nlin0207005/MHD4.tex
1: \documentclass[pre,aps,superscriptaddress,showpacs,twocolumn]{revtex4}
2: \usepackage{epsfig}
3: \usepackage{bm}
4: %\input REVTeXDraft1.tex
5: 
6: \begin{document}
7: 
8: \def\B.#1{{\bm #1}}
9: \def\C.#1{{\cal #1}}
10: \title{Active and Passive Fields in Turbulent Transport: the Role of
11: Statistically Preserved
12: Structures}
13: \author{Emily S.C. Ching}
14: \affiliation{Department of Physics, The Chinese University of Hong Kong,
15: Sha Tin, Hong Kong}
16: \author{Yoram Cohen}
17: \author{Thomas Gilbert}
18: \author{Itamar Procaccia}
19: \affiliation{Dept. of Chemical Physics, The Weizmann Institute
20: of Science, Rehovot 76100, Israel}
21: \date{\today}
22: \begin{abstract}
23: We have recently proposed that the statistics of {\em active}
24: fields (which affect the velocity field itself) in
25: well-developed turbulence
26: are also dominated by the Statistically Preserved Structures
27: of auxiliary {\em passive} fields which are
28: advected by the same velocity field.
29: The Statistically Preserved Structures are eigenmodes
30: of eigenvalue 1 of an appropriate propagator of the
31: decaying (unforced) passive field, or equivalently,
32: the zero modes of a related operator.
33: In this paper we investigate
34: further this surprising finding via two examples, one akin to turbulent
35: convection in which the temperature is the active scalar, and the other
36: akin to magneto-hydrodynamics in which the magnetic field is the active
37: vector. In the first example, all the even correlation functions
38: of the active and passive fields exhibit identical
39: scaling behavior.  The second example appears at first sight to be a
40: counter-example: the statistical objects of the active and passive fields
41: have entirely different scaling exponents. We demonstrate
42: nevertheless that the Statistically Preserved Structures
43: of the passive vector dominate again the statistics of the active
44: field, except that due to a dynamical conservation law the amplitude
45: of the leading zero mode cancels exactly. The active vector is then
46: dominated by the sub-leading zero mode of the passive vector. Our work
47: thus suggests that the statistical properties of active fields in
48: turbulence can be understood with the same generality as those of passive
49: fields.
50: \end{abstract}
51: \pacs{47.27.-i}
52: \maketitle
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
54: \section{Introduction}
55: \label{intro}
56: 
57: The aim of this paper is to address the statistical physics of so called
58: ``active" fields in developed fluid turbulence. These are fields that
59: differ
60: from the
61: fundamental fluid velocity field $\B.u(\B.r,t)$, but that interact with
62: the
63: velocity field
64: in an essential way, for example effecting a significant change in the
65: scaling exponents
66: of the velocity correlation functions from the classical Kolmogorov
67: exponents.
68: For the sake of concreteness we will focus on two generic examples with
69: very different interactions between the active and the velocity fields.
70: 
71: The first is thermal turbulent convection, in which the temperature field
72: $T(\B.r,t)$ is driving the velocity field through buoyancy effects. In the
73: Boussinesq
74: approximation the temperature equation reads like a standard forced
75: scalar advection problem,
76: \begin{equation}
77: \frac{\partial T(\B.r,t)}{\partial t}\!+\!\B.u(\B.r,t)\cdot
78: \B.\nabla T(\B.r,t)\!=\!\kappa \nabla^2 T(\B.r,t)\!+\!f(\B.r,t)  .
79: \label{active}
80: \end{equation}
81: Here $\kappa$ is the thermal diffusivity
82: and $f(\B.r,t)$ is a white random force of zero mean
83: with compact support in $\B.k$-space, acting on the largest
84: scales of the order of the outer scale $L$ only.
85: The velocity field is affected by the temperature. For
86: an incompressible fluid of unit density \cite{71MY} (dropping
87: the dependence on $(\B.r,t)$ for brevity),
88: \begin{equation}
89: \frac{\partial \B.u}{\partial t}+\B.u\cdot
90: \B.\nabla \B.u=-\B.\nabla p+\nu \nabla^2 \B.u
91: +\alpha g T\hat z \ .
92: \label{Boussinesq}
93: \end{equation}
94: Here $p$, $\nu$, $\alpha$, $g$ and $\hat z$ are the pressure,
95: kinematic viscosity, volume
96: expansion coefficient, acceleration due to gravity and a unit
97: vector in the upward direction respectively. The
98: appearance of $T$ in the equation for $\B.u$ is crucial,
99: and changes the scaling exponents of $\B.u$. When the conditions
100: are right it may even change the scaling exponents from Kolmogorov 
101: to Bolgiano (up to anomalies) \cite{71MY}.
102: 
103: The second example is that of magneto-hydrodynamics (MHD), in which the
104: magnetic field $\B.b(\B.r,t)$ is driving the velocity field $\B.u(\B.r,t)$
105: according to \cite{Zeldovich}
106: \begin{eqnarray}
107: \frac{\partial \B.u}{\partial t}+\B.u\cdot
108: \B.\nabla \B.u&=&-\B.\nabla p +\B.b\cdot\B.\nabla \B.b +\nu \nabla^2 \B.u
109:  \ , \nonumber\\ 
110: \frac{\partial \B.b}{\partial t}+\B.u\cdot
111: \B.\nabla \B.b&=& \B.b\cdot\B.\nabla \B.u +\kappa \nabla^2 \B.b+\B.f \ .
112: \label{MHD}
113: \end{eqnarray}
114: These equations of motion conserve (in the inviscid, unforced limit)
115: three quadratic invariants, i.~e. the energy, magnetic helicity and cross
116: helicity \cite{93Bis}
117: 
118: Our main interest is in the properties of the statistical
119: objects characterizing the active fields, including their anomalous
120: scaling. Here ``anomalous scaling'' means that
121: multi-point correlation functions are homogeneous functions of their
122: arguments, with exponents that cannot be guessed from dimensional
123: analysis. Thus for example the field $\phi(\B.r,t)$ (with $\phi$ being $T$
124: or
125: $\B.b$ respectively) has simultaneous multi-point correlation functions
126: \begin{equation}
127: F^{(m)}(\B.r_1,\B.r_2,\!\cdots\!, \B.r_m)\equiv
128: \langle \phi(\B.r_1,t)\phi(\B.r_2,t)\!\cdots\!
129: \phi(\B.r_m,t)\rangle_f \ , \label{corr}
130: \end{equation}
131: where pointed brackets with subscript $f$ refer to
132: averaging over the statistics
133: of the advecting velocity field {\em and} of the forcing. The forcing
134: is taken to be white random noise with zero mean. When the
135: forcing is stationary in time this object is {\em time independent}.
136: Anomalous scaling means that
137: \begin{equation}
138: F^{(m)}(\lambda \B.r_1, \cdots, \lambda\B.r_m)
139: =\lambda^{\zeta_m} F^{(m)}(\B.r_1,\cdots, \B.r_m) \ , \label{zeta}
140: \end{equation}
141: with $\zeta_m$ having a non-trivial dependence on $m$. In what
142: follows we will assume that the advecting velocity field
143: itself is fully turbulent, and that its correlation functions are also
144: exhibiting scaling behavior like Eq.~(\ref{zeta}).
145: 
146: The main point of this paper is that the statistical
147: theory of the active fields calls for consideration of 
148: auxiliary passive fields that satisfy the same equations 
149: of motion as the active fields, but
150: {\em do not} affect the velocity field itself. In other words,
151: For the two problems at hand we consider the following equations of motion~:
152: \begin{eqnarray}
153: \frac{\partial \B.u}{\partial t}+\B.u\cdot
154: \B.\nabla \B.u&=&-\B.\nabla p+\nu \nabla^2 \B.u
155: +\alpha g T\hat z \ ,\nonumber\\
156: \frac{\partial T}{\partial t} + \B.u \cdot
157: \B.\nabla T &=& \kappa \nabla^2 T + f  \ ,\nonumber\\
158: \frac{\partial C}{\partial t}+\B.u\cdot
159: \B.\nabla C&=&\kappa \nabla^2 C+\tilde f \ , \label{passives}
160: \end{eqnarray}
161: on the one hand, and
162: \begin{eqnarray}
163: \frac{\partial \B.u}{\partial t}+\B.u\cdot
164: \B.\nabla \B.u&=&-\B.\nabla p +\B.b\cdot\B.\nabla \B.b +\nu \nabla^2 \B.u
165:  \ , \nonumber\\ 
166: \frac{\partial \B.b}{\partial t}+\B.u\cdot
167: \B.\nabla \B.b&=& \B.b\cdot\B.\nabla \B.u +\kappa \nabla^2 \B.b+\B.f\ ,
168: \nonumber\\
169: \frac{\partial \B.q}{\partial t}+\B.u\cdot
170: \B.\nabla \B.q&=& \B.q\cdot\B.\nabla \B.u +\kappa \nabla^2 \B.q+\tilde\B.f \
171: ,
172: \label{passivev}
173: \end{eqnarray}
174: on the other.
175: 
176: Note that the velocity field that appears in the equations for
177: the passive fields is {\em the same} as the velocity field that results
178: from solving the coupled equations of the  associated equations for
179: the active fields. The forcing terms $\tilde f$ in Eq.~(\ref{passives})
180: (resp. $\tilde\B.f$ in Eq.~(\ref{passivev})) have the same statistics as the
181: forcing
182: terms $f$ in Eq. (\ref{active})  (resp. $\B.f$ in Eq. (\ref{MHD})), but they
183: must have
184: different realizations. While it is not true of course that the statistics
185: of the passive fields are independent of the statistics of the velocity 
186: fields, it is 
187: true that the statistics of the velocity fields  are independent of the
188: statistics of the passive forcing terms. This is, however, not the case
189: with the active forcing terms since these forcing terms affect the
190: active fields that affect in their turn the velocity fields. It is thus
191: not at all evident at first sight that there should be any relation,
192: apriori,
193: between the statistics of the active fields and their passive
194: counterparts. On the other hand, if there were such a relationship,
195: this would be very advantageous, since the statistics of the
196: passive fields is understood as explained next.
197: 
198: To understand the progress made in the context of passive fields
199: \cite{01FGV,01CV}, note
200: that the passive fields satisfy a linear equation of motion that
201: can be written as
202: \begin{equation}
203: \frac{\partial \phi(\B.r,t)}{\partial t}=\C.L \phi(\B.r,t)+f(\B.r,t) \ ,
204: \label{defL}
205: \end{equation}
206: with the actual form of the operator $\C.L$ determined by the problem at
207: hand.
208: In recent work \cite{01ABCPV,01CGP} it was clarified why and how passive
209: fields
210: exhibit anomalous scaling, when the velocity field is a generic
211: turbulent field. The key is to consider a
212: problem  associated with Eq. (\ref{defL}) which is the {\em decaying
213: problem} in which the forcing $f(\B.r,t)$ is put to zero.
214: The problem becomes then a linear initial value problem,
215: \begin{equation}
216: \partial \phi/\partial t =\C.L \phi\ , \label{decay}
217: \end{equation}
218: with a formal solution
219: \begin{equation}
220: \phi(\B.r,t) = \int d\B.r' \B.R(\B.r,\B.r',t) \phi(\B.r',0) \ , \label{oper}
221: \end{equation} 
222: with the operator 
223: \begin{equation}
224: \B.R\equiv T^+ \exp[{\int_0^t ds \C.L(s)}]\ , \label{defR}
225: \end{equation}
226: and $T^+$ being the time ordering operator.
227: Define next the {\em time dependent} correlation
228: functions of the decaying problem:
229: \begin{equation}
230: G^{(m)}(\B.r_1,\cdots ,\B.r_m,t)\equiv
231: \langle \phi(\B.r_1,t)\cdots
232: \phi(\B.r_m,t)\rangle \ . \label{defG}
233: \end{equation}
234: Here pointed brackets without subscript $f$ refer to the decaying
235: object in which averaging is taken with respect to realizations of
236: the velocity field only. As a result of Eq. (\ref{oper}) the decaying
237: correlation functions are developed by a propagator
238: $\C.P^{(m)}_{\underline{\B.r}|\underline{\B.\rho}}$, (with
239: $\underline {\B.r}\equiv \B.r_1,\B.r_2,\!\cdots\!,\B.r_m$)~:
240: \begin{equation}
241: G^{(m)}(\B.r_1,\!\cdots\!, \B.r_m,t)=\int\!\! d\underline{\B.\rho}
242:  \C.P^{(m)}_{\underline{\B.r}|\underline{\B.\rho}}(t)~
243: G^{(m)}(\B.\rho_1,\!\cdots\!, \B.\rho_m,0) \ . \label{defprop}
244: \end{equation}
245: 
246: In writing this equation we made explicit use of the fact that
247: the {\em initial} distribution of the passive field $\phi(\B.r,0)$
248: is statistically independent of the advecting velocity field. Thus
249: the operator $\C.P^{(m)}_{\underline{\B.r}|\underline{\B.\rho}}$
250: can be written explicitly
251: \begin{equation}
252: \C.P^{(m)}_{\underline{\B.r}|\underline{\B.\rho}}(t)\equiv
253: \langle  \B.R(\B.r_1,\B.\rho_1,t) \B.R(\B.r_2,\B.\rho_2,t)\cdots
254: \B.R(\B.r_m,\B.\rho_m,t)\rangle \ .
255: \end{equation}
256: 
257: The key finding \cite{01ABCPV,01CGP} is that the operator
258: $\C.P^{(m)}_{\underline{\B.r}|
259: \underline{\B.\rho}}$ possesses {\em left} eigenfunctions of eigenvalue
260: 1, i.~e. there exist time-independent functions
261: $Z^{(m)}(\B.r_1,\B.r_2,\cdots ,\B.r_m)$ satisfying
262: \begin{equation}
263: Z^{(m)}(\B.r_1,\cdots, \B.r_m)=\int d\underline {\B.\rho}
264: \C.P^{(m)}_{\underline{\B.\rho}|\underline{\B.r}}(t)
265: Z^{(m)}(\B.\rho_1,\cdots, \B.\rho_m) \ .\label{zeromode}
266: \end{equation}
267: The functions $Z^{(m)}$ are referred to as ``Statistically Preserved
268: Structures'', being invariant to the dynamics, even though
269: {\em the operator is strongly time dependent and decaying}. How to form,
270: from these
271: functions, infinitely many conserved variables in the decaying
272: problem was shown in \cite{01ABCPV}, and is discussed again in
273: Sect. \ref{anal}. The functions
274: $Z^{(m)}(\underline{\B.r})$ are homogeneous functions of
275: their arguments, with anomalous scaling exponents $\zeta_m$:
276: \begin{equation}
277: Z^{(m)}(\lambda\underline{ \B.r}) = \lambda^{\zeta_m}
278: Z^{(m)}(\underline{\B.r})+\dots
279: \end{equation}
280: where ``$\dots$'' stand for subleading scaling terms. Since
281: Eq.~(\ref{zeromode})
282: contains $Z^{(m)}(\underline{\B.r})$ on both sides, the scaling exponent
283: $\zeta_m$
284: cannot be determined
285: from dimensional considerations, and it can be anomalous.
286: More importantly,
287: it was shown that the correlation functions of the forced case,
288: $F^{(m)}(\underline{\B.r})$ Eq. (\ref{corr}), have exactly the same scaling
289: exponents
290: as $Z^{(m)}(\underline{\B.r})$ \cite{01CGP}. In the scaling sense
291: \begin{equation}
292: F^{(m)}(\underline{\B.r})\sim Z^{(m)}(\underline{\B.r}) \ .
293: \label{scalesame}
294: \end{equation}
295: This is how anomalous scaling in passive fields is understood.
296: Lastly, we note that for the operator governing the time derivative
297: of Eq. (\ref{defG}), $Z^{(m)}(\underline{\B.r})$ is a zero mode. We
298: will use the terms ``Statistically Preserved Structures" and ``zero modes"
299: interchangeably.
300: 
301: Of course, returning to the active fields, it makes no sense to consider the
302: decaying
303: problem; as the active field decays, the statistics of the velocity field
304: changes, and there is very little to say. On the other
305: hand, we propose that it is possible to learn a great deal from considering
306: the forced
307: solutions, comparing the forced correlation functions of the active field
308: with those of the passive field when advected
309: by the same velocity field \cite{01CMMV}. The rest of this paper is devoted
310: to making
311: this point clear and solid.
312: 
313: In Sect. \ref{convection} we discuss the active problem (\ref{Boussinesq})
314: in comparison with the passive problem (\ref{passives}). A preliminary
315: report of the correspondence between these problems was presented in
316: \cite{02CCGP}. Since we are interested in points of principle rather than
317: quantitative details, we opt to work with a shell model of the turbulent
318: convection problem. We
319: will argue (cf. Sect. \ref{summary}) that there are excellent reasons to
320: believe that the
321: results found for  the shell model translate verbatim to the partial
322: differential
323: equations. The main result of Sect. \ref{convection} is that the forced
324: $2m$th-order
325: correlation functions of the active and passive fields are both dominated by
326: the
327: Statistically Preserved Structures of the decaying passive problem,
328: i.~e. the functions  $Z^{(2m)}(\underline{\B.r})$ of Eq. (\ref{zeromode}).
329: The anomalous
330: scaling exponents are the same for the passive and active {\em forced}
331: correlation
332: functions, they are universal (independent of the forcing $f(\B.r,t)$) and
333: determined by
334: the scaling exponents of  $Z^{(2m)}(\underline{\B.r})$. We present a careful
335: discussion of the role of the statistical correlations between the forcing
336: and the velocity field that exist in the active case, but are absent
337: in the passive case. In the present problem the net result of these
338: correlations is just an amplitude factor relating the moments of the
339: two fields. In Sect. \ref{vector} we turn to a shell model of
340: magneto-hydrodynamics.
341: On the face of it, this is a counter-example to the previous case: the
342: active and
343: passive fields exhibit radically different scaling exponents. The main
344: result
345: of this section is that nevertheless the Statistically Preserved Structures
346: of the passive problem are shown to dominate the statistics of the active
347: problem, but the existence of a conservation law in the latter results
348: in an exact cancellation of the amplitude of the leading zero mode.
349: We identify analytically the leading and subleading exponents of the
350: passive problem, and then observe the cancellation of the leading
351: contribution by the dynamics. The Summary section \ref{summary} presents
352: the general lesson for the statistical physics of the (nonlinear)
353: active problem. We propose that the zero modes of the auxiliary 
354: passive fields will always have a dominant role in the
355: statistics of active fields. The active fields will thus
356: share the same scaling exponents as the passive fields unless there
357: exist additional conservation laws for the active fields. In all cases
358: the calculation of the active scaling exponents can be achieved in the
359: context of the passive problem, which boils down to finding the 
360: zero modes of a linear operator. 
361: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
362: \section{Active and passive scalars in a model of turbulent convection}
363: \label{convection}
364: \subsection{Model and numerical results}
365: In this section we 
366: examine in detail a shell model of active and passive scalars
367: for which the statistical object can be computed to high
368: accuracy. We consider a model that reproduces the conservation laws and
369: the form of coupling between the active field and the velocity field in Eqs.
370: (\ref{active}) and (\ref{Boussinesq}). Our model is a variant
371: of the shell model studied in ref. \cite{Brand}:
372: \begin{widetext}
373: \begin{eqnarray}
374: \frac{\partial u_n}{\partial t}&=&ak_n(u_{n-1}^2-\lambda u_nu_{n+1})
375: +bk_n(u_nu_{n-1}-\lambda u^2_{n+1})-\nu k_n^2 u_n+T_n\ ,\label{convu}\\
376: \frac{\partial T_n}{\partial t}&=&\tilde ak_n(u_{n-1}T_{n-1}-\lambda
377: u_nT_{n+1})
378: +\tilde b k_n(u_nT_{n-1}-\lambda u_{n+1}T_{n+1})-\kappa k_n^2
379: T_n+f_0\delta_{n,0}\
380: ,\label{convT}\\
381: \frac{\partial C_n}{\partial t}&=&\tilde ak_n(u_{n-1}C_{n-1}-\lambda
382: u_nC_{n+1})
383: +\tilde bk_n(u_n C_{n-1}-\lambda u_{n+1}C_{n+1})-\kappa k_n^2
384: C_n+f_0\delta_{n,0} \ .
385: \label{convC}
386: \end{eqnarray}
387: \end{widetext}
388: In this model all the field variables are real and $n$ stands for the index
389: of a 
390: shell of wavevector
391: $k_n=k_0\lambda^n$, with $n=0,1,\cdots, N-1$. We take $\lambda=2$, and the
392: parameters used
393: in the simulation are $a=0.01$, $\tilde a=\tilde b=b=1$,
394: $k_0=1$, $\kappa=\nu=5\times 10^{-4}$. The number of shells
395: is $N=30$, and the forcing is white noise of zero mean on the first shell.
396: 
397: Without the coupling to $T_n$, the velocity equation has
398: an inviscid unstable Kolmogorov fixed point, $u_n\sim k_n^{-1/3}$.
399: This is changed by the coupling \cite{Brand}, and the system of equations
400: for $T_n$ and $u_n$ exhibits an inviscid unstable Bolgiano
401: fixed point, $u_n\sim k_n^{-3/5}$, $T_n\sim k_n^{-1/5}$.
402: The chaotic dynamics renders the statistics of the velocity field strongly
403: non-Gaussian (cf. inset in Fig. \ref{scexp}). The exponents $\zeta_n^T$
404: for the active scalar are markedly anomalous, whereas, for the velocity,
405: they appear closer to normal (see Fig. \ref{scexp}).  The equation of motion
406: for the passive field $C$ is identical to the equation of motion of $T$,
407: but it does not affect the velocity field $u$. This equation has
408: a $C\to -C$ symmetry, whereas the coupled system of $T$ and $u$ lacks
409: this symmetry. This difference is reflected in the statistics of the two
410: fields.
411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
412: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
413: \begin{figure}
414: \centering
415: \includegraphics[width=.45\textwidth]{MHDFig1.eps}
416: \caption{The scaling exponents $\zeta_n^u$ of the velocity field (circles)
417: and $\zeta_n^T$ of the active scalar field (squares) for even $n$'s.
418: The solid lines are respectively $3n/5$ and $n/5$ for the velocity
419: and the active scalar fields. Shown in the inset is
420: the PDF of $z\equiv u_n/\langle  u_n^2 \rangle^{1/2}$
421: at shell $n=14$.}
422: \label{scexp}
423: \end{figure}
424: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
425: 
426: To demonstrate this
427: difference between the active and passive fields, we show in Fig. \ref{PDF}
428: the probability distribution functions (PDF) of $x= \phi_n/\langle
429: \phi_n^2\rangle^{1/2}$  where $ \phi_n$ is $ T_n$ or $C_n$,
430: for $n=14$. One clearly sees the symmetry of the PDF
431: of the passive scalar, in contradistinction to the asymmetry
432: of the PDF of the active scalar. This is typical to all $n$
433: in the inertial range. This is a demonstration of the
434: discussion after Eq. (\ref{phioft}). For the passive scalar the
435: odd moments vanish, whereas for the active scalar they all exist.
436: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
437: \begin{figure}
438: \centering
439: \includegraphics[width=.45\textwidth]{MHDFig2.eps}
440: \caption{The PDF's of the active (solid) and passive (dashed)
441: scalars at shell $n=14$. Note that the PDF of
442: the active scalar is asymmetric.}
443: \label{PDF}
444: \end{figure}
445: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
446: The situation is altogether different for the statistics of
447: even moments. To demonstrate the difference we plot in Fig. \ref{PDFsq}
448: the (typical) PDF of $\tilde T_n^2$ and $C_n^2$ for $n=9$ and 14,
449: where $\tilde T_n\equiv T_n-\langle T_n\rangle$. In plotting
450: we realize that the passive scalar is defined up to a constant,
451: so for the passive scalar the PDF is plotted for the rescaled
452: variable $\beta C^2_n$, where
453: \begin{equation}\beta=\langle \tilde T _n^2\rangle_f /
454: \langle C_n^2\rangle_f \approx 0.6327 \ . \label{defbeta}
455: \end{equation} 
456: Note that there is only one
457: numerical freedom $\beta$, constant for all $n$ in the inertial range.
458: An understanding of this numerical constant based on dynamical
459: considerations
460: is given in the next subsection.
461: 
462: We find very close agreement of all the PDF's in the
463: inertial range.
464: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
465: \begin{figure}
466: \centering
467: \includegraphics[width=.45\textwidth]{MHDFig3.eps}
468: \caption{
469: The PDF's of $y$ where $y=\tilde T_n^2$ (solid)
470: or $\beta C_n^2$ (dashed) at shells $n=9$ and 14.}
471: \label{PDFsq}
472: \end{figure}
473: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
474: The identity of the PDF's of $\tilde T_n^2$ and $C_n^2$ translates
475: automatically to the identity of the even-order
476: structure functions $F^{(2m)}(k_n)\equiv \langle \tilde
477: \phi_n^{2m}\rangle_f$,
478: where $\tilde \phi_n = \tilde T_n$ or $C_n$ (up to a constant $\beta^m$).
479: This is demonstrated in Fig. \ref{structure}.
480: We see that the 2nd, 4th and 6th-order
481: structure functions are barely distinguishable, with the
482: same scaling exponents in the inertial range.
483: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
484: \begin{figure}
485: \centering
486: \includegraphics[width=.45\textwidth]{MHDFig4.eps}
487: \caption{
488: The even-order structure functions $\langle \tilde T_n^{2m} \rangle_f$
489: (circles)
490: and $\langle \beta^m C_n^{2m} \rangle_f$ (squares), with
491: $m=1$, 2 and 3, from top to bottom.}
492: \label{structure}
493: \end{figure}
494: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
495: Finally, we demonstrate that the identity of the statistics
496: of the squares of the passive and
497: active scalars transcends structure functions. Consider
498: for example the multi-point correlation functions
499:  $\langle \tilde T_n^2
500: \tilde T^2_{n+5}\rangle$
501: and $\langle \tilde T^2_n \tilde T^2_{n+5}\tilde T^2_{2n}\rangle$.
502: In Fig. \ref{multi} these  correlation functions
503: are compared to their passive counterparts.
504: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
505: \begin{figure}
506: \centering
507: \includegraphics[width=.45\textwidth]{MHDFig5.eps}
508: \caption{Upper:
509: $\log_{10}\langle \tilde T_n^2\tilde T_{n+5}^2 \rangle_f$ (circles) and
510: $\log_{10}\langle \beta^2 C_n^2C_{n+5}^2 \rangle_f$ (squares). Lower:
511: $\log_{10} \langle \tilde T_n^2\tilde T_{n+5}^2\tilde T_{2n}^2 \rangle_f$
512: (circles) 
513: and $\log_{10} \langle \beta^3C_n^2C_{n+5}^2C_{2n}^2 \rangle_f$ (squares).}
514: \label{multi}
515: \end{figure}
516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
517: The conclusion is that again the multi-point correlation
518: functions are indistinguishable once the passive
519: ones are rescaled by $\beta^q$ where $q$ is the overall
520: order of the correlation function.
521: %%%%%%%%%%%%%%%%%%%%%%%%%
522: \subsection{Analysis of the results}
523: \label{anal}
524: To understand the results we start with the passive field,
525: demonstrating that its forced structure functions are
526: actually Statistically Preserved Structures. Consider then
527: the decaying passive problem, i.~e. Eq.~(\ref{convC}) without the
528: forcing term. The initial value problem for the time dependent
529: structure functions $G^{(2m)}(k_n,t)\equiv \langle C_n^{2m}(t) \rangle$
530: is the shell analog of Eq. (\ref{defprop}),
531: \begin{equation}
532: G^{(2m)}(k_n,t) =  \C.P^{(2m)}_{n,n'}(t) G^{(2m)}(k_{n'},t=0) \ ,
533: \label{defprop2}
534: \end{equation}
535: which defines the $2m$th-order
536: propagator $\C.P^{(2m)}_{n,n'}(t)$. Here and below, repeated indices are
537: being summed over. In Fig. \ref{SPSpas}
538: we show typical decay plots for the quantity $K^{(2m)}\equiv \sum_n
539: G^{(2m)}(k_n,t)$ for
540: $m=1,2,3$, starting from the initial conditions
541: $G^{(2m)}(k_n,t=0)=\delta_{n,16}$.
542: 
543: Statistically Preserved Structures in this case 
544: represent left-eigenfunctions
545: $Z^{(2m)}(k_n)$ of eigenvalue 1 satisfying
546: \begin{equation}
547: Z^{(2m)}(k_{n'}) =  Z^{(2m)}(k_n)\C.P^{(2m)}_{n,n'}(t)  \ . \label{defZ}
548: \end{equation}
549: The statement that we want to demonstrate is that the forced
550: structure functions $F_n^{(2m)}$ of the passive scalar scale like these
551: eigen-modes of the decaying problem:
552: \begin{equation}
553: F^{(2m)}(k_n) \equiv \langle C^{2m}_n\rangle_f \sim Z^{(2m)}(k_n) \ .
554: \label{true}
555: \end{equation}
556: To demonstrate this we use the method of \cite{01ABCPV} and
557: define the quantities $I^{(2m)}$,
558: \begin{equation}
559: I^{(2m)} = \sum_n G^{(2m)}(k_n,t) F^{(2m)}(k_n) \ .
560: \end{equation}
561: Using Eqs. (\ref{defprop2}) and (\ref{defZ}) we see that if
562: Eq.~(\ref{true}) is obeyed, than the quantities $I^{(2m)}$ are time
563: independent. Indeed, in Fig. \ref{SPSpas} we demonstrate the stationarity
564: of these objects, thus supporting Eq. (\ref{true}). The analytic
565: explanation as to why the forced solutions agree with the Statistically
566: Preserved Structures of the decaying problem was provided in \cite{01CGP}.
567: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
568: \begin{figure}
569: \centering
570: \includegraphics[width=.45\textwidth]{MHDFig6a.eps}
571: \centering
572: \includegraphics[width=.45\textwidth]{MHDFig6b.eps}
573: \centering
574: \includegraphics[width=.45\textwidth]{MHDFig6c.eps}
575: \caption{The decaying objects $K^{(2m)}$ (solid lines) and the conserved
576: objects
577: $I^{(2m)}$ (dashed lines) as a function of time, for $m=1,2$ and 3. Time is
578: measured
579: here in units of the largest scale eddy turn over time $\tau_0\equiv
580: [k_0 \sqrt{S_2(k_0)}]^{-1}\approx 22$. In
581: panels b and c we include in dotted lines the quantity $I^{(2m)}$ in which
582: we replaced
583: $F^{(2m)}$ by its dimensional prediction $[F^{(2)}]^m$. We see that using
584: the
585: dimensional exponent does not make $I^{(2m)}$ time invariant.}
586: \label{SPSpas}
587: \end{figure}
588: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
589: Before turning to the active field, it is worthwhile to observe
590: how any initial condition of the decaying passive field lands
591: on the scaling solution that is represented by the Statistically
592: Preserved Structure. Consider the initial value experiment that is
593: reported in Fig. \ref{initsc}.  Here we start,
594: as an example, from the initial value
595: $C_n(t=0)\propto k_n^{2/3}$. In this initial condition the order
596: of the amplitudes is inverted with respect to the spectrum of
597: the passive scalar.  We plot, as a function of time,
598: the trajectories of
599: $C_n(t)$ as computed just from this initial condition, averaged over
600: 650 realizations. We see that the trajectories
601: land on a decaying scaling solution in which the order of the amplitudes and
602: the ratios between them are identical to the spectrum of the zero mode of
603: the 
604: passive field; the decay that we see, at a rate proportional to $1/t^2$,
605: is entirely due to dissipative effects, as explained in some detail
606: in \cite{01CGP}.
607: 
608: Finally, we need to understand how the forced active scalar $T$ falls on
609: the Statistically Preserved Structure of the decaying passive
610: problem, and what is the origin of the factor $\beta$ in (\ref{defbeta}).
611: To this aim we note that both equations for passive and active fields can
612: be written as
613: \begin{equation}
614: \frac{\partial \phi_n}{\partial t}=\C. L_{n,n'}\phi_{n'}
615: +f_0\delta_{n,0}\ . \label{convphi}
616: \end{equation}
617: This equation has a formal solution in the form
618: \begin{equation}
619: \phi_n(t) = R_{n,n'}(t|0)\phi_{n'}(t=0)
620: +\int_0^t d\tau R_{n,0}(t|\tau)f_0(\tau)\ , 
621: \end{equation}
622: where $R_{n,n'}(t|\tau)$ is the shell analog of the operator (\ref{defR}),
623: \begin{equation}
624: R_{n,n'}(t|\tau) \equiv \left\{T^+ \int_\tau^t \exp
625: [\C.L(s)]ds\right\}_{n,n'} \ .
626: \label{phioft}
627: \end{equation}
628: The first difference between the
629: active and passive fields is encountered when we take the
630: average of this equation. For the passive case, the average
631: can be taken by decorrelating $f_0$ and $R_{n,n'}$. 
632: Since the mean of the force $f_0$ vanishes, we get
633: \begin{equation}
634: \langle C_n \rangle_f=0  
635: \label{meanC}
636: \end{equation}
637: Such a decorrelation is, however, not 
638: allowed in the active case since the forcing $f_0$
639: is correlated with $T$, which is itself correlated 
640: with $\B.u$ and thus with $R_{n,n'}$. Hence 
641: \begin{equation}
642: \langle T_n \rangle_f=\langle R_{n,n'}(t|0)T_{n'}(t=0)\rangle_f
643: \nonumber\\+\int_0^t d\tau \langle R_{n,0}(t|\tau)f_0(\tau)\rangle_f 
644: \label{meanT}
645: \end{equation}
646: and $T$ has a non-zero mean.
647: Similarly, the passive scalar has zero odd moments and its PDF is
648: symmetric. On the other hand, the active scalar has nonvanishing odd
649: moments and its PDF is asymmetric.
650: 
651: In spite of this great difference between the active and passive scalars,
652: there is a close affinity between the active field and
653: the Statistically Preserved Structures of the passive field. 
654: To see this, we note that
655: the first term on the RHS of Eq. (\ref{meanT}) represents a decaying field.
656: We expect that if the initial condition $T_n(t=0)$ has any component on
657: the Statistically Preserved Structure of the passive field, it will quickly
658: relax everything else and will land exactly on that solution. In this
659: respect it is just the same as the initial value experiment reported
660: in Fig. \ref{initsc}. In terms of the relative amplitudes of
661: the different $n$ shells
662: there is nothing in the fate of the initial value term to distinguish the
663: active and the
664: passive fields.
665: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
666: \begin{figure}
667: \centering
668: \includegraphics[width=.45\textwidth]{MHDFig7.eps}
669: \caption{An example of the fate of an initial value term as a function of
670: time,
671: in units of $\tau_0$. The initial amplitudes are inverted in order compared
672: to the 2nd order zero modes of the passive field. Shown are shells
673: $n=5,\:7,\dots,\:19$. }
674: \label{initsc}
675: \end{figure}
676: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
677: The second term on the RHS of Eq.~(\ref{meanT}) is more subtle. First, we
678: note
679: that for every value of $\tau$ we again face a decaying experiment that
680: takes
681: place between the times $\tau$ and $t$. In the language of the passive
682: field, the integrand can be read from a decaying field with initial
683: condition $C_n(t=\tau)=f_0(\tau)\delta_{n,0}$. Indeed, in our simulations
684: below
685: this is precisely how we evaluate integrals of this type. We break the
686: interval $[0,t]$ into $N$ sub-intervals $\{\tau_i\}_{i=1}^N$, $\tau_i =
687: (i/N)t$,
688: and start a decaying experiment with
689: initial conditions $f(\tau_i)\delta_{n,0}$. Measuring $C_n(t)$ and summing
690: up
691: all the contributions yields an approximation to the integral. Every term in
692: the integrand is expected to land, for most of the time $t-\tau$, on the
693: scaling solution of the passive field, in much of the same way that the
694: initial value term does.
695: Thus both terms in Eq. (\ref{meanT}) are expected to scale like the passive
696: field, which nevertheless itself has zero amplitude due to the symmetry.
697: 
698: The correlation effects that play a role for the active scalar will be
699: responsible for the factor $\beta$ that we discovered in the numerics.
700: To see this we need to consider the 2nd order structure functions:
701: \begin{eqnarray}
702: &&F^{(2)}(k_n)= \langle [R_{n,n'}(t|0)\phi_{n'}(t=0)]^2\rangle_f\label{F2}\\
703: &&+2\langle R_{n,n'}(t|0)\phi_{n'}(t=0)\int_0^t d\tau
704: R_{n,0}(t|\tau)f_0(\tau)\rangle_f
705: \nonumber\\
706: &&+\int_0^t d\tau' d\tau''\label{second}\langle
707: R_{n,0}(t|\tau')R_{n,0}(t|\tau'')f_0(\tau')f_0(\tau'')\rangle_f \ .
708: \nonumber
709: \end{eqnarray}
710: For sufficiently long time the first two terms, denoted below as
711: $\C.I_1$ and $\C.I_2$ respectively, do not contribute to the
712: structure functions, and any difference between the active and passive
713: fields must be ascribed to the last term. The last term, denoted here
714: as $\C.I_3$, has a ``diagonal"
715: contribution, which is obtained for $\tau'=\tau''$ and an ``off-diagonal"
716: contribution, which is the rest of the integral for which  $\tau'\ne
717: \tau''$.
718: For the passive field, $R_{n,n'}$ and $f_0$ decouple and only the 
719: diagonal part exists. For the active field there is no decoupling. 
720: Denoting this term $\C.I_{3,d}$, it
721: reads respectively
722: \begin{eqnarray}
723: \!\!\!\!\C.I_{3,d}\!\!&=&\!\!\int ds\langle
724: R_{n,0}(t|s)R_{n,0}(t|s)\rangle_f f_0^2 \quad \text{(passive)}\ ,
725: \label{one}\\
726: \!\!\!\!\C.I_{3,d}\!\!&=&\!\!\!\int_0^t \!\!ds\langle
727: R_{n,0}(t|s)R_{n,0}(t|s)f_0(s)f_0(s)\rangle_f \ \text{(active)}. \label{two}
728: \end{eqnarray}
729: In Fig. \ref{diagonal} we compare the integrands of these two expressions,
730: measured directly in our simulation as explained above.
731: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
732: \begin{figure}
733: \centering
734: \includegraphics[width=.45\textwidth]{MHDFig8.eps}
735: \caption{Comparison of the integrands in Eqs (\ref{one}) (the passive case)
736: and (\ref{two}) (the active case) for $n=10$. The plots are indistinguishable.}
737: \label{diagonal}
738: \end{figure}
739: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
740: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
741: \begin{figure}
742: \centering
743: \includegraphics[width=.45\textwidth]{MHDFig9.eps}
744: \caption{The integral $\C.I_3$ in comparison to the diagonal
745: part $\C.I_{3,d}$ for $n=10$. Upper panel: the passive field. The integral agrees
746: with the diagonal part at all times. Lower panel: the active field. The
747: deviations are due to the non-vanishing contribution of the off-diagonal
748: integral, which is also displayed in the next figure. The dashed
749: line in both panels represents the stationary value of the
750: corresponding second order structure function.}
751: \label{intmindiag}
752: \end{figure}
753: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
754: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
755: \begin{figure}
756: \centering
757: \includegraphics[width=.45\textwidth]{MHDFig10.eps}
758: \caption{The off-diagonal integral for $n=10$, computed as
759: $\langle [\phi_n(t)-R_{n,n'}(t|0)\phi_{n'}(t=0)]^2\rangle -\C.I_{3,d}$  for
760: the passive and active fields
761: respectively. For the passive field (dotted line) it fluctuates around zero,
762: while for
763: the active field it begins positive, and then turns negative. For longer
764: times it saturates at a constant negative value, giving rise to the
765: factor $\beta$.}
766: \label{beta}
767: \end{figure}
768: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
769: We can see that there is not much difference
770: between them; the diagonal term cannot be blamed for the factor $\beta$.
771: On the other hand, in Fig. \ref{intmindiag} we show the full integral
772: $\C.I_3$ and compare it with its diagonal term. We see that in the passive
773: case the diagonal part is everything, whereas in the active case there
774: is a difference. Lastly, we show that this difference is precisely
775: the source of the factor $\beta$. In Fig. \ref{beta} we show
776: $\langle [\phi_n(t)-R_{n,n'}(t|0)\phi_{n'}(t=0)]^2\rangle
777: -\C.I_{3,d}$ for the passive and the
778: active fields. The former fluctuates all the time around zero. The latter
779: is positive initially, and then becomes negative. For later times it
780: saturates
781: at a negative value that is precisely responsible for the factor $\beta$.
782: 
783: In summary, we find that the even correlation functions of the active
784: and passive scalars share the same scaling exponents simply because
785: the zero modes of the decaying passive problem dominate the statistics
786: of both fields. It is thus possible to understand the anomalous
787: statistics of the active field in the same way as that of the passive
788: field. We believe that this is a significant result that should be
789: put to further experimental and numerical tests in the PDE version
790: of the problem. We will return to this point in the discussion.
791: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
792: \section{Active and passive fields in a model of magnetohydrodynamics}
793: \label{vector}
794: \subsection{Model and Numerical Results}
795: 
796: In this section we examine a shell model that reproduces the type
797: of coupling and the conservation laws of Eqs.~(\ref{MHD}).  We need to be
798: careful about the dynamo effect which we want to avoid in order to
799: have stationary statistics. We thus construct the model to mimic
800: 2-dimensional
801: MHD, in which there is an inverse cascade of energy. Accordingly we need to
802: have large scale damping terms in the velocity equation, and
803: a force at intermediate scales. In all our simulations below we force both
804: the velocity and the active
805: fields on shells 10 and 11 (denoted by $n_f$), using white noise of zero
806: mean. We run the
807: model with 35 shells. The equations are an adaptation of the MHD
808: shell model of \cite{98GC} to the Sabra shell model \cite{98LPPPV}. All
809: field variables are complex numbers:
810: \begin{widetext}
811: \begin{eqnarray}
812: \label{sabrau}
813: \frac{d u_n}{dt}&=& i k_n [a \lambda (u_{n+1}^*u_{n+2}-b_{n+1}^*b_{n+2})
814: + b (u_{n-1}^*u_{n+1} - b_{n-1}^*b_{n+1})-c \lambda^{-1}(u_{n-2}u_{n-1} -
815: b_{n-2}b_{n-1})]\nonumber\\
816: &&+ f'_n \delta_{n,n_f} + \nu k_n^2 u_n+\tilde \nu k_n^{-4} u_n\ , \\
817: \frac{d b_n}{dt}&=& i k_n [\tilde{a} \lambda
818: (u_{n+1}^*b_{n+2}-b_{n+1}^*u_{n+2})
819: + \tilde{b} (u_{n-1}^*b_{n+1} - b_{n-1}^*u_{n+1})
820: -\tilde{c} \lambda^{-1}(u_{n-2}b_{n-1} - b_{n-2}u_{n-1})]\nonumber\\
821: &&+ f_n \delta_{n,n_f} +\kappa k_n^2 b_n +\tilde \kappa k_n^{-4} b_n\ .
822: \label{sabrab}
823: \end{eqnarray}
824: \end{widetext}
825: The coefficients $a$, $b$, $c$, $\tilde{a}$, $\tilde{b}$ and $\tilde{c}$
826: can be parametrized as follows~:
827: \begin{equation}
828: \begin{array}{l@{\quad}l@{\quad}l}
829: a = 1\ ,&b = -\delta\ ,&c = -(1-\delta)\ ,\\
830: \tilde{a}=1-\delta-\delta_m\ ,&\tilde{b}=\delta_m\ ,& \tilde{c}=1-\delta_m\
831: .
832: \end{array}
833: \label{sabraparam}
834: \end{equation}
835: \label{sixparam}
836: This choice ensures the conservation of the total energy and ``cross
837: helicity"
838: in the inviscid limit $\nu=\tilde \nu=\kappa=\tilde \kappa = f = f' =0$,
839: \begin{eqnarray}
840: E&=&\frac{1}{2}\sum_n (|u_n|^2 + |b_n|^2)\ , \label{toten}\\
841: K&=&\sum_n \Re(u_n^* b_n)\ .\label{crosshe}
842: \end{eqnarray}
843: To mimic the magnetic helicity, we can write down a generalized quantity
844: \begin{equation}
845: H=\frac{1}{2}\sum_n \mathrm{sign}(\delta-1)^n\frac{|b_n|^2}{k_n^\alpha},
846: \label{magnhe}
847: \end{equation}
848: with $\alpha>0$ a fixed parameter. We demand conservation of this
849: generalized
850: ``magnetic helicity", together with absence of dynamo effect. This implies
851: \begin{eqnarray}
852: \delta>1&\rightarrow&
853: \left\{\begin{array}{l}
854: \delta = 1 + \lambda^{-\alpha}\ ,\\
855: \delta_m=- 1/(\lambda^\alpha - 1)\ ,
856: \end{array}\right.
857: \label{deltagt1}
858: \end{eqnarray}
859: On the other hand, when $\delta<1$ one can have dynamo, and therefore
860: no stationary statistics.
861: 
862: In addition to the conservation laws the equations
863: of motion remain invariant to the phase transformations
864: $u_n\to u_n\exp(i\phi_n)$ and $b_n\to b_n\exp(i\psi_n)$. The conditions are
865: \begin{eqnarray}
866: \phi_n + \phi_{n+1} - \phi_{n+2} &=& 0\ ,\label{phases1}\\
867: \phi_{n} + \psi_{n+1} - \psi_{n+2} &=& 0\ ,\label{phases2}\\
868: \psi_n + \phi_{n+1}-\psi_{n+2} &=&0\ ,\label{phases3}\\
869: \psi_{n} + \psi_{n+1} - \phi_{n+2} &=& 0\ .\label{phases4}
870: \end{eqnarray}
871: This implies $\psi_n = \phi_n$ $\forall n$.
872: 
873: The passive field is denoted by $q_n$, whose evolution is given by an
874: equation similar to Eq. (\ref{sabrab}), i.~e.
875: \begin{widetext}
876: \begin{eqnarray}
877: &&\frac{d q_n}{dt}= i k_n [\tilde{a} \lambda
878: (u_{n+1}^*q_{n+2}-q_{n+1}^*u_{n+2})
879: + \tilde{b} (u_{n-1}^*q_{n+1} - q_{n-1}^*u_{n+1})- \tilde{c}
880: \lambda^{-1}(u_{n-2}q_{n-1} - q_{n-2}u_{n-1})]\nonumber\\
881: &&+\kappa k_n^2 q_n +f_n\delta_{n,n_f}\ . \label{sabraq}
882: \end{eqnarray}
883: \end{widetext}
884: The fields $q_n$ and $b_n$ are advected by the same velocity field, however
885: $b_n$ is
886: active, while $q_n$ is passive. The inviscid passive equation has only
887: one conserved variable, i.~e. Eq. (\ref{magnhe}) with $q_n$ replacing
888: $b_n$. It also satisfies the same phase relations as the active
889: field. We want to know whether the
890: scaling properties of $b_n$ are determined once again by the Statistically
891: Preserved
892: Structures of the decaying problem of the passive field $q_n$.
893: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
894: \begin{figure}
895: \centering
896: \includegraphics[width=.45\textwidth]{MHDFig11.eps}
897: \caption{Structure functions of order 2 ($\times$), 3 (+)
898: and 4 ($\diamond$) for the passive field (dotted line), active
899: field (dashed line) and velocity field (solid line). The two vertical
900: lines denote the forcing shells. Note
901: that the scaling exponents of the active field and the velocity field
902: coincide.
903: The parameters are $N=35$, $\alpha=2$, $k_0=0.0625$, $\nu=10^{-12}$,
904: $\tilde\nu
905: =10^{-3}$. The forcing is white noise on shells 10,11.}
906: \label{spectra}
907: \end{figure}
908: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
909: 
910: In Fig. \ref{spectra} we show the spectra of the passive and active fields
911: respectively, obtained from a direct numerical simulation with the
912: parameters as detailed in the figure legend. This appears like a
913: striking counter-example to the results of the previous section: the
914: two fields have totally different scaling behaviors. The active field
915: has ``standard" scaling exponents $\eta_p$, 
916: defined by $\langle |b_n|^p\rangle \sim k_n^{-\eta_p}$, that coincide
917: with those of the velocity field, defined by
918: $\langle |u_n|^p\rangle \sim k_n^{-\zeta_p}$,
919: and the spectrum decays like a power
920: law in the ``inertial range" which is between the forcing and the
921: dissipative scales. We estimate from the numerics $\eta_2=\zeta_2\approx .67$,
922: $\eta_3=\zeta_3\approx 1.0$ and $\eta_4=\zeta_4\approx 1.33$, 
923: in close correspondence
924: with the Kolmogorov dimensional predictions.  The passive
925: field has exponents, defined similarly by 
926: $\langle |q_n|^p\rangle \sim k_n^{-\beta_p}$,
927: that are with a different sign! Its spectrum is an increasing
928: function of
929: $k_n$ in the inertial range. We measure $\beta_2\approx -1.33$,
930: $\beta_3\approx -2$ and $\beta_4\approx -2.67$. If we assume that the
931: passive field
932: lands on the Statistically Preserved Structures of the passive decaying problem,
933: then it
934: appears that the active field does not do so.
935: 
936: In the rest of this section we will show that this is actually not a
937: counterexample
938: to the proposition that the active field lands on Statistically Preserved
939: Structures of the decaying passive field. It does. What happens here is
940: that,
941: due to the conservation law Eq. (\ref{crosshe}), the amplitude of the
942: leading
943: Statistically Preserved Structure with the negative scaling exponent
944: is exactly zero. The active
945: field then lands on a sub-leading zero mode, which has standard, positive
946: scaling exponent (the positive sign refers to $\B.r$-space representation,
947: as in Eq. (\ref{zeta})). 
948: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
949: \subsection{Analysis of the results}
950: 
951: To gain insight into this interesting situation we note that the
952: analog of Eq.~(\ref{phioft}) describes the dynamics of our active
953: field $b_n$:
954: \begin{equation}
955: b_n(t) = R_{n,n'}(t|0)b_{n'}(t=0)
956: +\int_0^t d\tau R_{n,n_f}(t|\tau)f_{n_f}(\tau)\ , \label{boft}
957: \end{equation}
958: with an obvious re-definition of the present operator $R_{n,n'}$.
959: It is very revealing to examine the time dependence of the two
960: terms on the RHS of this equation. We measure time in units
961: of the eddy turn over time of the forcing shell 10. This is
962: defined as $\tau_{10}\equiv 
963: [k_{10} \sqrt {\langle |u_{10}|^2 \rangle}]^{-1}\approx 3.35 $.
964: We will examine a forced system
965: which began running at $t=-\infty$,  denoting a generic
966: time as $t=0$. In Fig. \ref{twoterms} panel a
967: we show the time-dependence of the first term for 6 values of $n$
968: in the inertial interval. We see that the initial conditions
969: represent, as expected, a 'standard' spectrum in which the
970: amplitude $b_n$ decreases as a function of $n$. As time proceeds,
971: the decaying term cannot recognize its being `active' from being
972: `passive', and it switches rapidly to the Statistically Preserved
973: Structure of the decaying passive field, characterized by a negative
974: exponent. If it were not for the second term on the RHS of
975: Eq. (\ref{boft}), then $b_n$ would have landed on the same
976: solution as $q_n$. What about the second term? In panel b of
977: Fig. \ref{twoterms} we show the $n$ dependence of the
978: term at time $t=3\times 10^{-4}$. We see that also this term agrees,
979: in its $n$ dependence, with the negative exponent of the
980: passive field. Yet, the LHS $b_n(t)$ fluctuates around
981: {\em decreasing} amplitudes as $n$ increases, meaning that
982: the leading (negative) exponent exactly cancels between
983: the two terms on the RHS of Eq. (\ref{boft}). We demonstrate this
984: cancellation in Fig. \ref{cancel}. There we plot the real
985: parts of the initial value term and the integral term
986: at time $t= 0.3$. We see that the two terms cancel each other.
987: The imaginary parts exhibit the same behavior.
988: 
989: Next we need
990: to understand this cancellation from the analysis of the
991: equations of motion. With this analysis we will also show
992: that the solution on which $b_n(t)$ is landing is also
993: a Statistically Preserved Structure of the decaying passive
994: field, albeit with a sub-leading scaling exponent.
995: \subsection{Statistically Preserved Structures of the passive
996: field}
997: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
998: \begin{figure}
999: \centering
1000: \includegraphics[width=.45\textwidth]{MHDFig12a.eps}
1001: \includegraphics[width=.45\textwidth]{MHDFig12b.eps}
1002: \caption{Panel a: The fate of the modulus of the initial value term, averaged over
1003: $2,000$  realizations. Shown
1004: are shells 15, 17, 19, 21, 23 and 25 (top to bottom from the left-most
1005: side). 
1006: At time $t=0$ their
1007: relative amplitudes agree with the scaling exponent of the
1008: {\em active} field. As time progresses the decaying field switches
1009: to the relative amplitudes which agree with the scaling exponent
1010: of the {\em passive} field. Panel b: The modulus of four realizations of the integral
1011: as a function of $n$, for time $t=3\times 10^{-4}$ (in unit of
1012: $\tau_{10}$). Note that both terms of the
1013: RHS of Eq. (\ref{boft}) exhibit the same {\em leading} scaling
1014: behavior. This is canceled exactly as is demonstrated in Fig. \ref{cancel}}
1015: \label{twoterms}
1016: \end{figure}
1017: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1018: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1019: \begin{figure}
1020: \centering
1021: \includegraphics[width=.45\textwidth]{MHDFig13.eps}
1022: \caption{the real part of the initial value term (dashed line) and
1023: integral term (solid line) as a function of $k_n$. This is a demonstration
1024: of the cancellation of the leading order term in favor of the subleading
1025: one. The imaginary parts behave in the same way.}
1026: \label{cancel}
1027: \end{figure}
1028: In the next subsection we will show that the velocity
1029: field attains a scaling solution with $\zeta_3=1$:
1030: \begin{equation}
1031: S_3(k_n)\equiv 
1032: \Im<u_{n-1}u_{n}u_{n+1}^*>_f \sim k_n^{-1} \ .
1033: \label{S3ofk}
1034: \end{equation}
1035: where $\Im$ denotes the imaginary
1036: part. In this section we will assume this, and examine what are the
1037: scaling solutions that agree with the existence of a 2nd order
1038: Statistically Preserved Structure for the passive field.
1039: We are not going to compute the anomalous scaling exponent exactly,
1040: but rather obtain their dimensional estimates. Since we are after a
1041: radically different apparent behavior, small numerical corrections
1042: are not our main concern. With this caveat in mind,
1043: we can calculate the exponent $\beta_3$ characterizing the third order
1044: structure function 
1045: \begin{equation}
1046: P_3(k_n)\equiv \Im \langle q_{n-1}q_n q^*_{n+1}\rangle \sim k_n^{-\beta_3} \ .
1047: \label{P3ofk}
1048: \end{equation}
1049: The condition for the existence of the 2nd order Statistically Preserved
1050: Structure is
1051: \begin{equation}
1052: \frac{d}{dt}\langle |q_n|^2\rangle =0\ , \quad \forall n \ ,
1053: \label{condzero}
1054: \end{equation}
1055: in the inviscid limit. Using Eq.~(\ref{sabraq})
1056: this condition generates a number of third order quantities that need
1057: to be analyzed first. Denote therefore
1058: \begin{eqnarray}
1059: Q_{3,1}(k_n) &\equiv& \Im<u_{n-1}q_{n}q_{n+1}^*>\ , \nonumber\\
1060: Q_{3,2}(k_n) &\equiv& \Im<q_{n-1}u_{n}q_{n+1}^*>\ , \nonumber\\
1061: Q_{3,3}(k_n) &\equiv& \Im<q_{n-1}q_{n}u_{n+1}^*>\  . \label{Qdef}
1062: \end{eqnarray}
1063: In order to construct scaling solutions for these objects, Dimensional
1064: consideration imply that the
1065: fields involved in the averages above have scalings $u_n\propto
1066: \lambda^{- n/3}$ and $q_n \propto \lambda^{-\beta_3 n/3}$. We infer
1067: the expressions
1068: \begin{eqnarray}
1069: Q_{3,1}(k_n)&=&|q_0|^2|u_0|k_n^{-(2\beta_3+1)/3}
1070: \lambda^{-(\beta_3-1)/3} \ ,\nonumber\\
1071: Q_{3,2}(k_n)&=&|q_0|^2|u_0|k_n^{-(2\beta_3+1)/3}\ , \nonumber\\
1072: Q_{3,3}(k_n)&=&|q_0|^2|u_0|k_n^{-(2\beta_3+1)n/3}
1073: \lambda^{(\beta_3-1)/3} \ .
1074: \end{eqnarray}
1075: We can thus rewrite
1076: \begin{eqnarray}
1077: Q_{3,1}(k_n) &\equiv& \lambda^{-(\beta_3-1)/3} \tilde Q_3(k_n)\
1078: ,\label{redefQ31}\\
1079: Q_{3,2}(k_n) &\equiv& \tilde Q_3(k_n)\ ,\label{redefQ32}\\
1080: Q_{3,3}(k_n) &\equiv& \lambda^{(\beta_3-1)/3} \tilde Q_3(k_n)\
1081: ,\label{redefQ33}
1082: \end{eqnarray}
1083: where $\tilde Q_3(k_n)$ scales like
1084: \begin{equation}
1085: \tilde Q_3(k_n)\equiv \lambda^{-(2\beta_3+1)n/3}\ .
1086: \label{scalingQ3}
1087: \end{equation}
1088: having these definitions in mind we derive, by demanding
1089: Eq.~(\ref{condzero})
1090: and substituting the scaling form of $\tilde Q_3$ Eq. (\ref{scalingQ3}),
1091: the equation
1092: \begin{widetext}
1093: \begin{equation}
1094: \tilde{a}\lambda^{-(2\beta_3-2)/3}
1095: (1 - \lambda^{(\beta_3-1)/3})+
1096: \tilde{b}(\lambda^{-(\beta_3-1)/3}-\lambda^{(\beta_3-1)/3})
1097: +\tilde{c}\lambda^{(2\beta_3-2)/3}
1098: (\lambda^{-(\beta_3-1)/3}-1) = 0\ .
1099: \label{eqeta2}
1100: \end{equation}
1101: \end{widetext}
1102: This is  a fourth order polynomial in $\lambda^{(\beta_3-1)/3}$. The
1103: four roots are
1104: \begin{equation}
1105: \lambda^{(\beta_3-1)/3} = \left\{
1106: \begin{array}{l}
1107: 1\ ,\\
1108: \lambda^{-\alpha}\ ,\\
1109: \pm \lambda^{-\alpha/2}\ .
1110: \end{array}\right.
1111: \label{passscal2d}
1112: \end{equation}
1113: Here three of the roots correspond to a priori physical solutions:
1114: \begin{equation}
1115: \beta_3 = \left\{
1116: \begin{array}{l}
1117: 1\ ,\\
1118: 1 - 3\alpha/2\ ,\\
1119: 1 - 3\alpha\ .
1120: \end{array}\right.
1121: \label{solbeta}
1122: \end{equation}
1123: In our simulations with $\alpha=2$ these results are $\beta_3=1$, -2 and -5
1124: respectively. This is in agreement with spectral exponents $\beta_2$
1125: of the order of (neglecting anomalies) $\beta_2=2/3, -4/3, -10/3$.
1126: To know which of these is physical, we need to consider the fluxes
1127: supported by these solutions.
1128: 
1129: The only flux that is relevant for the passive field is the magnetic
1130: helicity. For the case considered here with $\delta>1$ 
1131: it can be conveniently computed at the shell
1132: $M$ by evaluating
1133: \begin{equation}
1134: \Phi^H_M \equiv - \frac{1}{2}\frac{d}{dt}\sum_{n=0}^M
1135: \left<\frac{|q_n|^2}{k_n^\alpha}\right>
1136: \ .\label{hflux}
1137: \end{equation}
1138: Using the equations of motion to evaluate this object we find
1139: \begin{widetext}
1140: \begin{eqnarray}
1141: \Phi^H_M = &-& \delta_m
1142: \Big[ k_{M+1}^{1-\alpha}
1143: \big(\Im<q_{M}u_{M+1}q_{M+2}^*> - \Im<q_{M}q_{M+1}u_{M+2}^*>\big)\nonumber\\
1144: &+& k_M^{1-\alpha}\big(\Im<q_{M-1}u_{M}q_{M+1}^*>
1145: + \Im<u_{M-1}q_{M}q_{M+1}^*>\big)\Big]\ . \label{exphflux}
1146: \end{eqnarray}
1147: \end{widetext}
1148: We can evaluate now the magnetic helicity flux for the three
1149: scaling solutions (\ref{solbeta}). We find
1150: \begin{equation}
1151: \Phi^H_M \propto \left\{
1152: \begin{array}{l}
1153: \lambda^{-\alpha M}\ ,\\
1154: 1\ ,\\
1155: \lambda^{\alpha M} \ .
1156: \end{array}\right.
1157: \label{solpassiveflux}
1158: \end{equation}
1159: We conclude that the third solution is unphysical, since it supports
1160: a flux that diverges with $M$. The first two solutions are allowed.
1161: With $\beta_3=-2$ we get a constant flux; this is the leading
1162: scaling solution, and is indeed realized in the simulations. The
1163: solution $\beta_3=1$ is subleading, it is associated with a decaying
1164: flux, and is asymptotically allowed. It is not observed in the passive
1165: field simulations simply because it is subleading.
1166: 
1167: Our main point will be that the {\em active} field will in its turn
1168: land on the subleading Statistically Preserved Structure
1169: because the additional conservation laws exclude the leading
1170: one. We demonstrate this phenomenon in the next subsection.
1171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1172: \subsection{Why does the active field fall on a subleading zero mode?}
1173: 
1174: If we accept the general philosophy that active fields exhibit
1175: scaling behaviors that are determined by the zero modes of the
1176: auxiliary passive fields, then we should explain here why
1177: in the present case the active field avoids the {\em leading}
1178: zero mode, and appears to land on the sub-leading one. The answer
1179: is hidden of course in the conservation laws, as we expose now.
1180: 
1181: We first repeat the analysis performed in the previous
1182: section to find the consequence of the equation
1183: \begin{equation}
1184: \frac{d}{dt} <|b_n|^2> = 0\ ,
1185: \label{ddtb2}
1186: \end{equation}
1187: or, equivalently,
1188: \begin{equation}
1189: \frac{d}{dt} <|u_n|^2> = 0\ .
1190: \label{ddtu2}
1191: \end{equation}
1192: Using now the definitions
1193: \begin{eqnarray}
1194: B_3(k_n) &=& \Im \langle b_{n-1} b_n b^*_{n+1}\rangle \sim
1195: k_n^{-\eta_3}\nonumber\\
1196: Q_3(k_n)&\equiv&  k_n^{-(2\eta_3+\zeta_3)/3} \ , \
1197: \end{eqnarray}
1198: we obtain an equation that is analogous to Eq.~(\ref{eqeta2}),
1199: \begin{widetext}
1200: \begin{equation}
1201: \tilde{a}\lambda^{1-(2\eta_3+\zeta_3)/3}
1202: (1 - \lambda^{(\eta_3-\zeta_3)/3})+
1203: \tilde{b}(\lambda^{-(\eta_3-\zeta_3)/3}-\lambda^{(\eta_3-\zeta_3)/3})
1204: +\tilde{c}\lambda^{-1+(2\eta_3+\zeta_3)/3}
1205: (\lambda^{-(\eta_3-\zeta_3)/3}-1) = 0\ .
1206: \label{eqeta}
1207: \end{equation}
1208: \end{widetext}
1209: This is a fourth degree polynomial for $\lambda^{\eta_3/3}$ if $\zeta_3$
1210: is known. Obviously,
1211: if we simply substituted here $\zeta_3=1$ we would get the same predictions
1212: for $\eta_3$ as obtained for $\beta_3$ in Eq. (\ref{solbeta}). However, we
1213: have
1214: in this case an important additional constraint that is absent in the case
1215: of the passive field, which can be inferred from the additional
1216: conservation equation
1217: \begin{equation}
1218: \frac{d}{dt} \Re<u_n^*b_n> = 0
1219: \label{ddtub}
1220: \end{equation}
1221: Repeating the analysis as above, and introducing a  new object
1222: $A_3(k_n)\equiv k_n^{-(\eta_3+2\zeta_3)}$ yields the two equations
1223: \begin{eqnarray}
1224: &&a\lambda B_3(k_{n+1}) + b B_3(k_n) + c\lambda^{-1}B_3(k_{n-1}) =0\ ,
1225: \label{eqP3}\\
1226: &&a\lambda^{1-(\eta_3-\zeta_3)/3} A_3(k_{n+1})+ b A_3(k_n)\nonumber\\
1227: &&\quad + c\lambda^{-1+(\eta_3-\zeta_3)/3}A_3(k_{n-1})=0\ ,
1228: \label{eqR3}
1229: \end{eqnarray}
1230: Solving this system together with Eq. (\ref{eqeta}) yields the scaling
1231: exponents
1232: \begin{equation}
1233: \zeta_3,\:\eta_3 = \left\{\begin{array}{l@{\quad}l}1&\\
1234: 1+ \log_{\lambda}(a/c)& \ .
1235: \end{array}
1236: \right.
1237: \label{scalingsol}
1238: \end{equation}
1239: %With our parameters $\log_{\lambda}(a/c)=2$ so the second solution is
1240: %strongly subleading.
1241: It is easy to check that, among the four possible combinations of
1242: this equation, the only solution allowed by Eq. (\ref{eqeta}) is
1243: \begin{equation}
1244: \zeta_3=\eta_3=1\ .
1245: \label{zetaeta}
1246: \end{equation}
1247: We thus conclude that as far as the active field
1248: is considered, the additional conservation law rules out the leading
1249: zero mode of the passive problem, leaving us only with the subleading mode
1250: which is observed in the simulations.
1251: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1252: \section{Summary and Conclusions}
1253: \label{summary}
1254: In this paper we considered the correspondence between the statistics
1255: of active fields that are advected by a turbulent velocity field,
1256: and the statistics of an auxiliary passive field that is advected
1257: by the same velocity, but does not affect it. The two
1258: examples were akin to turbulent convection and to magneto-hydrodynamics
1259: respectively. In the first example the conserved variables for
1260: the equations of the passive and active fields are the same. For the
1261: second example the active problem exhibits additional conservation laws.
1262: This was shown to be very significant in determining the respective
1263: statistical physics of the two problems.
1264: 
1265: The two examples appear very different in superficial
1266: examination. In the first example the even-order statistics of the
1267: passive and active fields turned out to be the same up to a single
1268: multiplicative factor $\beta$, common to all orders. The forced structure
1269: functions of the active field scale with exactly the same exponents
1270: as the passive field, which in turn are dominated by the leading zero modes
1271: of the decaying problem. We analyzed in detail the source of the
1272: multiplicative factor $\beta$ and showed that it stems from the
1273: additional correlation effects between the forcing and the velocity field
1274: that are absent in the passive case.
1275: Nevertheless these correlation effects do not cause a change in the
1276: scaling exponents. The general lesson that we would propose on the
1277: basis of this example is that whenever there exist a problem in
1278: which the equation of motion of the active field does not satisfy
1279: additional conservation
1280: laws compared to the passive case, the former field
1281: will exhibit structure functions that are dominated by the leading
1282: zero modes of the latter.  This point is also pertinent to the second
1283: example.
1284: Here the active equations possess additional conservation laws,
1285: and indeed the active and passive fields exhibit different scaling
1286: exponents.
1287: Nevertheless we argued that the structure functions of the active field
1288: are still dominated by the zero modes of the passive problem, but
1289: not the leading ones. The additional conservation laws results in
1290: exact cancellations in the contributions of the leading zero modes,
1291: and the active problem lands on the next allowed sub-leading zero mode
1292: of the passive problem.
1293: 
1294: As a generalization, consider then a sufficiently
1295: turbulent velocity field which advects an active field, scalar or
1296: vector, which in its turn is forced by a force having a compact
1297: support in $\B.k$ space. An auxiliary passive field which is advected
1298: by the same velocity field can be employed to find the
1299: zero modes of the operator involved in the passive decay
1300: problem. On the basis of the intuition gained
1301: with the examples presented above, we offer the following tentative
1302: conjecture: the forced structure function of the active field
1303: will exhibit scaling exponents that are the scaling exponents of the
1304: aforementioned 
1305: zero modes. Whenever the conservation laws of the active and passive
1306: problems coincide, these will be the exponents of the leading zero modes.
1307: When the active problem has additional conservation laws, these will be
1308: the next-leading zero modes, as allowed by the conservation laws.
1309: 
1310: Finally, we need to consider the relation of our shell models
1311: to the physical problems and the PDE's that motivate these models. It is
1312: important to test the
1313: conjecture stated here in that context. In light of the above discussion
1314: we expect that much of what has been found here will translate literally
1315: to the continuous problems. After all, the crucial aspects are the linearity
1316: of the advection equation, and the existence of conservation laws. These
1317: are unchanged in the continuous problems. Of course, one can expect
1318: many more numerical difficulties, especially due to the role of angles
1319: in the multi-point correlation functions. Nevertheless, the idea that
1320: the understanding of the anomalous scaling exponent of active fields boils
1321: down to the analysis of eigenfunctions of a linear operator is expected
1322: to hold verbatim.
1323: %%%%%%%%%%%%%%%%%%%%%%%%%%%
1324: \acknowledgments
1325: This work has been supported in part by the
1326: Research Grants Council of Hong Kong SAR, China
1327: (CUHK 4119/98P and CUHK 4286/00P), the European Commission
1328: under a TMR grant, by the Minerva Foundation, Munich, Germany, the German
1329: Israeli Foundation, and the Naftali and Anna Backenroth-Bronicki Fund for
1330: Research in Chaos and Complexity. TG thanks the Israeli Council for Higher
1331: Education and the Feinberg 
1332: postdoctoral Fellowships program at the WIS for financial support.
1333: \begin{thebibliography}{99}
1334: 
1335: \bibitem{71MY}
1336: A. S. Monin and A. M. Yaglom, {\em Statistical Fluid Mechanics},
1337: (MIT, Cambridge  1971).
1338: 
1339: \bibitem{Zeldovich} Ya.~B.~Zeldovich, A.~A.~Ruzmaikin and
1340: D.~D.~Sokoloff, {\it Magnetic Fields in Astrophysics}
1341: Gordon and Breach Publ. (1983).
1342: 
1343: \bibitem{93Bis}
1344: D. Biskamp, {\em Nonlinear Magnetohydrodynamics} (Cambridge, Cambridge UK,
1345: 1993)
1346: 
1347: \bibitem{01FGV}
1348: G. Falkovich, K. Gawedzki and M. Vergassola, Rev. Mod. Phys. {\bf 73}, 2001,
1349: and references therein.
1350: 
1351: \bibitem{01CV}
1352: A. Celani and M. Vergassola, Phys. Rev. Lett. {\bf 86}, 424 (2001).
1353: 
1354: \bibitem{01ABCPV}
1355: I. Arad, L. Biferale, A. Celani, I. Procaccia and M. Vergassola,
1356: Phys. Rev. Lett. {\bf 87}, 164502 (2001).
1357: 
1358: \bibitem{01CGP}
1359: Y. Cohen, T. Gilbert and I. Procaccia, Phys. Rev. E, {\bf 65}, 026314
1360: (2002).
1361: 
1362: 
1363: \bibitem{01CMMV}
1364: A. Celani, T. Matsumoto, A. Mazzino and M. Vergassola,
1365: Phys. Rev. Lett. {\bf 88}, 054503 (2002).
1366: 
1367: 
1368: \bibitem{02CCGP}
1369: E. S. C Ching, Y. Cohen, T. Gilbert and I. Procaccia
1370: ``Statistically Preserved Structures and
1371: Anomalous Scaling in Turbulent Active Scalar Advection", submitted to 
1372: Europhys. Lett., arXiv:nlin/CD/0111030.
1373: 
1374: \bibitem{Brand}
1375: A. Brandenburg, Phys. Rev. Lett., {\bf 69}, 605 (1992).
1376: 
1377: \bibitem{98GC}
1378: P. Giuliani and V. Carbone, Europhys. Lett. {\bf 43}, 527 (1998).
1379: 
1380: \bibitem{98LPPPV}
1381: V. S. L'vov, E. Podivilov, A. Pomyalov, I. Procaccia
1382: and D. Vandembroucq, Phys. Rev. E{\bf 58},
1383: 
1384: 
1385: 
1386:  1811 (1998).
1387: \end{thebibliography}
1388: 
1389: \end{document}
1390: 
1391: 
1392: %\begin{widetext}
1393: %\begin{equation}
1394: %\tilde{a}\lambda(1 - \lambda^{(\beta_3-1)/3})\tilde Q_3(k_{n+1})
1395: %+
1396: %\tilde{b}(\lambda^{-(\beta_3-1)/3}-\lambda^{(\beta_3-1)/3})\tilde Q_3(k_n)
1397: %+\tilde{c}\lambda^{-1}(\lambda^{-(\beta_3-1)/3}-1)\tilde Q_3(k_{n-1}) = 0\
1398: ,
1399: %\label{eqQ3b}
1400: %\end{equation}
1401: %\end{widetext}
1402: 
1403: 
1404: %\begin{equation}
1405: %\tilde{a}\lambda(1 - \lambda^{(\eta_3-\zeta_3)/3})Q_3(k_{n+1})+
1406: %\tilde{b}(\lambda^{-(\eta_3-\zeta_3)/3}-\lambda^{(\eta_3-\zeta_3)/3})Q_3(k_
1407: n)
1408: %+\tilde{c}\lambda^{-1}(\lambda^{-(\eta_3-\zeta_3)/3}-1)Q_3(k_{n-1}) = 0\ ,
1409: %\label{eqQ3b2}
1410: %\end{equation}
1411: %Substituting the scaling form of $Q_3$ in this
1412: %expression gives 
1413: 
1414: