpha
7668: }\partial ^{\beta }+r^{\beta }\partial ^{\alpha }\right] \Phi _{jm}(${\bf
7669: $r$
7670: }$)$,
7671: \item $B_{9,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j-2}r^{\alpha
7672: }r^{\beta }\Phi _{jm}(${\bf $r$}$)$,
7673: \item $B_{5,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j+2}\partial
7674: ^{\alpha }\partial ^{\beta }\Phi _{jm}(${\bf $r$}$)$.
7675: \end{itemize}
7676: \item $(-)^{j}$ parity, anti-symmetric tensors:
7677: \begin{itemize}
7678: \item $B_{3,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j}\left[ r^{\alpha
7679: }\partial ^{\beta }-r^{\beta }\partial ^{\alpha }\right] \Phi _{jm}(${\bf
7680: $r$
7681: }$)$.
7682: \end{itemize}
7683: \item $(-)^{j+1}$ parity, symmetric tensors
7684: \begin{itemize}
7685: \item $B_{8,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j-1}\left[
7686: r^{\alpha }\epsilon ^{\beta \mu \nu }r_{\mu }\partial _{\nu }+r^{\beta
7687: }\epsilon ^{\alpha \mu \nu }r_{\mu }\partial _{\nu }\right] \Phi _{jm}(${\bf
7688: $r$}$)$,
7689: \item $B_{6,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j+1}\left[
7690: \epsilon ^{\beta \mu \nu }r_{\mu }\partial _{\nu }\partial ^{\alpha
7691: }+\epsilon ^{\alpha \mu \nu }r_{\mu }\partial _{\nu }\partial ^{\beta }
7692: \right] \Phi _{jm}(${\bf $r$}$)$.
7693: \end{itemize}
7694: \item $(-)^{j+1}$ parity, anti-symmetric tensors:
7695: \begin{itemize}
7696: \item $B_{4,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j-1}\epsilon
7697: ^{\alpha \beta \mu }r_{\mu }\Phi _{jm}(${\bf $r$}$)$,
7698: \item $B_{2,jm}^{\alpha \beta }(\hat{{\bf r}})\equiv r^{-j+1}\epsilon
7699: ^{\alpha \beta \mu }\partial _{\mu }\Phi _{jm}(${\bf $r$}$)$.
7700: \end{itemize}
7701: \end{enumerate}
7702: In order to differentiate these expressions we can use the following
7703: identities:
7704: \begin{eqnarray*}
7705: r^{\alpha }\partial _{\alpha }r^{\zeta }Y_{jm}(\hat{{\bf r}}) &=&\zeta
7706: r^{\zeta }Y_{jm}(\hat{{\bf r}}) \ ,\\
7707: \partial ^{\alpha }\partial _{\alpha }r^{\zeta }Y_{jm}(\hat{{\bf r}}) &=&
7708: \left[ \zeta \left( \zeta +1\right) -j(j+1)\right] r^{\zeta -2}Y_{jm}(
7709: \hat{x})
7710: \end{eqnarray*}
7711: which give rise to:
7712: \begin{eqnarray*}
7713: r^{\alpha }\partial _{\alpha }\Phi _{jm}({\bf r}) &=&j\Phi _{jm}({\bf r}) \
7714: .\\
7715: \partial ^{\alpha }\partial _{\alpha }\Phi _{jm}({\bf r}) &=&0 \ .
7716: \end{eqnarray*}
7717: From this point, it is a matter of simple (though somewhat lengthy)
7718: algebra to derive the differential constraints among $a_{q,jm}(r)$. The
7719: results are as follows:
7720: \begin{enumerate}
7721: \item $q\in \{1,7,9,5\}$
7722: \begin{eqnarray}
7723: &a&_{1,jm}^{\prime }(r)-jr^{-1}a_{1,jm}+ja_{7,jm}^{\prime
7724: }-j^{2}r^{-1}a_{7,jm}+a_{9,jm}^{\prime }+2r^{-1}a_{9,jm} =0 \ ,
7725: \label{eq:in1795} \\
7726: &r&^{-1}a_{1,jm}+a_{7,jm}^{\prime }+3r^{-1}a_{7,jm}+\left( j-1\right)
7727: a_{5,jm}^{\prime }-\left( j^{2}-3j+2\right) r^{-1}a_{5,jm} =0 \ .
7728: \nonumber
7729: \end{eqnarray}
7730: \item $q\in \{3\}$
7731: \begin{eqnarray}
7732: a_{3,jm}^{\prime }-jr^{-1}a_{3,jm} &=&0 \label{eq:in3} \ , \\
7733: a_{3,jm}^{\prime }+r^{-1}a_{3,jm} &=&0 \ . \nonumber
7734: \end{eqnarray}
7735: These equations have no solutions other than: $a_{3,jm}(r)=0$.
7736: \item $q\in \{8,6\}$
7737: \begin{equation}
7738: a_{8,jm}^{\prime }+3r^{-1}a_{8,jm}+(j-1)a_{6,jm}^{\prime }-\left(
7739: j^{2}-2j+1\right) r^{-1}a_{6,jm}=0 \ . \label{eq:in86}
7740: \end{equation}
7741: \item $q\in \{4,2\}$
7742: \begin{equation}
7743: r^{-1}a_{4,jm}-a_{2,jm}^{\prime }+(j-1)r^{-1}a_{2,jm}=0 \ . \label{eq:in42}
7744: \end{equation}
7745: \end{enumerate}
7746: There are obviously more unknowns than equations, since we merely exploited
7747: the incompressibility conditions. Nevertheless, we believe
7748: that the missing equations that arise from the dynamical hierarchy of
7749: equations will preserve the distinction between $a_{q,jm}$ of different
7750: $j,m$
7751: (again, due to the isotropy of these equations).
7752: Note also, that the above analysis holds also for the second-order structure
7753: function
7754: $$
7755: S^{\alpha \beta }({\bf r})\equiv \left\langle \left[ u^{\alpha }({\bf x}+
7756: {\bf r})-u^{\alpha }({\bf x})\right] \left[ u^{\beta }({\bf x}+{\bf r}
7757: )-u^{\beta }({\bf x})\right] \right\rangle .
7758: $$
7759: Only that in this case we should only consider the representations $
7760: q=1,7,9,5 $ for even $j$ and the representations $q=8,6$ for odd $j$. This
7761: follows from the fact that $S^{\alpha \beta }({\bf r})$ is symmetric with
7762: respect to its indices and it has an even parity in \ ${\bf r}$. Also, in
7763: that case, it is possible to go one step further by assuming a specific
7764: functional form for the $a_{q,jm}(r)$. We know that the $S^{\alpha \beta }(
7765: {\bf r})$ is expected scale in the inertial range, and we therefore may {\em
7766: assume}:
7767: $$
7768: a_{q,jm}(r)\equiv c_{q,jm}r^{\zeta _{2}^{(j)}}.
7769: $$
7770: where $c_{q,jm}$ are just numerical constants. If we now substitute this
7771: definition into the equations (\ref{eq:in1795},\ref{eq:in86}), we get a set
7772: of linear equations among the $c_{q,jm}$. These relations can be easily
7773: solved and give us two possible tensors for even $j$ ($q=1,7,9,5$) and one
7774: tensor form for odd $j$ (from $q=8,6$). This kind of approach was taken in
7775: the two-probes experiment which is described in Sect.~\ref{chap:experiment}.
7776: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7777: \section{Anisotropy in d-dimensions}
7778: \label{app:d-dim}
7779:
7780: To deal with anisotropy in d-dimensions we need classify the
7781: irreducible representations of the group of all $d$-dimensional
7782: rotations, SO(d) \cite{62Ham}, and then to find a proper basis
7783: for these representations. The main linear space that we work in
7784: (the carrier space) is the space of constant tensors with $n$
7785: indices. This space possesses a natural representation of SO(d),
7786: given by the well known transformation of tensors under
7787: $d$-dimensional rotation.
7788:
7789: The traditional method to find a basis for the irreducible
7790: representations of SO(d) in this space, is using the Young
7791: tableaux machinery on the subspace of traceless tensors
7792: \cite{62Ham,horvai}. It turns out that in the context of the present
7793: work, we do not need the explicit structure of these tensors.
7794: Instead, all that matters are some relations among them. A
7795: convenient way to derive these relations is to construct the
7796: basis tensors from functions on the unit $d$-dimensional sphere
7797: which belong to a specific irreducible representation. Here also,
7798: the explicit form of these functions in unimportant. All that
7799: matters for the calculations is the action of the Laplacian
7800: operator on these functions.
7801:
7802: Let us therefore consider first the space $\mathcal{S}_d$ of
7803: functions over the
7804: unit
7805: $d$-dimensional sphere. The representation of SO(d) over this
7806: space is naturally defined by:
7807: \begin{equation}
7808: \mathcal{O}_{\mathcal R} \Psi(\hat{u}) \equiv \Psi({\mathcal
7809: R}^{-1}\hat{u}) \ ,
7810: \label{d-Rot}
7811: \end{equation}
7812: where $\Psi(\hat{u})$ is any function on the $d$-dimensional
7813: sphere, and ${\mathcal R}$ is a $d$-dimensional rotation.
7814:
7815: ${\mathcal S}_d$ can be spanned by polynomials of the unit vector
7816: $\hat{u}$. Obviously (\ref{d-Rot}) does not change the degree of a
7817: polynomial, and therefore each irreducible representation in this
7818: space can be characterized by an integer $j=0,1,2,\ldots$,
7819: specifying the degree of the polynomials that span this
7820: representation. At this point, we cannot rule out the possibility
7821: that some other integers are needed to fully specify all
7822: irreducible representations in ${\mathcal S}_d$ and therefore we will
7823: need below another set of indices to complete the specification.
7824:
7825: We can now choose a basis of polynomials $\{ Y_{j,\sigma}(\hat
7826: u) \}$ that span all the irreducible representations of SO(d)
7827: over ${\mathcal S}_d$. The index $\sigma$ counts all integers other
7828: than $j$ needed to fully specify all irreducible
7829: representations, and in addition, it labels the different
7830: functions within each irreducible representation.
7831:
7832: Let us demonstrate this construction in two and three dimensions.
7833: In two dimensions $\sigma$ is unneeded since all the irreducible
7834: representation are one-dimensional and are spanned by
7835: $Y_j(\hat u)= e^{ij\phi}$ with $\phi$ being the angle
7836: between $\hat u$ and the the vector $\hat e_1\equiv (1,0)$. Any
7837: rotation of the coordinates in an angle $\phi_0$ results in a
7838: multiplicative factor $e^{i\phi_0}$. It is clear that
7839: $Y_j(\hat u)$ is a polynomial in $\hat u$ since $Y_j(\hat
7840: u)=[\hat u\cdot \hat p]^j$ where $\hat p\equiv (1,i)$. In
7841: three dimensions $\sigma=m$ where $m$ takes on $2j+1$ values
7842: $m=-j, -j+1,\dots,j$. Here $Y_{j,m}\propto e^{im\phi}
7843: P^m_j(\cos\theta)$ where $\phi$ and $\theta$ are the usual
7844: spherical coordinates, and $P^m_j$ is the associated Legendre
7845: polynomial of degree $j-m$. Obviously we again have a
7846: polynomial in $\hat u$ of degree $j$.
7847:
7848: We now wish to calculate the action of the Laplacian operator
7849: with respect to $u$ on the $Y_{j,\sigma}(\hat u)$. We prove
7850: the following identity:
7851: \begin{equation}
7852: u^2\partial^\alpha\partial_\alpha Y_{j,\sigma}(\hat u) =
7853: -j(j+d-2)Y_{j,\sigma}(\hat u) \ . \label{LapY}
7854: \end{equation}
7855: One can easily check that for $d=3$ (\ref{LapY}) gives the factor
7856: $j (j + 1)$, well known from the theory of angular-momentum
7857: in Quantum Mechanics. To prove this identity for any $d$, note
7858: that
7859: \begin{equation}
7860: |u|^{2-j}\partial^2 |u|^j Y_{j,\sigma}(\hat u)) =0 \ .
7861: \label{firststep}
7862: \end{equation}
7863: This follows from the fact that the Laplacian is an isotropic
7864: operator, and therefore is diagonal in the $Y_{j,\sigma}$. The
7865: same is true for the operator $|u|^{2-j}\partial^2 |u|^j$.
7866: But this operator results in a polynomial in $\hat u$ of degree
7867: $j-2$, which is spanned by $Y_{j',\sigma'}$ such that
7868: $j'\le j -2$. Therefore the RHS of (\ref{firststep}) must
7869: vanish. Accordingly we write
7870: \begin{equation}
7871: \partial^2|u^j|Y_{j,\sigma}(\hat
7872: u)+2\partial^\alpha|u^j|\partial^\alpha
7873: Y_{j,\sigma}+|u^j|\partial^2 Y_{j,\sigma}(\hat u) = 0 \ .
7874: \end{equation}
7875: The second term vanishes since it contains a radial derivative
7876: $u^\alpha\partial_\alpha$ operating on $Y_{j,\sigma}(\hat u)$
7877: which depends on $\hat u$ only. The first and third terms, upon
7878: elementary manipulations, lead to (\ref{LapY}).
7879:
7880: Having the $Y_{j,\sigma}(\hat u)$ we can now construct the
7881: irreducible representations in the space of constant tensors. The
7882: method is based on acting on the $Y_{j,\sigma}(\hat u)$ with
7883: the {\em isotropic} operators $u^\alpha,~\partial^\alpha$ and
7884: $\delta^{\alpha\beta}$. Due to the isotropy of the above
7885: operators, the behavior of the resulting expressions under
7886: rotations is similar to the behavior of the scalar function we
7887: started with. For example, the tensor fields
7888: $\delta^{\alpha\beta}Y_{j,\sigma}(\hat u),
7889: \partial^{\alpha}\partial^{\beta}Y_{j,\sigma}(\hat u) $
7890: transform under rotations according to the $(j, \sigma)$
7891: sector of SO(d).
7892:
7893: Next, we wish to find the basis for the irreducible
7894: representations of the space of constant and fully symmetric
7895: tensors with $n$ indices. We form the basis
7896: \begin{equation}
7897: B^{\alpha_1,\dots,\alpha_{n}}_{n,j,\sigma}\equiv
7898: \partial^{\alpha_1} \dots \partial^{\alpha_{n}} u^{n}
7899: Y_{j,\sigma}(\hat u) , \quad j\le n \ . \label{Birr}
7900: \end{equation}
7901: Note that when $j$ {\em and} $n$ are even
7902: $B^{\alpha_1,\dots,\alpha_{n}}_{j,\sigma,n}$ no longer depends
7903: on $\hat u$, and is indeed fully symmetric by construction.
7904: Simple arguments can also prove that this basis is indeed
7905: complete, and spans {\em all} fully symmetric tensors with $n$
7906: indices. Other examples of this procedure for the other spaces
7907: are presented directly in the text.
7908:
7909: Finally let us introduce two identities involving the
7910: $B_{n,j,\sigma}$. The first one is
7911: \begin{eqnarray}
7912: \delta_{\alpha_1\alpha_2}B^{\alpha_1,\dots,\alpha_{n}}_{n,j,\sigma}
7913: &=&z_{n,j}B^{\alpha_3,\dots,\alpha_{n}}_{n-2,j,\sigma} \ ,
7914: \label{iden1}\\ z_{n,j}&=&[n(n+d-2)-j(j+d-2)] \ .
7915: \label{znl}
7916: \end{eqnarray}
7917: It is straightforward to derive this identity using (\ref{LapY}).
7918: The second identity is
7919: \begin{equation}
7920: \sum_{i\ne j} \delta^{\alpha_i\alpha_j}B^{\{\alpha_m\},m\ne
7921: i,j}_{n-2,j,\sigma}
7922: =B^{\alpha_1,\dots,\alpha_{n}}_{n,j,\sigma} \ , \quad j\le
7923: n-2 \ . \label{iden2}
7924: \end{equation}
7925: This identity is proven by writing $u^{n}$ in (\ref{Birr}) as
7926: $u^2 u^{n-2}$, and operating with the derivative on $u^2$. The
7927: term obtained as $u^2\partial^{\alpha_1} \dots
7928: \partial^{\alpha_{n}} u^{n-2} Y_{j,\sigma}(\hat u) $ vanishes
7929: because we have $n$ derivatives on a polynomial of degree $n-2$.
7930: It is worthwhile noticing that these identities connect tensors
7931: from two different spaces. The space of tensors with $n$ indices
7932: and the space of tensors with $n-2$ indices. Nevertheless, in
7933: both spaces, the tensors belong to the same $(j, \sigma)$
7934: sector of the SO(d) group. This is due to the isotropy of the
7935: contraction with $\delta^{\alpha_1\alpha_2}$ in the first
7936: identity, and the contraction with $\delta^{\alpha_i\alpha_j}$ in
7937: the second identity.
7938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7939: \section{Full Form for the $j=2$ Contribution for the Homogeneous Case}
7940: \label{app:fullj2}
7941: In this appendix we focus on the decomposition of second order
7942: tensorial structure functions up to $j=2$. For this purpose we define:
7943: $$
7944: S^{\alpha\beta}({\bf r}) = S_{j=0}^{\alpha\beta}({\B r})+
7945: S_{j=2}^{\alpha\beta}({\B r})
7946: $$
7947: The $j=0$ is well-known and given explicitly by:
7948: \begin{equation} S_{j=0}^{\alpha\beta}({\bf
7949: r})=c_0r^{\zeta_0^{(2)}} \left[(2+\zeta_0^{(2)}) \delta^
7950: {\alpha\beta}-\zeta_0^{(2)}{r^\alpha r^\beta\over r^2}\right]\ ,
7951: \label{Siso}
7952: \end{equation} where $\zeta_0^{(2)} \approx 0.68$
7953: is the known universal scaling
7954: exponent for the isotropic contribution and $c_0$ is
7955: an unknown coefficient that depends on the boundary conditions of
7956: the flow. For the $j=2$ sector which is the lowest contribution
7957: to anisotropy to the homogeneous structure function, the $m=0$
7958: (axisymmetric) terms were derived from constraints of symmetry,
7959: even parity (because of homogeneity) and incompressibility on the
7960: second order structure function \cite{ara98}
7961: \bea
7962: &S&^{\alpha\beta}_{j=2,m=0} ({\bf r}) =
7963: ar^{\zeta_2^{(2)}}\Big[(\zeta_2^{(2)} -2)\delta^{\alpha\beta} -
7964: \zeta_2^{(2)}(\zeta_2^{(2)}+6)
7965: \delta^{\alpha\beta}
7966: {(\B n\cdot \B r)^2 \over r^2}+2\zeta_2^{(2)}(\zeta_2^{(2)}-2){r^\alpha
7967: r^\beta(\B n\cdot \B r)^2 \over r^4} \nonumber \\\label{m0}
7968: &+&([\zeta_2^{(2)}]^2+3\zeta_2^{(2)}+6)n^\alpha n^\beta
7969: -{\zeta_2^{(2)}(\zeta_2^{(2)}-2)\over r^2}(r^\alpha n^\beta +
7970: r^\beta n^\alpha)(\B n\cdot \B r)\Big]\\ &+& \nonumber
7971: br^{\zeta_2^{(2)}}\Big [-(\zeta_2^{(2)}
7972: +3)(\zeta_2^{(2)}+2)\delta^ {\alpha\beta}(\B n\cdot \B r)^2 +
7973: {r^\alpha r^\beta \over r^2} + (\zeta_2^{(2)} +3)
7974: (\zeta_2^{(2)}+2)n^\alpha n^\beta\\& + &(2\zeta_2^ {(2)}+1)(\zeta_2^{(2)}-2)
7975: {{r^\alpha}{r^\beta}{(\B n\cdot \B r)^2} \over r^4}-
7976: ([\zeta_2^{(2)}]^2 - 4)(r^\alpha n^\beta + r^\beta n^\alpha)(\B
7977: n\cdot \B r)\Big] \ . \nonumber
7978: \end{eqnarray}
7979: where $\zeta_2^{(2)}$ is the universal scaling exponent for the
7980: $j=2$ anisotropic sector and $a$ and $b$ are independent unknown
7981: coefficients to be determined by the boundary conditions. We would
7982: now like to derive the remaining $m=\pm1$, and $m=\pm2$ components
7983: $$
7984: S_{2m}^{\alpha\beta}=
7985: \sum_q {a_{q,2, m}r^{\zeta_2^{(2)}}B_{q,2, m}^{\alpha\beta} (\bf {\hat
7986: r}
7987: )}
7988: $$
7989: As usual the $q$
7990: label denotes the different possible ways of arriving at the
7991: same $j$ and runs over all such terms with the same parity and
7992: symmetry (a consequence of homogeneity and hence the constraint
7993: of incompressibility).
7994: In all that follows, we work
7995: closely with the procedure outlined in \cite{ara99b}. Following the
7996: convention in \cite{ara99b} the $q$'s to sum over are
7997: $q=\{1,7,9,5\}$. The incompressibility condition $\partial_\alpha
7998: u^\alpha = 0$ coupled with homogeneity can be used to give
7999: relations between the $a_{q,jm}$ for a given $(j,m)$. That is,
8000: for $j=2$, $m=-2\dots 2$ \begin{eqnarray} (\zeta_2^{(2)} -
8001: 2)a_{1,2, m} + 2(\zeta_2^{(2)} - 2)a_{7,2 m} + (\zeta_2^{(2)} +2)a_{9,2, m}
8002: &=& 0 \\ \nonumber a_{1,2, m} + (\zeta_2^{(2)} +
8003: 3)a_{7,2, m} + \zeta_2^{(2)}a_{5,2,m} &=& 0.
8004: \end{eqnarray}
8005: We solve the above equations in order to obtain $a_{5,2,m}$ and
8006: $a_{7,2m}$ in terms of linear combinations of $a_{1,2,m}$ and
8007: $a_{9,2m}$. \begin{eqnarray} a_{5,2,m} &=&
8008: {a_{1,2m}([\zeta_2^{(2)}]^2 - \zeta_2^{(2)} - 2) + a_{9,2,m}
8009: ([\zeta_2^{(2)}]^2 + 5 \zeta_2^{(2)} + 6) \over
8010: 2\zeta_2^{(2)}(\zeta_2^{(2)} - 2)} \\ \nonumber a_{7,2,m} &=&
8011: {a_{1,2m}(2-\zeta_2^{(2)}) - a_{9,2,m}(2+\zeta_2^{(2)}) \over
8012: 2(\zeta_2^{(2)} - 2)}.
8013: \end{eqnarray}
8014: Using the above constraints on the coefficients, we are now left
8015: with a linear combination of just two linearly independent tensor
8016: forms {\em for each m}
8017: \begin{eqnarray}\label{genl-s2m}
8018: S^{\alpha\beta}_{2m} &=&
8019: a_{9,2,m}r^{\zeta_2^{(2)}}[-\zeta_2^{(2)}(2+\zeta_2^{(2)})
8020: B_{7,2,m}^{\alpha\beta}({\B {\hat r}}) +
8021: 2\zeta_2^{(2)}(\zeta_2^{(2)} - 2)
8022: B_{9,2,m}^{\alpha\beta}({\B {\hat r}}) \nonumber \\
8023: &+&([\zeta_2^{(2)}]^2+5\zeta_2^{(2)}+6)B_{5,2,m}^{\alpha\beta}({\B {\hat
8024: r}})] \nonumber\\
8025: &+&a_{1,2,m}r^{\zeta_2^{(2)}}[2\zeta_2^{(2)}(\zeta_2^{(2)} - 2)
8026: B_{1,2,m}^{\alpha\beta}({\B {\hat r}}) -
8027: \zeta_2^{(2)}(\zeta_2^{(2)}-2)B_{7,2,m}^{\alpha\beta}({\B {\hat
8028: r}})
8029: \nonumber\\
8030: &+&([\zeta_2^{(2)}]^2-\zeta_2^{(2)}-2)B_{5,2,m}^{\alpha\beta}({\B
8031: {\hat r}})]. \end{eqnarray}
8032:
8033: The task remains to find the explicit form of the basis tensor
8034: functions $B_{q,2,m}^{\alpha\beta}({\B {\hat r}})$,
8035: $q\in\{1,7,9,5\}$, $m\in\{\pm1,\pm2\}$
8036: \\
8037: $\bullet$\,$ B_{1,2,m}^{\alpha\beta}({\B {\hat r}}) \equiv
8038: r^{-2}\delta^{\alpha\beta}r^j Y_{2m}({\B {\hat r}})$ \\
8039: $\bullet$\,$ B_{7,2,m}^{\alpha\beta}({\B {\hat r}}) \equiv
8040: r^{-2}[r^\alpha \partial^\beta +
8041: r^\beta \partial^\alpha]r^2 Y_{2m}({\B {\hat r}})$ \\
8042: $\bullet$\,$ B_{9,2,m}^{\alpha\beta}({\B {\hat r}}) \equiv r^{-4}r^\alpha
8043: r^\beta r^2 Y_{2m}({\B {\hat r}})$ \\
8044: $\bullet$\,$ B_{5,2,m}^{\alpha\beta}({\B {\hat r}}) \equiv
8045: \partial^\alpha \partial^\beta r^2 Y_{jm}({\B {\hat r}})$
8046:
8047: We obtain the $m=\{\pm1,\pm2\}$ basis functions in the following
8048: derivation. We first note that it is more convenient to form a
8049: real basis from the $ r^2 Y_{2m}({\B {\hat r}})$ since we
8050: ultimately wish to fit to real quantities and extract real
8051: best-fit parameters. We therefore form the $r^2 {\widetilde
8052: Y}_{2k}({\B {\hat r}})$ ($k= -1,0,1$) as follows:
8053: \begin{eqnarray}
8054: r^2 {\widetilde Y}_{2\; 0}(\hat {\B r})&=&r^2 Y_{2\;0}(\hat {\B
8055: r}) =r^2 \cos^2 \theta = r{_3}^2\nonumber \\ r^2 {\widetilde
8056: Y}_{2\;-1}(\hat {\B r}) &=&r^2 {Y_{2\;-1}(\hat {\B r}) -
8057: Y_{2\;+1}(\hat {\B r}) \over 2}\nonumber\\ &=&r^2 {(\cos\phi -
8058: i\sin\phi)\cos\theta\sin\theta + (\cos\phi +
8059: i\sin\phi)\cos\theta\sin\theta \over 2} \nonumber\\ &=&r^2
8060: \cos\theta\sin\theta\cos\phi = r_3 r_1 \nonumber\\ r^2 {\widetilde
8061: Y}_{2\;+1}(\hat {\B r}) &=&r^2 {Y_{2\; -1}(\hat {\B r}) +
8062: Y_{2\; +1}(\hat {\B r}) \over -2i} \nonumber\\ &=&r^2 {(\cos\phi
8063: - i\sin\phi)\cos\theta\sin\theta - (\cos\phi +
8064: i\sin\phi)\cos\theta\sin\theta \over -2i}\nonumber\\ &=&r^2
8065: \cos\theta\sin\theta\sin\phi = r_3 r_2\nonumber\\ r^2 {\widetilde
8066: Y}_{2\;-2}(\hat {\B r}) &=& r^2 {Y_{2\; 2}(\hat {\B r}) -
8067: Y_{2\; -2}(\hat {\B r}) \over 2i} \nonumber\\ &=& r^2
8068: {(\cos2\phi + i\sin2\phi)\sin^2\theta - (\cos2\phi -
8069: i\sin2\phi)\sin^2\theta \over 2i} \nonumber \\ &=& r^2
8070: \sin2\phi\sin^2\theta = 2r_1 r_2 \nonumber \\ r^2 {\widetilde
8071: Y}_{2\;+2}(\hat {\B r}) &=& r^2 {Y_{2\; 2}(\hat {\B r}) +
8072: Y_{2\; -2}(\hat {\B r}) \over 2} \nonumber\\ &=& r^2 {(\cos2\phi
8073: + i\sin2\phi)\sin^2\theta + (\cos2\phi - i\sin2\phi)\sin^2\theta
8074: \over 2} \nonumber \\ &=& r^2 \cos2\phi\sin^2\theta = r_1^2 -
8075: r_2^2 \end{eqnarray} This new basis of $r^2 {\tilde Y}_{2k}{(\B
8076: r)}$ is equivalent to using the $r^2 Y_{jm}{(\B r)}$ themselves
8077: as they form a complete, orthogonal (in the new $k$'s) set. We omit
8078: the normalization constants for the spherical harmonics for
8079: notational convenience. The subscripts on $r$ denote its
8080: components along the 1 ($m$), 2 ($p$) and 3 ($n$) directions.
8081: ${\B m}$ denotes the shear direction, ${\B p}$ the horizontal
8082: direction parallel to the boundary and orthogonal to the mean
8083: wind direction and ${\B n}$ the direction of the mean wind. This
8084: notation makes it simple to take the derivatives when we form the
8085: different basis tensors and the only thing to remember is that
8086: \begin{eqnarray}
8087: &&\partial^\alpha r_1 = \partial^\alpha (\B {r \cdot m}) = m^\alpha
8088: \nonumber\\ &&\partial^\alpha r_2 = \partial^\alpha (\B {r \cdot
8089: p})= p^\alpha \nonumber\\ &&\partial^\alpha r_3 = \partial^\alpha
8090: (\B {r \cdot \B n}) = n^\alpha \end{eqnarray}
8091:
8092: We use the above identities to proceed to derive the basis tensor
8093: functions
8094: \begin{eqnarray}
8095: B_{1,2, -1}^{\alpha\beta}(\B{ {\hat r}}) &=&
8096: r^{-2}\delta^{\alpha\beta} (\B { r \cdot n})(\B{ r \cdot m})
8097: \nonumber\\ B_{7,2, -1}^{\alpha\beta} (\B{ {\hat r}}) &=&
8098: r^{-2}[(r^\alpha m^\beta + r^\beta m^\alpha)(\B{ r \cdot n}) +
8099: (r^\alpha n^\beta + r^\beta n^\alpha)(\B{ r \cdot m})]
8100: \nonumber\\ B_{9,2, -1}^{\alpha\beta}(\B{ {\hat r}}) &=&r^{-2}
8101: r^\alpha r^\beta (\B{ r \cdot n})(\B{ r \cdot m})\nonumber\\
8102: B_{5,2, -1}^{\alpha\beta}(\B{ {\hat r}}) &=&n^\alpha m^\beta +
8103: n^\beta m^\alpha \nonumber\\ B_{1,2, 1}^{\alpha\beta}(\B{ {\hat
8104: r}}) &=& r^{-2}\delta^{\alpha\beta}(\B{ r \cdot n})(\B{ r \cdot
8105: p})\nonumber\\ B_{7,2, 1}^{\alpha\beta}(\B{ {\hat r}}) &=&
8106: r^{-2}[(r^\alpha p^\beta + r^\beta p^\alpha)(\B{ r \cdot n}) +
8107: (r^\alpha n^\beta + r^\beta n^\alpha)(\B{ r \cdot p})]
8108: \nonumber\\ B_{9,2, 1}^{\alpha\beta}(\B{ {\hat r}}) &=&r^{-2}
8109: r^\alpha r^\beta (\B{ r \cdot n})(\B{ r \cdot p})\nonumber\\
8110: B_{5,2,1}^{\alpha\beta}(\B{ {\hat r}}) &=& n^\alpha p^\beta +
8111: n^\beta p^\alpha \nonumber\\ B_{1,2, -2}^{\alpha\beta}(\B{ {\hat
8112: r}}) &=&2 r^{-2}\delta^{\alpha\beta}(\B{ r \cdot m})(\B{ r
8113: \cdot p})\nonumber\\ B_{7,2, -2}^{\alpha\beta}(\B{ {\hat r}})
8114: &=&2 r^{-2}[(r^\alpha p^\beta + r^\beta p^\alpha)(\B{ r \cdot
8115: m}) + (r^\alpha m^\beta + r^\beta m^\alpha)(\B{ r \cdot
8116: p})]\nonumber \\ B_{9,2, -2}^{\alpha\beta}(\B{ {\hat r}}) &=& 2
8117: r^{-2} r^\alpha r^\beta (\B{ r \cdot m})(\B{ r \cdot
8118: p})\nonumber\\ B_{5,2,-2}^{\alpha\beta}(\B{ {\hat r}}) &=& 2
8119: (m^\alpha p^\beta + m^\beta p^\alpha) \nonumber\\ B_{1,2,
8120: 2}^{\alpha\beta}(\B{ {\hat r}}) &=&
8121: r^{-2}\delta^{\alpha\beta}[(\B{ r \cdot m})^2 - (\B{ r \cdot
8122: p})^2] \nonumber\\ B_{7,2, 2}^{\alpha\beta}(\B{ {\hat r}}) &=&2
8123: r^{-2}[(r^\alpha m^\beta + r^\beta m^\alpha)(\B{ r \cdot m}) -
8124: (r^\alpha p^\beta + r^\beta p^\alpha)(\B{ r \cdot p})]\nonumber
8125: \\ B_{9,2, 2}^{\alpha\beta}(\B{ {\hat r}}) &=& r^{-2} r^\alpha
8126: r^\beta [(\B{ r \cdot m})^2 - (\B{ r \cdot p})^2] \nonumber\\
8127: B_{5,2,2}^{\alpha\beta}(\B{ {\hat r}}) &=& 2 (m^\alpha m^\beta -
8128: p^\alpha p^\beta) \end{eqnarray}
8129:
8130: Note that for each dimension $k$ the tensor is bilinear in some
8131: combination of two basis vectors from the set $\B{ m}$, $\B{ p}$
8132: and $\B{ n}$.
8133: %{bf. can we say something here about why this is
8134: %so... using shear ${\partial u^\alpha over \partial r^\beta}$
8135: %etc...??}
8136: Substituting these tensors forms into Eq.
8137: \ref{genl-s2m} we obtain the full tensor forms for the $j=2$
8138: non-axisymmetric terms, with two independent coefficients for
8139: each $k$.
8140: \begin{eqnarray}\label{ktens}
8141: S^{\alpha\beta}_{j=2,k=-1}(\B{ r})
8142: &=&a_{9,2,-1}r^{\zeta_2^{(2)}}{\Big
8143: [}-\zeta_2^{(2)}(2+\zeta_2^{(2)}) r^{-2}[(r^\alpha m^\beta +
8144: r^\beta m^\alpha)(\B{ r \cdot n})\nonumber\\ &+& (r^\alpha
8145: n^\beta + r^\beta n^\alpha)(\B{ r \cdot m})] +
8146: 2\zeta_2^{(2)}(\zeta_2^{(2)} - 2) r^{-4} r^\alpha r^\beta (\B{ r
8147: \cdot n})(\B{ r \cdot m})\nonumber\\
8148: &+&([\zeta_2^{(2)}]^2+5\zeta_2^{(2)}+6)(n^\alpha m^\beta +
8149: n^\beta m^\alpha) {\Big ]}\nonumber\\&+&
8150: a_{1,2,-1}r^{\zeta_2^{(2)}}{\Big [}2\zeta_2^{(2)}(\zeta_2^{(2)} -
8151: 2) r^{-2}\delta^{\alpha\beta}(\B{ r \cdot n})(\B{ r \cdot m})
8152: \nonumber\\ &-& \zeta_2^{(2)}(\zeta_2^{(2)}-2) r^{-2}[(r^\alpha
8153: m^\beta + r^\beta m^\alpha)(\B{ r \cdot n})
8154: + (r^\alpha n^\beta + r^\beta n^\alpha)(\B{ r \cdot m})] \nonumber\\
8155: &+&([\zeta_2^{(2)}]^2-\zeta_2^{(2)}-2)(n^\alpha m^\beta + n^\beta m^\alpha)
8156: {\Big ]} \nonumber\\
8157: S^{\alpha\beta}_{j=2,k=1}(\B{ r})
8158: &=&a_{9,2,1}r^{\zeta_2^{(2)}}{\Big
8159: [}-\zeta_2^{(2)}(2+\zeta_2^{(2)}) r^{-2}[(r^\alpha p^\beta +
8160: r^\beta p^\alpha)(\B{ r \cdot n})\nonumber\\ &+& (r^\alpha
8161: n^\beta + r^\beta n^\alpha)(\B{ r \cdot p})] +
8162: 2\zeta_2^{(2)}(\zeta_2^{(2)} - 2)
8163: r^{-4} r^\alpha r^\beta (\B{ r \cdot n})(\B{ r \cdot p})\nonumber\\
8164: &+&([\zeta_2^{(2)}]^2+5\zeta_2^{(2)}+6)(n^\alpha p^\beta + n^\beta p^\alpha)
8165: {\Big ]}\nonumber\\
8166: &+&a_{1,2,1}r^{\zeta_2^{(2)}}{\Big [}2\zeta_2^{(2)}(\zeta_2^{(2)}
8167: - 2)
8168: r^{-2}\delta^{\alpha\beta}(\B{ r \cdot n})(\B{ r \cdot p}) \nonumber\\ &-&
8169: \zeta_2^{(2)}(\zeta_2^{(2)}-2) r^{-2}[(r^\alpha p^\beta + r^\beta p^\alpha)
8170: (\B{ r \cdot n}) + (r^\alpha n^\beta + r^\beta n^\alpha)(\B{ r \cdot p})]
8171: \nonumber\\
8172: &+&([\zeta_2^{(2)}]^2-\zeta_2^{(2)}-2)(n^\alpha p^\beta + n^\beta p^\alpha)
8173: {\Big ]}\nonumber\\
8174: S^{\alpha\beta}_{j=2,k=-2}(\B{ r})
8175: &=&a_{9,2,-2}r^{\zeta_2^{(2)}}{\Big
8176: [}-2\zeta_2^{(2)}(2+\zeta_2^{(2)}) r^{-2}[(r^\alpha p^\beta +
8177: r^\beta p^\alpha)(\B{ r \cdot m})\nonumber\\ &+& (r^\alpha
8178: m^\beta + r^\beta m^\alpha)(\B{ r \cdot p})] +
8179: 2\zeta_2^{(2)}(\zeta_2^{(2)} - 2) r^{-4} r^\alpha r^\beta (\B{ r
8180: \cdot p})(\B{ r \cdot m}) \nonumber\\ &+&
8181: ([\zeta_2^{(2)}]^2+5\zeta_2^{(2)}+6)(m^\alpha p^\beta + m^\beta
8182: p^\alpha) {\Big ]}\nonumber\\&+& a_{1,2,-2}r^{\zeta_2^{(2)}}{\Big
8183: [}2\zeta_2^{(2)}(\zeta_2^{(2)} - 2)
8184: r^{-2}\delta^{\alpha\beta}(\B{ r \cdot m})(\B{ r \cdot p}) \nonumber\\
8185: &-&2\zeta_2^{(2)}(\zeta_2^{(2)}-2) r^{-2}[(r^\alpha p^\beta + r^\beta
8186: p^\alpha) (\B{ r \cdot m}) + (r^\alpha m^\beta + r^\beta m^\alpha)(\B{ r
8187: \cdot p})] \nonumber\\
8188: &+&2([\zeta_2^{(2)}]^2-\zeta_2^{(2)}-2)(m^\alpha p^\beta + m^\beta p^\alpha)
8189: {\Big ]}\nonumber\\
8190: S^{\alpha\beta}_{j=2,k=2}(\B{ r})
8191: &=&a_{9,2,2}r^{\zeta_2^{(2)}}{\Big
8192: [}-2\zeta_2^{(2)}(2+\zeta_2^{(2)}) r^{-2}[(r^\alpha m^\beta +
8193: r^\beta m^\alpha)(\B{ r \cdot m})\nonumber\\ &-& (r^\alpha
8194: p^\beta + r^\beta p^\alpha)(\B{ r \cdot p})] +
8195: 2\zeta_2^{(2)}(\zeta_2^{(2)} - 2) r^{-4} r^\alpha r^\beta [(\B{
8196: r \cdot m})^2-(\B{ r \cdot p})^2]\nonumber\\
8197: &+&2([\zeta_2^{(2)}]^2+5\zeta_2^{(2)}+6)(m^\alpha m^\beta -
8198: p^\beta p^\alpha) {\Big ]}\nonumber\\&+&
8199: a_{1,2,2}r^{\zeta_2^{(2)}}{\Big [}2\zeta_2^{(2)}(\zeta_2^{(2)} -
8200: 2)
8201: r^{-2}\delta^{\alpha\beta}[(\B{ r \cdot m})^2-(\B{ r \cdot
8202: p})^2]\nonumber\\ &-&2\zeta_2^{(2)}(\zeta_2^{(2)}-2) r^{-2}[(r^\alpha
8203: m^\beta + r^\beta m^\alpha) (\B{ r \cdot m}) - (r^\alpha p^\beta + r^\beta
8204: p^\alpha)(\B{ r \cdot p})] \nonumber\\
8205: &+&2([\zeta_2^{(2)}]^2-\zeta_2^{(2)}-2)(m^\alpha m^\beta -
8206: p^\beta p^\alpha) {\Big ]}
8207: \end{eqnarray}
8208: Now we want to use this form to fit for the scaling exponent
8209: $\zeta_2^{(2)}$ in the structure function $S^{33}(\B r)$ from
8210: data set I where $\alpha=\beta=3$ and the azimuthal angle of $\B
8211: r$ in the geometry is $\phi = \pi/2$.
8212: \begin{eqnarray}
8213: &S&^{33}_{j=2,k=-1}(r,\theta,\pi/2)=0\nonumber\\
8214: &S&^{33}_{j=2,k=1}(r,\theta,\pi/2)
8215: =a_{9,2,1}r^{\zeta_2^{(2)}}[-2\zeta_2^{(2)}(\zeta_2^{(2)}+2)
8216: \sin\theta\cos\theta \nonumber +
8217: 2\zeta_2^{(2)}(\zeta_2^{(2)}-2)\cos^3\theta\sin\theta]
8218: \\ &S&^{33}_{j=2,k=-2}(r,\theta,\pi/2)=0\nonumber \\
8219: &S&^{33}_{j=2,k=2}(r,\theta,\pi/2)
8220: =a_{9,2,2}r^{\zeta_2^{(2)}}[-2\zeta_2^{(2)}(\zeta_2^{(2)}-2)
8221: \cos^2\theta\sin^2\theta]
8222: +a_{1,2,2}r^{\zeta_2^{(2)}}[-2\zeta_2^{(2)}(\zeta_2^{(2)}-2)\sin^2\theta]
8223: \nonumber
8224: \end{eqnarray}
8225: We see that choosing a particular geometry eliminates certain
8226: tensor contributions. In the case of set I we are left with 3
8227: independent coefficients for $m\ne0$, the 2 coefficients from the
8228: $m=0$ contribution (Eq. \ref{m0}), and the single coefficient
8229: from the isotropic sector \ref{Siso}, giving a total of 6 fit
8230: parameters. The general forms in \ref{ktens} can be used along
8231: with the $k=0$ (axisymmetric) contribution \ref{Siso} to fit to
8232: any second order tensor object. For convenience, the table shows
8233: the number of independent coefficients that a few different
8234: experimental geometries we have will allow in the $j=2$ sector.
8235: It must be kept in mind that these forms are to be used {\em
8236: only} when there is known to be homogeneity. If there is
8237: inhomogeneity, then we cannot apply the incompressibility
8238: condition to provide constraints in the various parity and
8239: symmetry sectors and we must in general mix different parity
8240: objects, using only the geometry of the experiment itself to
8241: eliminate any terms.
8242: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8243: \begin{table}
8244: \begin{tabular}{|c|c|c|c|c|c|c|c|c|}
8245: \hline
8246: &\multicolumn{2}{c|}{$\phi=\pi/2,\alpha=\beta=3$} &
8247: \multicolumn{2}{c|}{$\phi = 0,\alpha=\beta=3$} &
8248: \multicolumn{2}{c|}{$\phi = 0,\alpha=\beta=1$}&
8249: \multicolumn{2}{c|}{$\phi = 0,\alpha=3,\beta=1$}\\ \cline {2-9}
8250: $k$&$\theta \ne 0$ &$\theta = 0$ & $\theta \ne 0$ &$ \theta = 0$
8251: & $\theta \ne 0$ & $\theta = 0$ & $\theta \ne 0 $&$\theta = 0$ \\
8252: \hline 0 & 2 & 2 & 2 & 2 & 2 & 2 & 2 & 0 \\ \hline -1 & 0 & 0
8253: & 1 & 0 & 1 & 0 & 2 & 2 \\ \hline 1 & 1 & 0 & 0 & 0 & 0 & 0 & 0
8254: & 0 \\ \hline -2) & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\ \hline 2 & 2
8255: & 0 & 2 & 0 & 2 & 2 & 2 & 0 \\ \hline \hline
8256: Total & 5 & 2 & 5 & 2 & 5 & 4 & 6 & 2 \\
8257: \hline
8258: \end{tabular}
8259: \caption{The number of free coefficients in the $j=2$ sector for
8260: homogeneous turbulence and for different geometries}
8261: \end{table}
8262: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8263:
8264: %
8265: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8266: \section{The j=1 Component in the Inhomogeneous Case}
8267: \label{app:fullj1}
8268: \subsection{Antisymmetric Contribution}
8269: We consider the tensor
8270: $$
8271: T^{\alpha\beta}(\B{ r}) = \langle u^\alpha({\B x} + {\B r}) -
8272: u^\alpha({\B x})) (u^\beta({\B x} + {\B r}) + u^\beta({\B
8273: x}))\rangle. $$ This object is trivially zero for
8274: $\alpha=\beta$. In the experimental setup, we measure at points
8275: separated in the shear direction and therefore have inhomogeneity
8276: which makes the object of mixed parity and symmetry. We cannot
8277: apply the incompressibility condition in same parity/symmetry
8278: sectors as before to provide constraints. We must in general use
8279: all 7 irreducible tensor forms. This would mean fitting for $7 \times
8280: 3 = 21$ independent coefficients plus 1 exponent $\zeta_1^{(2)}$
8281: in the anisotropic sector, together with 2 coefficients in the
8282: isotropic sector. In order to pare down the number of parameter
8283: we are fitting for, we look at the antisymmetric part of
8284: $T^{\alpha\beta}({\B r})$
8285: $$
8286: {\widetilde T}^{\alpha\beta}({\B r}) = {T^{\alpha\beta}({\B r})
8287: - T^{\beta\alpha}({\B r}) \over 2} = \langle u^\alpha({\B
8288: x})u^\beta({\B x} + {\B r})\rangle - \langle u^\beta({\B
8289: x})u^\alpha({\B x} + {\B r})\rangle $$
8290: which will only have contributions from the antisymmetric $j=1$ basis
8291: tensors. These are \\
8292: $\bullet$ Antisymmetric, odd parity
8293: \begin{equation}
8294: B_{3,1,m}^{\alpha\beta} = r^{-1}[r^\alpha\partial^\beta-
8295: r^\beta\partial^\alpha] rY_{1,m}(\B {\hat r})
8296: \end{equation}
8297: $\bullet$ Antisymmetric, even parity
8298: \begin{eqnarray}
8299: B_{4,1,m}^{\alpha\beta} &=& r^{-2}\epsilon^{\alpha\beta\mu}r_\mu
8300: r Y_{1,m}(\B {\hat r}) \nonumber\\ B_{2,1,m}^{\alpha\beta} &=&
8301: r^{-2}\epsilon^{\alpha\beta\mu}\partial_\mu r Y_{1,m}(\B {\hat
8302: r}) \end{eqnarray}
8303: As with the $j=2$ case we form a real basis $r {\tilde
8304: Y}_{1,k}(\B {\hat r})$ from the (in general) complex $r
8305: Y_{1,m}(\B {\hat r})$ in order to obtain real coefficients in
8306: the fits.
8307: \begin{eqnarray}
8308: r {\tilde Y}_{1,k=0}(\B {\hat r}) &=& r Y_{1,0}(\B {\hat r})= r\cos\theta
8309: = r_3 \nonumber\\
8310: r {\tilde Y}_{1,k=1}(\B {\hat r}) &=& r {Y_{1,1}(\B {\hat r})
8311: +Y_{1,1} (\B {\hat r}) \over 2i }
8312: =r\sin\theta\sin\phi = r_2 \nonumber\\ r {\tilde
8313: Y}_{1,k=-1}(\B {\hat r}) &=& r
8314: {Y_{1,-1}(\B {\hat r}) - Y_{1,1}(\B {\hat r}) \over 2 }
8315: =r\sin\theta\cos\phi = r_1 \nonumber
8316: \end{eqnarray}
8317: And the final forms are
8318: \begin{eqnarray}
8319: B_{3,1,0}^{\alpha\beta}(\B {\hat r}) &=& r^{-1}[r^\alpha
8320: n^\beta- r^\beta n^\alpha] \nonumber\\
8321: B_{4,1,0}^{\alpha\beta}(\B {\hat r})
8322: &=&r^{-2}\epsilon^{\alpha\beta\mu}r_\mu (\B{ r \cdot n}) \nonumber\\
8323: B_{2,1,0}^{\alpha\beta}(\B {\hat r}) &=&
8324: r^{-2}\epsilon^{\alpha\beta\mu}n_\mu \nonumber\\
8325: B_{3,1,1}^{\alpha\beta}(\B {\hat r}) &=& r^{-1}[r^\alpha
8326: p^\beta- r^\beta p^\alpha] \nonumber\\
8327: B_{4,1,1}^{\alpha\beta}(\B {\hat r})
8328: &=&r^{-2}\epsilon^{\alpha\beta\mu}r_\mu (\B{ r \cdot p}) \nonumber\\
8329: B_{2,1,1}^{\alpha\beta}(\B {\hat r}) &=&
8330: r^{-2}\epsilon^{\alpha\beta\mu}p_\mu \nonumber\\
8331: B_{3,1,-1}^{\alpha\beta}(\B {\hat r}) &=& r^{-1}[r^\alpha
8332: m^\beta- r^\beta m^\alpha] \nonumber\\
8333: B_{4,1,-1}^{\alpha\beta}(\B {\hat r})
8334: &=&r^{-2}\epsilon^{\alpha\beta\mu}r_\mu (\B{ r \cdot m}) \nonumber\\
8335: B_{2,1,-1}^{\alpha\beta}(\B {\hat r}) &=&
8336: r^{-2}\epsilon^{\alpha\beta\mu}m_\mu \end{eqnarray}
8337: Note: For a given $k$ the representations is symmetric about a
8338: particular axis in the coordinate system chosen (1=m (shear), 2=p
8339: (horizontal), 3=n (mean-wind))
8340: We now have 9 independent terms and we cannot apply
8341: incompressibility in order to reduce the number of independent
8342: coefficients in the fitting procedure. We use the geometric
8343: constraints of the experiment to do this.\\
8344: $\bullet$ $\phi = 0$ (vertical separation), $\alpha = 3, \beta =
8345: 3$ \begin{eqnarray} B_{3,1,0}^{31}(r,\theta,\phi=0) = -\sin\theta
8346: \nonumber
8347: \\ B_{2,1,1}^{31}(r,\theta,\phi=0) = 1 \nonumber\\
8348: B_{3,1,-1}^{31}(r,\theta,\phi=0) = \cos\theta \end{eqnarray}
8349: There are no contributions from the reflection-symmetric terms in
8350: the $j=0$ isotropic sector since these are symmetric in the
8351: indices. The helicity term in $j=0$ also doesn't contribute
8352: because of the geometry. So, to lowest order
8353: $$
8354: {\widetilde T}^{\alpha\beta}({\B r}) = {\widetilde
8355: T}_{j=1}^{\alpha\beta}({\B r})
8356: = a_{3,1,0}(r)(-\sin\theta) + a_{2,1,1}(r) + a_{3,1,-1}(r)\cos\theta
8357: $$
8358: We have 3 unknown independent coefficients and 1 unknown exponent
8359: to fit for in the data.
8360: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8361: \subsection{Symmetric Contribution}
8362: We consider the structure function
8363: $$
8364: S^{\alpha\beta}({\B r}) = \langle (u^\alpha({\B x} + {\B r}) -
8365: u^\alpha({\B x})) (u^\beta({\B x} + {\B r}) - u^\beta({\B
8366: x}))\rangle $$ in the case where we have homogeneous flow.
8367: This object is symmetric in the indices by construction, and it
8368: is easily seen that homogeneity implies even parity in $r$:
8369: $
8370: S^{\alpha\beta}({\B r}) = S^{\beta\alpha}({\B r})$ and
8371: $S^{\alpha\beta}({\B -r}) = S^{\alpha\beta}({\B r}).$
8372: We reason that this object cannot exhibit a $j=1$
8373: contribution from the $SO(3)$ representation in the following
8374: manner. Homogeneity allows us to use the incompressibility
8375: condition: $ \partial_\alpha S^{\alpha\beta} = 0$ and $ \partial_\beta
8376: S^{\alpha\beta} = 0$,
8377: separately on the basis tensors of a given parity and symmetry in
8378: order to give relationships between their coefficients. For the
8379: even parity, symmetric case we have for general $j \geq 2$ just
8380: two basis tensors and they must occur in some linear combination
8381: with incompressibility providing a constraint between the two
8382: coefficients. However, for $j=1$ we only have one such tensor in
8383: the even parity, symmetric group. Therefore, by
8384: incompressibility, its coefficient must vanish. Consequently, we
8385: cannot have a $j=1$ contribution for the even parity
8386: (homogeneous), symmetric structure function. Now, we consider
8387: the case as available in experiment when ${\B r}$ has some
8388: component in the inhomogeneous direction. Now, it is no longer
8389: true that $S^{\alpha\beta}({\B r})$ is of even parity and
8390: moreover it is also not possible to use incompressibility as
8391: above to exclude the existence of a $j=1$ contribution. We must
8392: look at all $j=1$ basis tensors that are symmetric, but not
8393: confined to even
8394: parity. These are \\
8395: $\bullet$ Odd parity, symmetric
8396: \begin{eqnarray}
8397: B_{1,1,k}^{\alpha\beta}({\B {\hat r}}) &\equiv&
8398: r^{-1}\delta^{\alpha\beta}r {\tilde Y}_{1k}({\B {\hat
8399: r}})\nonumber\\ B_{7,1,k}^{\alpha\beta}({\B {\hat r}}) &\equiv&
8400: r^{-1}[r^\alpha \partial^\beta + r^\beta \partial^\alpha]
8401: r{\tilde Y}_{1k}(\B {\hat r}) \nonumber \\
8402: B_{9,1,k}^{\alpha\beta}({\B {\hat r}}) &\equiv& r^{-3}r^\alpha
8403: r^\beta r {\tilde Y}_{1k}({\B {\hat r}})\nonumber \\
8404: B_{5,1,k}^{\alpha\beta}({\B {\hat r}}) &\equiv& r
8405: \partial^\alpha \partial^\beta r{\tilde Y}_{1k}({\B {\hat r}})
8406: \equiv 0
8407: \end{eqnarray}
8408: $\bullet$ Even parity, symmetric
8409: \begin{eqnarray}
8410: B_{8,1,k}^{\alpha\beta}({\B {\hat r}}) &\equiv& r^-2[r^\alpha
8411: \epsilon^{\beta\mu\nu} r_\mu \partial_\nu + r^\beta
8412: \epsilon^{\alpha\mu\nu} r_\mu \partial_\nu] r{\tilde Y}_{1k}(\B
8413: {\hat r}) \nonumber\\ B_{6,1,k}^{\alpha\beta}({\B {\hat r}})
8414: &\equiv& [\epsilon^{\beta\mu\nu} r_\mu \partial_\nu
8415: \partial_\alpha + \epsilon^{\beta\mu\nu} r_\mu \partial_\nu
8416: \partial_\beta] r{\tilde Y}_{1k}(\B {\hat r}) \equiv 0
8417: \end{eqnarray}
8418: We use the real basis of $r^{-1}{\tilde Y}_{1k}({\B {\hat r}})$
8419: which are formed from the $r^{-1}Y_{1m}({\B {\hat r}})$. Both
8420: $B_{5,1,k}^{\alpha\beta}({\B {\hat r}})$ and
8421: $B_{6,1,k}^{\alpha\beta}({\B {\hat r}})$ vanish because of the
8422: taking of the double derivative of an object of single power in
8423: $r$. We thus have 4 different contributions to symmetric $j=1$
8424: and each of these is of 3 dimensions $(k= -1,0,1)$ giving in
8425: general 12 terms in all. \begin{eqnarray}
8426: B_{1,1,0}^{\alpha\beta}(\B{ {\hat r}}) &=&
8427: r^{-1}\delta^{\alpha\beta}(\B{ r \cdot n})\nonumber\\
8428: B_{7,1,0}^{\alpha\beta}(\B{ {\hat r}}) &=& r^{-1}[r^\alpha
8429: n^\beta + r^\beta n^\alpha] \nonumber\\
8430: B_{9,1,0}^{\alpha\beta}(\B{ {\hat r}}) &=&r^{-3}r^\alpha r^\beta
8431: (\B{ r \cdot n})\nonumber\\ B_{8,1,0}^{\alpha\beta}(\B{ {\hat
8432: r}}) &\equiv& r^{-2}[(r^\alpha m^\beta + r^\beta m^\alpha)(\B{ r
8433: \cdot p}) -(r^\alpha p^\beta + r^\beta p^\alpha)(\B{ r \cdot
8434: m})]\nonumber \\ B_{1,1,1}^{\alpha\beta}(\B{ {\hat r}}) &=&
8435: r^{-1}\delta^{\alpha\beta}(\B{ r \cdot p})\nonumber\\
8436: B_{7,1,1}^{\alpha\beta}(\B{ {\hat r}}) &=& r^{-1}[r^\alpha
8437: p^\beta + r^\beta p^\alpha] \nonumber\\
8438: B_{9,1,1}^{\alpha\beta}(\B{ {\hat r}}) &=&r^{-3}r^\alpha r^\beta
8439: (\B{ r \cdot p}) \nonumber\\ B_{8,1,1}^{\alpha\beta}(\B{ {\hat
8440: r}}) &\equiv& r^{-2}[(r^\alpha m^\beta + r^\beta m^\alpha)(\B{ r
8441: \cdot n}) -(r^\alpha n^\beta + r^\beta n^\alpha)(\B{ r \cdot
8442: m})] \nonumber \\ B_{1,1,-1}^{\alpha\beta}(\B{ {\hat r}}) &=&
8443: r^{-1}\delta^{\alpha\beta}(\B{ r \cdot m})\nonumber\\
8444: B_{7,1,-1}^{\alpha\beta}(\B{ {\hat r}}) &=& r^{-1}[r^\alpha
8445: m^\beta + r^\beta m^\alpha] \nonumber\\
8446: B_{9,1,-1}^{\alpha\beta}(\B{ {\hat r}}) &=&r^{-3}r^\alpha
8447: r^\beta (\B{ r \cdot m}) \nonumber\\
8448: B_{8,1,-1}^{\alpha\beta}(\B{ {\hat r}}) &\equiv&
8449: r^{-2}[(r^\alpha p^\beta + r^\beta p^\alpha)(\B{ r \cdot n})
8450: -(r^\alpha n^\beta + r^\beta n^\alpha)(\B{ r \cdot p})]
8451: \end{eqnarray}
8452:
8453:
8454:
8455: These are all the possible $j=1$ contributions to the symmetric,
8456: mixed parity (inhomogeneous) structure function.
8457:
8458: For the experimental setup II, we want to analyze the
8459: inhomogeneous structure function in the case $\alpha = \beta =
8460: 3$, and azimuthal angle $\phi=0$ (which corresponds to vertical
8461: separation) and we obtain the basis tensors
8462: \begin{eqnarray}
8463: B_{1,1,0}^{33}(\theta) &=& \cos\theta \nonumber\\
8464: B_{7,1,0}^{33}(\theta) &=& 2\cos\theta \nonumber\\
8465: B_{9,1,0}^{33}(\theta) &=& \cos^3\theta \nonumber\\
8466: B_{8,1,1}^{33}(\theta) &=& -2\cos\theta\sin\theta \nonumber\\
8467: B_{1,1,-1}^{33}(\theta) &=& \sin\theta \nonumber\\
8468: B_{9,1,-1}^{33}(\theta) &=& \cos^2\theta\sin\theta
8469: \label{b15}
8470: \end{eqnarray}
8471: Table~D.1 gives the number of free coefficients in the symmetric
8472: $j=1$ sector in the fit to the inhomogeneous structure function
8473: for various geometric configurations.
8474: \begin{table}
8475: \begin{tabular}{|c|c|c|c|c|c|c|}
8476: \hline
8477: &\multicolumn{2}{c|}{$\phi= 0,\alpha=\beta=3$} &
8478: \multicolumn{2}{c|}{$\phi = 0,\alpha=\beta=1$} &
8479: \multicolumn{2}{c|}{$\phi = 0,\alpha=3,\beta=1$}\\ \cline {2-7}
8480: $k$&$\theta \ne 0$ &$\theta = 0$ & $\theta \ne 0$ &$ \theta = 0$
8481: & $\theta \ne 0$ & $\theta = 0$ \\ \hline 0 & 3 & 3 & 2 & 1 & 2
8482: & 0 \\ \hline 1 & 1 & 0 & 1 & 0 & 0 & 0 \\ \hline -1 & 2 & 0 &
8483: 3 & 0 & 2 & 1 \\ \hline \hline Total & 6 & 3 & 6 & 1 & 4 & 1\\
8484: \hline
8485: \end{tabular}
8486: \caption{The number of free coefficients in the symmetric $j=1$
8487: sector for inhomogeneous turbulence and for different
8488: geometries.} \end{table}
8489: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8490: \section{The Matrix Form of the Operator of the Linear Pressure Model}
8491: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%55555555555555555555
8492: \label{app:lp1}
8493: Using the basic identities of the $\Phi_{j m}(\B r)$ functions
8494: (see \cite{ara99b}),
8495: \begin{eqnarray*}
8496: \partial^2 \Phi_{j m}(\B r) &=& 0 \ , \\
8497: r^\mu \partial_\mu \Phi_{j m}(\B r) &=& j \Phi_{j m}(\B r) \ ,
8498: \end{eqnarray*}
8499: a short calculation yields:
8500: \begin{eqnarray}
8501: && \KO C^\alpha(\B r) \equiv
8502: K^{\mu\nu}(\B r)\partial_\mu\partial_\nu C^\alpha(\B r) =
8503: Dx^\epsilon\Big[
8504: 2c''_1 +
8505: 2(2+\epsilon)\frac{c'_1}{r} -
8506: (2+\epsilon)(j+1)(j+2)\frac{c_1}{r^2}
8507: \Big]B_{1jm}^\alpha(\B{\hat r}) \nonumber \\
8508: && \quad + Dx^\epsilon\Big[ 2c''_2 + 2(2+\epsilon)\frac{c'_2}{r}
8509: + 2(2+\epsilon)\frac{c_1}{r^2} - 2(2+\epsilon)j(j-1)\frac{c_2}{r^2}
8510: \Big]
8511: B_{2jm}^\alpha(\B{ \hat r}) \nonumber \ .
8512: \end{eqnarray}
8513: Therefore, in matrix notation, the Kraichnan operator can be
8514: written as:
8515: \begin{eqnarray}
8516: && \KO \VecII{c_1}{c_2} = 2Dr^\epsilon
8517: \MatII{1}{0}{0}{1}\VecII{c''_1}{c''_2}
8518: + 2D(2+\epsilon)r^{\epsilon-1} \MatII{1}{0}{0}{1}\VecII{c'_1}{c'_2}
8519: \nonumber \\
8520: && \quad - D(2+\epsilon)r^{\epsilon-2} \MatII{(j+1)(j+2)}{0}{-2}{j(j-1)}
8521: \VecII{c_1}{c_2} \nonumber \\
8522: && \equiv r^\epsilon \KM_2\VecII{c''_1}{c''_2} +
8523: r^{\epsilon-1}\KM_1\VecII{c'_1}{c'_2} +
8524: r^{\epsilon-2}\KM_0\VecII{c_1}{c_2} \ .
8525: \label{def:K-mat}
8526: \end{eqnarray}
8527: Letting
8528: \begin{equation}
8529: T^\alpha(\B r) = t_1(r)B_{1jm}^\alpha(\B{\hat r}) +
8530: t_2(r)B_{2jm}^\alpha(\B{ \hat r}) \ ,
8531: \end{equation}
8532: and applying a Laplacian to $\PO T^\alpha$, we get,
8533: \begin{eqnarray}
8534: \label{eq:explicit-P}
8535: && \partial^2\PO T^\alpha =
8536: \Big[ -j t''_2 + j\frac{t'_1}{r} + j(2j-1)\frac{t'_2}{r}-
8537: j(j+1)\frac{t_1}{r^2}
8538: - j(j-1)(j+1)\frac{t_2}{r^2}\Big]B_{1jm}^\alpha \nonumber\\&&+ \left[
8539: t''_2
8540: -
8541: \frac{t'_1}{r} +
8542: (2-j)\frac{t'_2}{r}\right]
8543: B_{2jm}^\alpha \ . \nonumber
8544: \end{eqnarray}
8545: Hence in matrix notation,
8546: \begin{eqnarray}
8547: \label{def:P-mat}
8548: && \partial^2\PO\VecII{t_1}{t_2} = \MatII{0}{-j}{0}{1}
8549: \VecII{t''_1}{t''_2} +
8550: \frac{1}{r}\MatII{j}{j(2j-1)}{-1}{2-j}
8551: \VecII{t'_1}{t'_2} \\&&
8552: \quad - \frac{1}{r^2}\MatII{j(j+1)}{j(j-1)(j+1)}{0}{0}
8553: \VecII{t_1}{t_2} \equiv \PM_2\VecII{t^{''}_1}{t^{''}_2} +
8554: \frac{1}{r}\PM_1
8555: \VecII{t'_1}{t'_2} + \frac{1}{r^2}\PM_0\VecII{t_1}{t_2} \ .\nonumber
8556: \end{eqnarray}
8557: Now that the matrix forms of the Kraichnan operator and of the
8558: Laplacian of the projection operator have been found, we can
8559: combine these two results to find the matrix form of the LHS of
8560: \Eq{eq:diff-C}. To this aim let us define,
8561: $$
8562: \VecII{t_1}{t_2} = \KO \VecII{c_1}{c_2} \ ,$$
8563: and from Eq.~(\ref{def:K-mat},\ref{def:P-mat}) we get,
8564: \begin{eqnarray}
8565: && \partial^2 \PO \KO \VecII{c_1}{c_2} =
8566: r^\epsilon \MM_4 \VecII{c^{(4)}_1}{c^{(4)}_2} +
8567: r^{\epsilon-1} \MM_3 \VecII{c^{(3)}_1}{c^{(3)}_2} +\ r^{\epsilon-2}
8568: \MM_2
8569: \VecII{c^{(2)}_1}{c^{(2)}_2}
8570: \nonumber\\&&+ r^{\epsilon-3} \MM_1 \VecII{c^{(1)}_1}{c^{(1)}_2} +\
8571: r^{\epsilon-4} \MM_0
8572: \VecII{c_1}{c_2} \ ,
8573: \nonumber
8574: \end{eqnarray}
8575: where the number in parenthesis denotes the order of the
8576: derivative. The matrices $\MM_i$ are given by:
8577: \begin{eqnarray}
8578: \label{def:M-mat}
8579: \MM_4 &\equiv& \PM_2\KM_2 \ , \\
8580: \MM_3 &\equiv& 2\epsilon\PM_2\KM_2 + \PM_2\KM_1 + \PM_1\KM_2 \ , \nonumber
8581: \\
8582: \MM_2 &\equiv& \epsilon(\epsilon-1)\PM_2\KM_2 + 2(\epsilon-1)\PM_2\KM_1 +
8583: \PM_2\KM_0
8584: \nonumber \\
8585: && \quad \ + \epsilon\PM_1\KM_2 + \PM_1\KM_1 + \PM_0\KM_2 \ ,
8586: \nonumber \\
8587: \MM_1 &\equiv& (\epsilon-1)(\epsilon-2)\PM_2\KM_1 + 2(\epsilon-2)\PM_2\KM_0
8588: \nonumber \\
8589: && \quad \ + (\epsilon -1)\PM_1\KM_1 +\PM_1\KM_0 + \PM_0\KM_1 \
8590: ,\nonumber
8591: \\
8592: \MM_0 &\equiv& (\epsilon-2)(\epsilon-3)\PM_2\KM_0 + (\epsilon-2)\PM_1\KM_0
8593: + \PM_0\KM_0 \ . \nonumber
8594: \end{eqnarray}
8595: To find the RHS of \Eq{eq:diff-C} we expand the ``forcing''
8596: $A^\alpha(\B r)$ in terms of the spherical vectors $\B B_{1jm}, \B
8597: B_{2jm}$,
8598: \begin{equation}
8599: A^\alpha(\B r) = f_1(r) B^\alpha_{1jm}(\B{ \hat r}) +
8600: f_2(r) B^\alpha_{2jm}(\B{ \hat r}) \ ,
8601: \end{equation}
8602: and applying a Laplacian we find the matrix form of $\partial^2
8603: A^\alpha(\B r)$:
8604: \begin{eqnarray}
8605: && \partial^2 \VecII{f_1}{f_2} =
8606: \left( \begin{array}{c}
8607: f''_1 + \frac{2}{r}f'_1 - (j+1)(j+2)\frac{1}{r^2}f_1 \\ \ \\
8608: f''_2 + \frac{2}{r}f'_2 + \frac{2}{r^2}f_1 - j(j-1)\frac{1}{r^2}f_2
8609: \end{array} \right) \equiv \VecII{\rho_1}{\rho_2} \ .
8610: \end{eqnarray}
8611: At this point it is worthwhile to remember that the forcing term
8612: $A^\alpha(\B r/L)$ is assumed to be analytic. As a result for
8613: $r/L \ll 1$ its leading contribution in the $(j,m)$ sector is
8614: proportional to $\partial^\alpha r^j Y_{j m}(\hat{\B r})
8615: \sim r^{j-1}$. However $\partial^2 A^{\alpha}(\B r/L)$ is also
8616: analytic, and must therefore also scale like $r^{j-1}$ for
8617: small $r$, instead of like $r^{j-3}$ which could be the naive
8618: dimensional guess.
8619:
8620: To proceed we restrict ourselves to finding the solution in the
8621: inertial range and beyond. In these ranges the dissipative term
8622: $\kappa
8623: \partial^2\partial^2 C^\alpha(\B r)$ is negligible and can be omitted,
8624: thus reaching Eq.~(\ref{eq:C-mat}) for $c_1(r)$ and $c_2(r)$.
8625:
8626: %\bibliography{literatur}
8627: \begin{thebibliography}{100}
8628:
8629: \bibitem{my75}
8630: A.~S. Monin and A.~M. Yaglom, {\em Statistical Fluid Mechanics} (The MIT Press,
8631: Cambridge, Massachusetts, 1975).
8632:
8633: \bibitem{tay35}
8634: G.~I. Taylor, Proc.~Roy.~Soc.~A {\bf 151}, 421 (1935).
8635:
8636: \bibitem{kar38}
8637: T.~D. K\'arm\'an and L. Howarth, Proc.~Roy.~Soc.~A {\bf 164}, 192 (1938).
8638:
8639: \bibitem{kol41}
8640: A.~N. Kolmogorov, CR. Acad. Sci. USSR. {\bf 30}, 299 (1941).
8641:
8642: \bibitem{fri95}
8643: U. Frisch, {\em Turbulence: The legacy of A.N. Kolmogorov} (Cambridge
8644: University Press, Cambridge, 1995).
8645:
8646: \bibitem{hin75}
8647: J.~O. Hinze, {\em Turbulence} (McGraw-Hill, New York, 1975).
8648:
8649: \bibitem{she00}
8650: X. Shen and Z. Warhaft, Phys.~Fluids {\bf 12}, 2976 (2000).
8651:
8652: \bibitem{lum67}
8653: J.~L. Lumley, Phys. Fluids {\bf 10}, 855 (1967).
8654:
8655: \bibitem{sad94}
8656: S.~G. Saddoughi and S.~V. Veeravalli, J. Fluid Mech. {\bf 268}, 333 (1994).
8657:
8658: \bibitem{kur00}
8659: S. Kurien and K.~R. Sreenivasan, Phys. Rev. E {\bf 62}, 2206 (2000).
8660:
8661: \bibitem{lum65}
8662: J. Lumley, Phys. Fluids {\bf 8}, 1056 (1965).
8663:
8664: \bibitem{tav81}
8665: S. Tavoularis and S. Corrsin, J. Fluid Mech. {\bf 104}, 311 (1981).
8666:
8667: \bibitem{kur00a}
8668: S. Kurien, V. L'vov, I. Procaccia, and K. Sreenivasan, Phys. Rev. E {\bf 62},
8669: 2206 (2000).
8670:
8671: \bibitem{fer00}
8672: M. Ferchichi and S. Tavoularis, Phys. Fluids {\bf 12}, 2942 (2000).
8673:
8674: \bibitem{bif02}
8675: L. Biferale, I. Daumont, A. Lanotte, and F. Toschi, Phys. Rev. E {\bf 66},
8676: 056306 (2002).
8677:
8678: \bibitem{ben02}
8679: R. Benzi, G. Amati, C.~M. Casciola, F. Toschi, and R. Piva, Phys. Fluids {\bf
8680: 11}, 1284 (1999).
8681:
8682: \bibitem{cas00}
8683: P. Gualtieri, C.~M. Casciola, R. Benzi, G. Amati, and R. Piva, Phys. Fluids
8684: {\bf 14}, 583 (2002).
8685:
8686: \bibitem{tos99}
8687: F. Toschi, G. Amati, S. Succi, R. Benzi, and R. Piva, Phys.~Rev.~Lett. {\bf
8688: 82}, 5044 (1999).
8689:
8690: \bibitem{tos00}
8691: F. Toschi, E. Leveque, and G. Ruiz-Chavarria, Phys. Rev. Lett. {\bf 85}, 1436
8692: (2000).
8693:
8694: \bibitem{pum95}
8695: A. Pumir and B.~I. Shraiman, Phys. Rev. Lett. {\bf 75}, 3114 (1995).
8696:
8697: \bibitem{pum96}
8698: A. Pumir, Phys.~Fluids {\bf 8}, 3112 (1996).
8699:
8700: \bibitem{sch02}
8701: J. Schumacher, K. Sreenivasan, and P. Yeung, Phys. Fluids {\bf 15}, 84
8702: (2003).
8703:
8704: \bibitem{ara99b}
8705: I. Arad, V. L'vov, and I. Procaccia, Phys. Rev. E {\bf 59}, 6753 (1999).
8706:
8707: \bibitem{fal01}
8708: G. Falkovich, K. Gaw\c{e}dzki, and M. Vergassola, Rev. Mod. Phys. {\bf 73},
8709: 913 (2001).
8710:
8711: \bibitem{ara00}
8712: I. Arad, V. L'vov, E. Podivilov, and I. Procaccia, Phys. Rev. E {\bf 62}, 4904
8713: (2000).
8714:
8715: \bibitem{ara00a}
8716: I. Arad, L. Biferale, and I. Procaccia, Phys. Rev. E {\bf 61}, 2654 (2000).
8717:
8718: \bibitem{lan99}
8719: A. Lanotte and A. Mazzino, Phys. Rev. E {\bf 60}, R3483 (1999).
8720:
8721: \bibitem{ara01}
8722: I. Arad and I. Procaccia, Phys. Rev. E {\bf 63}, 056302 (2001).
8723:
8724: \bibitem{ara98}
8725: I. Arad, B. Dhruva, S. Kurien, V.~S. L'vov, I. Procaccia, and K.~R.
8726: Sreenivasan, Phys. Rev. Lett. {\bf 81}, 5330 (1998).
8727:
8728: \bibitem{she02}
8729: X. Shen and Z. Warhaft, Phys. Fluids {\bf 14}, 370 (2002).
8730:
8731: \bibitem{she02a}
8732: X. Shen and Z. Warhaft, Phys.~Fluids {\bf 14}, 2432 (2002).
8733:
8734: \bibitem{wander}
8735: A. Staicu, B. Vorselaars, and W. van~de Water, Phys. Rev. E {\bf 68},
8736: (2003).
8737:
8738: \bibitem{iacob04}
8739: B. Jacob, L. Biferale, G. Iugo, and C. Casciola, Phys. Fluids (2004),
8740: submitted.
8741:
8742: \bibitem{ara99}
8743: I. Arad, L. Biferale, I. Mazzitelli, and I. Procaccia, Phys. Rev. Lett. {\bf
8744: 82}, 5040 (1999).
8745:
8746: \bibitem{bif01}
8747: L. Biferale and M. Vergassola, Phys.~Fluids {\bf 13}, 2139 (2001).
8748:
8749: \bibitem{bif01a}
8750: L. Biferale and F. Toschi, Phys. Rev. Lett. {\bf 86}, 4831 (2001).
8751:
8752: \bibitem{ish02}
8753: T. Ishihara, K. Yoshida, and Y. Kaneda, Phys. Rev. Lett. {\bf 88}, 154501
8754: (2002).
8755:
8756: \bibitem{sch00}
8757: J. Schumacher and B. Eckhardt, Europhys. Lett. {\bf 52}, 627 (2000).
8758:
8759: \bibitem{rob40}
8760: H. Robertson, Proc.~Camb.~Phil.~Soc. {\bf 36}, 209 (1940).
8761:
8762: \bibitem{chk96}
8763: O. Chkhetiani, JETP Lett. {\bf 63}, 808 (1996).
8764:
8765: \bibitem{lvo97a}
8766: V. L'vov, E. Podivilov, and I. Procaccia, chao-dyn/9705016 (1997).
8767:
8768: \bibitem{eyink1}
8769: Q. Chen, S. Chen, and G.~L. Eyink, Phys. Fluids {\bf 15}, 361 (2003).
8770:
8771: \bibitem{eyink2}
8772: Q. Chen, S. Chen, G.~L. Eyink, and D. Holm, Phys. Rev. Lett. {\bf 90}, 254501
8773: (2003).
8774:
8775: \bibitem{bif98c}
8776: L. Biferale, D. Pierotti, and F. Toschi, Phys. Rev. E {\bf 57}, R2515 (1998).
8777:
8778: \bibitem{kur03}
8779: S. Kurien, Physica D {\bf 175}, 167 (2003).
8780:
8781: \bibitem{pol03}
8782: H. Politano, T. Gomez, and A. Pouquet, Phys. Rev. E {\bf 68}, 026315 (2003).
8783:
8784: \bibitem{fri90}
8785: U. Frisch and Z. She, Fluid Dyn. Res. {\bf 8}, 139 (1991).
8786:
8787: \bibitem{ben91}
8788: R. Benzi, L. Biferale, G. Paladin, A. Vulpiani, and M. Vergassola, Phys. Rev.
8789: Lett. {\bf 67}, 2299 (1991).
8790:
8791: \bibitem{ans84}
8792: F. Anselmet, Y. Gagne, E.~J. Hopfinger, and R. Antonia, J. Fluid Mech. {\bf
8793: 140}, 63 (1984).
8794:
8795: \bibitem{kan03}
8796: Y. Kaneda, T. Ishihara, M. Yokokawa, K. Itakura, and A. Uno, Phys. Fluids {\bf
8797: 15}, L21 (2003).
8798:
8799: \bibitem{ben93b}
8800: R. Benzi, S. Ciliberto, R. Tripiccione, C. Baudet, F. Massaioli, and S. Succi,
8801: Phys. Rev. E {\bf 48}, R29 (1993).
8802:
8803: \bibitem{kol62}
8804: A.~N. Kolmogorov, J. Fluid Mech. {\bf 13}, 82 (1962).
8805:
8806: \bibitem{lvo96}
8807: V. L'vov and I. Procaccia, Phys. Rev. Lett {\bf 76}, 2898 (1996).
8808:
8809: \bibitem{ben98}
8810: R. Benzi, L. Biferale, and F. Toschi, Phys. Rev. Lett. {\bf 80}, 3244 (1998).
8811:
8812: \bibitem{gar98}
8813: S. Garg and Z. Warhaft, Phys.~Fluids {\bf 10}, 662 (1998).
8814:
8815: \bibitem{war00}
8816: Z. Warhaft, Annu. Rev. Fluid Mech. {\bf 32}, 203 (2000).
8817:
8818: \bibitem{gib77}
8819: C. Gibson, C. Friehe, and S. McConnel, Phys. Fluids {\bf 20}, S156 (1977).
8820:
8821: \bibitem{ant78}
8822: R. Antonia and C. van Atta, J. Fluid Mech. {\bf 84}, 561 (1978).
8823:
8824: \bibitem{sre91}
8825: K. Sreenivasan, Proc. R. Soc. London {\bf Ser. A 434}, 165 (1991).
8826:
8827: \bibitem{p94}
8828: A. Pumir, Phys. Fluids {\bf 6}, 6 (1994).
8829:
8830: \bibitem{cel01}
8831: A. Celani, A. Lanotte, A. Mazzino, and M. Vergassola, Phys. Fluids {\bf 13},
8832: 1768 (2001).
8833:
8834: \bibitem{che97}
8835: S. Chen, K.~R. Sreenivasan, M. Nelkin, and N. Cao, Phys. Rev. Lett. {\bf 79},
8836: 2253 (1997).
8837:
8838: \bibitem{dhr97}
8839: B. Dhruva, Y. Tsuji, and K.~R. Sreenivasan, Phys. Rev. E {\bf 56}, R4928
8840: (1997).
8841:
8842: \bibitem{got02}
8843: T. Gotoh, D. Fukayama, and T. Nakano, Phys. Fluids {\bf 14}, 1065 (2002).
8844:
8845: \bibitem{wat99}
8846: W. van~de Water and J. Herweijer, J. Fluid Mech. {\bf 387}, 3 (1999).
8847:
8848: \bibitem{zho99}
8849: T. Zhou and R. Antonia, J. Fluid Mech. {\bf 406}, 81 (1999).
8850:
8851: \bibitem{nou97}
8852: A. Noullez, G. Wallace, W. Lempert, R.~B. Miles, and U. Frisch, J. Fluid Mech.
8853: {\bf 339}, 287 (1997).
8854:
8855: \bibitem{cam96}
8856: R. Camussi, D. Barbagallo, G. Guj, and F. Stella, Phys. Fluids {\bf 8}, 1181
8857: (1996).
8858:
8859: \bibitem{bor97}
8860: O.~N. Boratav and R.~B. Pelz, Phys. Fluids {\bf 9}, 1400 (1997).
8861:
8862: \bibitem{ker01}
8863: R. Kerr, M. Meneguzzi, and T. Gotoh, Phys. Fluids {\bf 13}, 1985 (2001).
8864:
8865: \bibitem{gro98b}
8866: S. Grossmann, D. Lohse, and A. Reeh, J. Stat. Phys. {\bf 93}, 715 (1998).
8867:
8868: \bibitem{gau98}
8869: E. Gaudin, B. Protas, S. Gouion-Durand, J. Wojciechowski, and J. Wesfried,
8870: Phys. Rev. E {\bf 57}, R9 (1998).
8871:
8872: \bibitem{ono00}
8873: M. Onorato, R. Camussi, and G. Iuso, Phys. Rev. E {\bf 61}, 1447 (2000).
8874:
8875: \bibitem{str96}
8876: R. BEnzi, M. Struglia, and R. Tripiccione, Phys. Rev. E {\bf 53}, 5565
8877: (1996).
8878:
8879: \bibitem{bat46}
8880: G.~K. Batchelor, Proc.~Roy.~Sco.~Lond.~A {\bf 186}, 480 (1946).
8881:
8882: \bibitem{cha50}
8883: S. Chandrasekhar, Proc.~Roy.~Sco.~Lond.~A {\bf 242}, 557 (1950).
8884:
8885: \bibitem{lin95}
8886: E. Lindborg, J.~Fluid~Mech. {\bf 302}, 179 (1995).
8887:
8888: \bibitem{her74}
8889: J.~R. Herring, Phys.~Fluids {\bf 17}, 859 (1974).
8890:
8891: \bibitem{rot51}
8892: J.~C. Rotta, Z.~Phys. {\bf 129}, 547 (1951).
8893:
8894: \bibitem{lum77}
8895: J.~L. Lumley and G.~R. Newman, J.~Fluid~Mech. {\bf 82}, 161 (1977).
8896:
8897: \bibitem{lum70}
8898: J.~L. Lumley, {\em Stochastic tools in turbulence} (Acad.~Press, London, 1970).
8899:
8900: \bibitem{chu95}
8901: M.~K. Chung and S.~K. Kim, Phys.~Fluid {\bf 7}, 1425 (1995), and refernce
8902: therein.
8903:
8904: \bibitem{gen80}
8905: J.~N. Gence and J. Mathieu, J.~Fluid~Mech. {\bf 101}, 555 (1980).
8906:
8907: \bibitem{kra59}
8908: R.~H. Kraichnan, J. Fluid Mech. {\bf 5}, 497 (1959).
8909:
8910: \bibitem{sch76}
8911: U. Schumann and J.~R. Herring, J.~Fluid~Mech. {\bf 76}, 755 (1976).
8912:
8913: \bibitem{ors70}
8914: S.~A. Orszag, J. Fluid Mech. {\bf 41}, 363 (1970).
8915:
8916: \bibitem{ors77}
8917: S.~A. Orszag, in {\em Fluid Dynamics, Les Houches Summer school}, edited by R.
8918: Balian and J.~L. Peube (Gordon and Breach, New York, 1977), pp.\ 237--374.
8919:
8920: \bibitem{nak87}
8921: N. Nakauchi and H. Oshima, Phys.~Fluids {\bf 30}, 3653 (1987).
8922:
8923: \bibitem{sou95}
8924: F.~D. Souza, V. Nguyen, and S. Tavoularis, J. Fluid MEch. {\bf 303}, 155
8925: (1995).
8926:
8927: \bibitem{yak94}
8928: V. Yakhot, Phys. Rev. E {\bf 49}, 2887 (1994).
8929:
8930: \bibitem{gro94}
8931: S. Grossmann, D. Lohse, V. L'vov, and I. Procaccia, Phys. Rev. Lett. {\bf 73},
8932: 432 (1994).
8933:
8934: \bibitem{fal95}
8935: G. Falkovich and V. L'vov, Chaos,~Solitons~and~Fractals {\bf 5}, 1855 (1995).
8936:
8937: \bibitem{gro00}
8938: S. Grossmann, A. von~der Heydt, and D. Lohse, J. Fluid Mech. {\bf 440}, 381
8939: (2001).
8940:
8941: \bibitem{kra68}
8942: R.~H. Kraichnan, Phys.~Fluids {\bf 11}, 945 (1968).
8943:
8944: \bibitem{gaw95}
8945: K. Gaw\c{e}dzki and A. Kupiainen, Phys.~Rev.~Lett. {\bf 75}, 3834 (1995).
8946:
8947: \bibitem{che95}
8948: M. Chertkov, G. Falkovich, I. Kolokolov, and V. Lebedev, Phys.~Rev.~E {\bf 52},
8949: 4924 (1995).
8950:
8951: \bibitem{ber96}
8952: D. Bernard, K. Gaw\c{e}dzcki, and A. Kupiainen, Phys.~Rev.~E {\bf 54}, 2624
8953: (1996).
8954:
8955: \bibitem{gat}
8956: O. Gat, V.~S. L'vov, E. Podivilov, and I. Procaccia, Phys. Rev. E {\bf 55},
8957: 3836R (1997).
8958:
8959: \bibitem{fai96}
8960: A. Fairhall, O. Gat, V.~S. L'vov, and I. Procaccia, Phys.~Rev.~E {\bf 53},
8961: 3518 (1996).
8962:
8963: \bibitem{lvo96c}
8964: V.~S. L'vov and I. Procaccia, Phys.~Fluids {\bf 8}, 2565 (1996).
8965:
8966: \bibitem{tay38}
8967: G.~I. Taylor, Proc. R. Soc. London A {\bf 164}, 476 (1938).
8968:
8969: \bibitem{Corn}
8970: J. Cornwell, {\em Group Theory in Physics} (Academic, London, 1984).
8971:
8972: \bibitem{Stern}
8973: S. Sternberg, {\em Group Theory and Physics} (Canbridge University Press,
8974: Cambridge, 1994).
8975:
8976: \bibitem{bel87}
8977: V.~I. Belinicher and V.~S. L'vov, Sov. Phys. JETP {\bf 66}, 303 (1987).
8978:
8979: \bibitem{kra94}
8980: R. Kraichnan, Phys. Rev. Lett. {\bf 72}, 1016 (1994).
8981:
8982: \bibitem{ver96}
8983: M. Vergassola, Phys.~Rev.~E {\bf 53}, R3021 (1996).
8984:
8985: \bibitem{anto01}
8986: L.~T. Adzhemyan, N.~V. Antonov, M. Hnatich, and S.~V. Novikov, Phys. Rev. E
8987: {\bf 63}, 016309 (2001).
8988:
8989: \bibitem{anto01a}
8990: N.~V. Antonov and J. Honkonen, Phys. Rev. E {\bf 63}, 036302 (2001).
8991:
8992: \bibitem{anto00}
8993: N. Antonov, J. Honkonen, A. Mazzino, and P. Muratore-Ginanneschi, Phys. Rev. E
8994: {\bf 62}, R5891 (2000).
8995:
8996: \bibitem{anto99}
8997: N. Antonov, A. Lanotte, and A. Mazzino, Phys. Rev. E {\bf 61}, 6586 (2000).
8998:
8999: \bibitem{anto03}
9000: N. Antonov, M. Hnatich, J. Honkonen, and M. Jurcisin, Phys. Rev. E {\bf 68},
9001: 046306 (2003).
9002:
9003: \bibitem{lvo94c}
9004: V.~S. L'vov, I. Procaccia, and A.~L. Fairhall, Phys. Rev. E {\bf 50}, 4684
9005: (1994).
9006:
9007: \bibitem{ben97}
9008: R. Benzi, L. Biferale, and A. Wirth, Phys. Rev. Lett. {\bf 78}, 4926 (1997).
9009:
9010: \bibitem{ap01}
9011: I. Arad and I. Procaccia, IUTAM symposium, ed. T. Kambe 175 (2001).
9012:
9013: \bibitem{sra74}
9014: B.~I. Shraiman and E.~D. Siggia, Phys. Rev. E {\bf 49}, 2912 (1974).
9015:
9016: \bibitem{gat98}
9017: O. Gat and R. Zeitak, Phys. Rev. E {\bf 57}, 5331 (1998).
9018:
9019: \bibitem{gat98b}
9020: O. Gat, R. Zeitak, and I. Procaccia, Phys. Rev. Lett. {\bf 80}, 5536 (1998).
9021:
9022: \bibitem{fmnv99}
9023: U. Frisch, A. Mazzino, A. Noullez, and M. Vergassola, Phys. Fluids {\bf 11},
9024: 2178 (1999).
9025:
9026: \bibitem{cel01a}
9027: A. Celani and M. Vergassola, Phys. Rev. Lett. {\bf 86}, 424 (2001).
9028:
9029: \bibitem{ara01a}
9030: I. Arad, L. Biferale, A. Celani, I. Procaccia, and M. Vergassola, Phys. Rev.
9031: Lett. {\bf 87}, 164502 (2001).
9032:
9033: \bibitem{lvo96a}
9034: V.~S. L'vov and I. Procaccia, Phys. Rev. Lett. {\bf 77}, 3541 (1996).
9035:
9036: \bibitem{fri91}
9037: U. Frisch and M. Vergassola, Europhys. Lett. {\bf 14}, 439 (1991).
9038:
9039: \bibitem{ben96b}
9040: R. Benzi, L. Biferale, S. Ciliberto, M.~V. Struglia, and R. Tripiccione,
9041: Physica D {\bf 96}, 162 (1996).
9042:
9043: \bibitem{cel04}
9044: A. Celani, M. Cencini, M. Vergassola, E. Villermaux, and D. Vincenzi, J. Fluid
9045: Mech. (2004), submitted.
9046:
9047: \bibitem{chi95}
9048: S. Childress and A. Gilbert, {\em Twist Stretch Fold: the Fast Dynamo}
9049: (Springer, Berlin, 1995).
9050:
9051: \bibitem{lvo95c}
9052: V.~S. L'vov and I. Procaccia, Phys. Rev. E {\bf 52}, 3840 (1995).
9053:
9054: \bibitem{fuk00}
9055: D. Fukayama, T. Oyamada, T. Nakano, T. Gotoh, and K. Yamamoto, J. Phys. Soc.
9056: Japan {\bf 69}, 701 (200).
9057:
9058: \bibitem{ben01a}
9059: R. Benzi, L. Biferale, and F. Toschi, Eur. Phys. J. B {\bf 24}, 125 (2001).
9060:
9061: \bibitem{yos01}
9062: K. Yoshida and Y. Kaneda, Phys.~Rev.~E {\bf 63}, 016308 (2001).
9063:
9064: \bibitem{lvo03}
9065: V.~S. L'vov, I. Procaccia, and V. Tiberkevich, Phys. Rev. E {\bf 67}, 026312
9066: (2003).
9067:
9068: \bibitem{bif03a}
9069: L. Biferale, E. Calzavarini, F. Toschi, and R. Tripiccione, Europhys. Lett.
9070: {\bf 64}, 461 (2003).
9071:
9072: \bibitem{lpp99}
9073: V.~S. L'vov, A. Pomyalov, and I. Procaccia, Phys. Rev. E {\bf 60}, 4175
9074: (1999).
9075:
9076: \bibitem{kra67}
9077: R. Kraichnan, Phys. Fluids {\bf 10}, 2080 (1967).
9078:
9079: \bibitem{sch01}
9080: J. Schumacher, J. Fluid Mech. {\bf 441}, 109 (2001).
9081:
9082: \bibitem{sta02}
9083: A. Staicu, Ph.D. thesis, Univ. of Technology Eindhoven, 2002.
9084:
9085: \bibitem{sta03}
9086: A. Staicu and W. van~de Water, Phys. Rev. Lett. {\bf 90}, 094501 (2003).
9087:
9088: \bibitem{rog82}
9089: R. Rogallo, NASA Tech. Memo. {\bf 81315}, (1982).
9090:
9091: \bibitem{bif02a}
9092: L. Biferale, D. Lohse, I. Mazzitelli, and F. Toschi, J.~Fluid Mech. {\bf 542},
9093: 39 (2002).
9094:
9095: \bibitem{bif03b}
9096: L. Biferale, I. Daumont, A. Lanotte, and F. Toschi, Europ. Journ. Mech. {\bf B}
9097: (2004), in press.
9098:
9099: \bibitem{ama97}
9100: G. Amati, R. Benzi, and S. Succi, Phys.~Rev.~E {\bf 55}, 6985 (1997).
9101:
9102: \bibitem{ama97a}
9103: G. Amati, S. Succi, and R. Piva, Int.~Jour.~Mod.~Phys.~C {\bf 8}, 869 (1997).
9104:
9105: \bibitem{bor96}
9106: V. Borue and S.~A. Orszag, J. Fluid. Mech. {\bf 306}, 293 (1996).
9107:
9108: \bibitem{mene00}
9109: C. Meneveau and J. Katz, Annu. Rev. Fluid Mech. {\bf 32}, 1 (2000).
9110:
9111: \bibitem{gro02}
9112: S. Grossmann and D. Lohse, Phys. Rev. E {\bf 66}, 016305 (2002).
9113:
9114: \bibitem{kad01}
9115: L. Kadanoff, Physics Today {\bf 54}, 34 (2001).
9116:
9117: \bibitem{yak92}
9118: V. Yakhot, Phys. Rev. Lett. {\bf 69}, 769 (1992).
9119:
9120: \bibitem{che97b}
9121: S. Chen, K.~R. Sreenivasan, and M. Nelkin, Phys. Rev. Lett. {\bf 79}, 1253
9122: (1997).
9123:
9124: \bibitem{bat53}
9125: G.~K. Batchelor, {\em The Theory of Homogeneous Turbulence} (Cambridge
9126: University Press, Cambridge, 1953).
9127:
9128: \bibitem{skr00}
9129: L. Skrbek and S. Stalp, Phys. Fluids {\bf 12}, 1997 (2000).
9130:
9131: \bibitem{des94}
9132: I.~D. Silva and H. Fernando, Phys. Fluids {\bf 6}, 2455 (1994).
9133:
9134: \bibitem{bif03c}
9135: L. Biferale, G. Boffetta, A. Celani, A. Lanotte, F. Toschi, and M. Vergassola,
9136: Phys. Fluids {\bf 15}, 2105 (2003).
9137:
9138: \bibitem{bor95}
9139: V. Borue and S.~A. Orszag, Phys. Rev. E {\bf 51}, R856 (1995).
9140:
9141: \bibitem{62Ham}
9142: M. Hammermesh, {\em Group Theory and its Applications to Physical Problems}
9143: (Addison-Wesley, Reading, Ma, 1962).
9144:
9145: \bibitem{horvai}
9146: P. Horvai, Ph.D. thesis, Ecole Polytechnique, 2003.
9147:
9148: \end{thebibliography}
9149:
9150:
9151: %\input{biblio.tex}
9152: %\clearpage
9153: %%%%%%%%%%%%%%%%%% comment next line if you do not want figures
9154: %\input{figlist.tex}
9155:
9156:
9157:
9158: \end{document}
9159:
9160:
9161: