nlin0303006/SPS.tex
1: \documentclass[pre,aps,twocolumn]{revtex4}
2: \usepackage{epsfig,psfrag}
3: \usepackage{bm}
4: \newcommand{\B}[1]{{\bm{#1}}} \newcommand{\C}[1]{{\mathcal{#1}}}
5: \newcommand{\Onecol} {\begin{widetext} \onecolumngrid} %% 2 -> 1
6: \newcommand{\Twocol} {\end{widetext} \twocolumngrid} %% 1 -> 2
7: \newcommand{\ds}{\displaystyle}
8: \newcommand{\be}{\begin{equation}}
9: \newcommand{\ba}{\begin{array}} \newcommand{\bea}{\begin{eqnarray}}
10: \newcommand{\bfi}{\begin{figure}} \newcommand{\ee}{\end{equation}}
11: \newcommand{\ea}{\end{array}} \newcommand{\eea}{\end{eqnarray}}
12: \newcommand{\efi}{\end{figure}} \newcommand{\dV}{\delta u}
13: \newcommand{\T}{\theta} \newcommand{\lp}{\left(} \newcommand{\rp}{\right)}
14: \newcommand{\ra}{\right\rangle} \newcommand{\la}{\left\langle}
15: % \def\Fbox#1{\vskip1ex\hbox to 8.5cm{\hfil\fboxsep0.3cm\fbox{%
16: %\parbox{8.0cm}{#1}}\hfil}\vskip1ex\noindent}  %%  {TEXT} in BOX
17: %\input REVTeXDraft1.tex
18:  \begin{document}
19: \title{Eulerian Statistically Preserved Structures in Passive Scalar Advection}
20:   \author{Yoram Cohen$^{1,2}$ Anna Pomyalov$^1$ and
21:   Itamar Procaccia$^{1,2}$}
22:   \affiliation{$^1$ Dept. of Chemical Physics, The Weizmann Institute of
23:   Science, Rehovot 76100, Israel.\\
24: $^2$ Dept. of Physics, the Chinese University of Hong Kong, Shatin, Hong Kong.
25: } \pacs{}
26: \begin{abstract}
27: We analyze numerically the time-dependent linear operators that
28: govern the dynamics of Eulerian correlation functions  of a
29: decaying passive scalar advected by a stationary, forced
30: 2-dimensional Navier-Stokes  turbulence.  We show how to naturally
31: discuss the dynamics in terms of effective compact operators that
32: display Eulerian Statistically Preserved Structures which
33: determine the anomalous scaling of the correlation functions. In
34: passing we point out a bonus of the present approach, in
35: providing analytic predictions for the time-dependent correlation
36: functions in decaying turbulent transport.
37: \end{abstract}
38: \maketitle
39: 
40: \section{Introduction}
41: The aim of this paper is to discuss the statistical physics of turbulent advection
42: of passive scalars \cite{79MY}. We are interested in scalar fields $\theta(\B r,t)$ which are
43: advected by a velocity field $\B u(\B r,t)$ such that together they solve
44: the set of equations
45: \begin{eqnarray}
46: \frac{\partial \B u}{\partial t}&+&(\B u\cdot \B \nabla)\B u=
47: -\B \nabla p +\nu \Delta \B u+\B f\ , \label{NS}\\
48: \frac{\partial \theta}{\partial t} &+&(\B u\cdot \B \nabla)\theta=
49:  \kappa \Delta \theta+f \ .\label{PS}
50: \end{eqnarray}
51: In these equations $p(\B r,t)$, $\nu$ and $\kappa$ are the pressure field, the kinematic viscosity
52: and the scalar diffusivity respectively. In this paper
53: the Navier-Stokes equations will be always forced
54: with a time-stationary forcing $\B f(\B r,t)$.
55: The scalar field will be forced or unforced (decaying), depending on our interests below.
56: We focus on the case of high Reynolds number Re and high Peclet number Pe, i.e.
57: $\nu, \kappa \to 0$, where the turbulence of the velocity field is fully developed and where
58: both fields display a significant range of scaling behavior at scales that are
59: sufficiently far from the forcing scales.
60: 
61: A major theoretical question that had been answered recently has to do
62: with the scaling properties of the correlation functions of the advected field \cite{98BGK,01CV}.
63: Define the simultaneous many-point correlation function $F^{(N)}(\B r_1, \dots \B r_N)$
64: in the forced case by
65: \begin{equation}
66: F^{(N)}(\B r_1, \dots \B r_N) \equiv \langle \theta(\B
67: r_1,t)\theta(\B r_2,t)\cdots \theta(\B r_N,t)\rangle_f \ ,
68: \label{defFn}
69: \end{equation}
70: with $\langle \cdots \rangle_f$ denoting an average with respect to realizations of the
71: stationary forcing $f(\B r,t)$ and of the velocity field $\B u (\B r,t)$.
72: It had been known for a long time that for high Re and Pe these functions
73: are homogeneous functions of their arguments, i.e.
74: \begin{equation}
75: F^{(N)}(\lambda \B r_1, \dots \lambda \B r_N) =\lambda^{\zeta_N} F^{(N)}(\B r_1, \dots \B r_N) \ ,
76: \end{equation}
77: with $\zeta_N$ being a scaling exponent that in general is {\em anomalous}, i.e. cannot be
78: guessed from dimensional considerations. But only recently it became clear how
79: these exponents are determined by the dynamical processes.
80: 
81: To understand the progress made, we rewrite the dynamical
82: equation for $\theta (\B r,t)$ in the shorthand form
83: \begin{equation}
84: \frac{\partial \theta(\B r,t)}{\partial t}=\C L \theta(\B
85: r,t)+f(\B r,t) \ , \label{defL}
86: \end{equation}
87: where in the present case $\C L\equiv \B u\cdot \B \nabla-\kappa
88: \Delta$.  In recent work \cite{01ABCPV,02CGP} it was clarified why
89: and how passive fields exhibit anomalous scaling, when the
90: velocity field is a generic turbulent field. The key is to
91: consider a {\em decaying problem} associated with Eq.
92: (\ref{defL}), in which the forcing $f(\B r,t)$ is put to zero.
93: The problem becomes then a linear initial value problem,
94: \begin{equation}
95: \partial \theta/\partial t =\C L \theta\ , \label{decay}
96: \end{equation}
97: with a formal solution
98: \begin{equation}
99: \theta(\B r,t) = \int d\B r' R(\B r,\B r',t) \theta(\B r',0) \
100: , \label{oper}
101: \end{equation}
102: with the operator
103: \begin{equation}
104: R(\B r, \B r', t)\equiv T^+ \exp[{\int_0^t ds \C L(s)}]\Big|_{\B r, \B r'}\ , \label{defR}
105: \end{equation}
106: and $T^+$ being the time ordering operator.
107: Define next the {\em time dependent} correlation
108: functions of the decaying problem:
109: \begin{equation}
110: C^{(N)}(\B r_1,\cdots ,\B r_N,t)\equiv \langle \theta(\B
111: r_1,t)\cdots \theta(\B r_N,t)\rangle \ . \label{defG}
112: \end{equation}
113: Here pointed brackets without subscript $f$ refer to the decaying
114: object in which averaging is taken with respect to realizations of
115: the velocity field and initial conditions. As a result of Eq. (\ref{oper}) the decaying
116: correlation functions are evolved by a propagator
117: $\C P^{(N)}(\underline{\B r},\underline{\B \rho},t)$, (with
118: $\underline {\B r}\equiv \B r_1,\B r_2,\!\cdots\!,\B r_N$ and $\underline {\B \rho}\equiv \B \rho_1,\B \rho_2,\!\cdots\!,\B \rho_N$)~:
119: \begin{equation}
120: C^{(N)}(\underline{\B r},t)=\int\!\! d\underline{\B \rho}
121:  \C P^{(N)}(\underline{\B r},\underline{\B \rho},t)~
122: C^{(m)}(\underline{\B \rho},0) \ . \label{defprop}
123: \end{equation}
124: 
125: In writing this equation we made explicit use of the fact that
126: the {\em initial} distribution of the passive field $\theta(\B r,0)$
127: is statistically independent of the advecting velocity field. Thus
128: the operator $\C P^{(N)}(\underline{\B r},\underline{\B \rho},t)$
129: can be written explicitly
130: \begin{equation}
131: \C P^{(N)}(\underline{\B r},\underline{\B \rho},t)\equiv
132: \langle  R(\B r_1,\B \rho_1,t)  R(\B r_2,\B \rho_2,t)\cdots
133: R(\B r_N,\B \rho_N,t)\rangle \ .
134: \end{equation}
135: 
136: The key finding \cite{01ABCPV,02CGP} is that the operator $\C
137: P^{(N)}(\underline{\B r},\underline{\B \rho},t)$ possesses a {\em
138: left} eigenfunction of eigenvalue 1, i.~e. there exists (for each
139: $N$) a time-independent function $Z^{(N)}(\underline{\B r})$ satisfying
140: \begin{equation}
141: Z^{(N)}(\underline{\B r})=\int d\underline {\B \rho}
142: Z^{(N)}(\underline{\B \rho}) \C P^{(N)}
143: (\underline{\B r},\underline{\B \rho},t)\ .\label{zeromode}
144: \end{equation}
145: The functions $Z^{(N)}$ are referred to as ``Statistically Preserved
146: Structures'', being invariant to the dynamics, even though
147: {\em the operator is strongly time dependent and decaying}. How to form,
148: from these functions, infinitely many conserved variables in the decaying
149: problem was shown in \cite{01ABCPV}, and is discussed again in
150: Sect. III. The functions
151: $Z^{(N)}(\underline{\B r})$ are homogeneous functions of
152: their arguments, with anomalous scaling exponents $\zeta_N$:
153: \begin{equation}
154: Z^{(N)}(\lambda\underline{ \B r}) = \lambda^{\zeta_N}
155: Z^{(N)}(\underline{\B r})+\dots
156: \end{equation}
157: where ``$\dots$'' stand for sub-leading scaling terms. Since
158: Eq.~(\ref{zeromode}) contains $Z^{(N)}(\underline{\B r})$ on both
159: sides, the scaling exponent $\zeta_N$ cannot be determined from
160: dimensional considerations, and it can be anomalous. More
161: importantly, it was shown that the correlation functions of the
162: forced case, $F^{(N)}(\underline{\B r})$ Eq. (\ref{defFn}), have
163: exactly the same scaling exponents as $Z^{(N)}(\underline{\B r})$
164: \cite{02CGP}. In the scaling sense
165: \begin{equation}
166: F^{(N)}(\underline{\B r})\sim Z^{(N)}(\underline{\B r}) \ .
167: \label{scalesame}
168: \end{equation}
169: This is how anomalous scaling in passive fields is understood.
170: 
171: Besides exactly soluble examples in which the advecting velocity
172: field is non-generic (i.e. $\delta$-correlated in time) the
173: existence of eigenfunctions of eigenvalue 1 of the time dependent
174: propagator was demonstrated fully only in shell models of
175: turbulence. While the present authors believe that shell models
176: contain a lot of the robust properties of real turbulence, this
177: belief is not universally accepted in the community. It is
178: therefore necessary to demonstrate that the mechanism sketched
179: above exists indeed in the full problem Eqs.
180: (\ref{NS})-(\ref{PS}). This had been done for 3'rd order
181: correlations within the Lagrangian formulation in \cite{01CV}.
182: The aim of this paper is to demonstrate this in the Eulerian
183: frame, and for correlation functions of order 2, 4 and 6.
184: 
185: In Section II we describe the simulations of Eqs. (\ref{NS}) and
186: (\ref{PS}) and the type of measurements that we performed. Sect.
187: III is a theoretical digression, in which we analyze an exactly
188: soluble simple model to guide ourselves as to how to analyze the
189: numerical results to find the scaling forms of the $n$th oder
190: propagators and their eigenfunctions. Sect. IV describes the
191: analysis of the data, and Sect. V offers a discussion and a
192: summary of the paper.
193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
194: \section{Simulations}
195: We performed a Direct Numerical Simulation (DNS) of Eqs.
196: (\ref{NS}) and (\ref{PS}) on a $2048\times 2048$ 2-D array. The
197: forcing $\B f$ in Eq. (\ref{NS}) is random, $\delta$ correlated
198: in time, isotropic and homogeneous. Its $k$ depdendence is
199: \begin{equation}
200: \langle|\B f(\B k)|\rangle \propto k\;exp[-0.5(k/1024)^2]\ .
201: \end{equation}
202: This forcing is biased towards the small scales; this is done
203: because of the inverse energy cascade that characterizes
204: 2-dimensional turbulence. The fluid dissipation is modeled by a
205: hyper-viscosity  term proportional to $\Delta^8 \B u$. In
206: addition we employed  a ``friction'' term proportional to $\B u$
207: in order to stabilize the velocity field on the largest scales.
208: The passive field $\theta$ dissipates normally as shown in Eq.
209: (\ref{PS}).
210: 
211: The simulations were performed for a decaying passive field
212: $\theta$, that is, the forcing $f$ in Eq. (\ref{PS}) was put to
213: zero. The initial conditions for the $\theta$ field were of the
214: form
215: \begin{equation}\label{init}
216: \tilde \theta(\B k,t=0)=\delta(k-k_0)exp(i\gamma(\B k))\, ,
217: \end{equation}
218: where $\tilde \theta(\B k,t)$ is the Fourier transform of the
219: real space variable defined as:
220: \begin{equation}
221: \tilde h(\B k)=\frac{1}{2\pi}\int d \B r h(\B r) e^{-i \B k \B r}
222: \end{equation}
223: (from now on we will omit the tilde above the functions and denote the k
224: space functions only by their variables).
225: $\gamma(\B k)$ is a random variable in the interval
226: $\gamma\in [0,2 \pi]$ where $\gamma(\B k)=-\gamma(-\B k)$, insuring that
227: $\theta^*(-\B k,0)=\theta(\B k,0)$, and therefore that $\theta(\B
228: r,t)$ is real.
229: 
230: As the initial conditions for $\theta$ and the forcing of the $\B
231: u$ fields are both homogeneous and isotropic so are the scalar
232: correlation functions $C^{(N)}(\B r_1 \dots \B r_N,t)$ defined in
233: Eq. (\ref{defG}), at all times $t$. We measured the $\B k$ space
234: correlation functions:
235: \begin{eqnarray}
236: &&C^{(N)}(\B k_1,\dots,\B k_{N},t)\delta(\B k_1+\dots+\B
237: k_{N})\nonumber\\ &&=\langle \theta(\B k_1,t)\dots \theta(\B
238: k_N,t)\rangle \ , \label{defCm}
239: \end{eqnarray}
240: where again the average $\langle \dots\rangle$ is over initial
241: conditions and over realizations of the $\B u$ field. The delta
242: function appears due to translational invariance.
243: 
244: In accordance to Eq. (\ref{defprop}) we define the $\B k$ space propagator:
245:  \begin{equation}
246: C^{(N)}(\underline{\B k},t)=\int\!\! d\underline{\B k}'
247:  \C P^{(N)}(\underline{\B k},\underline{\B k}',t)~
248: C^{(N)}(\underline{\B k}',0) \ . \label{defpropk}
249: \end{equation}
250: 
251: \subsection{The 2-point propagator}
252: The only two point correlator that is not zero is $C^{(2)}(\B
253: k,-\B k,t)=\langle\theta(\B k)\theta^*(\B k)\rangle$. Because of
254: the isotropy of the initial conditions and the driving field, the
255: correlator depends only on the magnitude of $\B k$. We can
256: therefore consider the second order structure function
257: $S^{(2)}(k,t)=C^{(2)}(\B k,-\B k,t)$, and it's propagator $\hat \C
258: P^{(2)}(k,k',t)$.
259: 
260: In discrete $k$-space (as in our simulation on a grid) the
261: propagator $\hat \C P^{(2)}(k,k',t)$ has a matrix representation.
262: For the choice of initial conditions as in Eq. (\ref{init}),
263: $S^{(2)}(k,t)$ is simply the $k_0$th column of the
264: propagator :
265: \begin{equation}
266: S^{(2)}(k,t)=\int\!\! dk'
267:  \hat \C P^{(2)}(k,k',t)~
268: S^{(2)}(k',0) =\hat \C P^{(2)}(k,k_0,t)\ . \label{defpropd2}
269: \end{equation}
270: In Figure \ref{fig1} we plot such a column of the 2-point
271: propagator, at 10 different times. Two properties of the
272: propagator should be noticed: as time progresses the overall
273: amplitude decreases due to the decay, while the maximum moves
274: from the initial $k_0$ to lower values of $k$. In Sects. III and
275: IV we will find the scaling form of this propagator.
276: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
277: \begin{figure}
278: \centering
279: \psfrag{yl}{$\hat\C P^{(2)}(k,300,t)$}
280: \psfrag{xl}{$k$}
281: \includegraphics[width=.45\textwidth]{SPSfig1.eps}
282: \caption{$\hat \C P^{(2)}(k,300,t)$ for 10 different times. The
283: first time is $\tau_{30}$, i.e. the eddy turn over time at scale
284: $k=30$, in lightest grey. Later times, in darker and darker greys,
285: are at 2$\tau_{30}$, 3$\tau_{30}$, etc. until 10$\tau_{30}$.}
286: \label{fig1}
287: \end{figure}
288: %%%%%%%%%%%%%%%%%%%%%%%%
289: \subsection{Multi-point propagators}
290: In the case of the multi-point correlator the overall $\delta$
291: function and the isotropy of the fields do not reduce the
292: dependence to a one variable function. The propagators are
293: therefore functions of many $k$-vectors. It turns out however that
294: measurement of the statistics for objects depending on many
295: $k$-vectors are very taxing. We opted therefore the extract from
296: the DNS partial information on the dynamics of the $2N$ order
297: structure functions with $N>1$:
298: \begin{equation}\label{gfus}
299: S^{(2N)}(k,t)=\langle|\theta(\B k,t)|^{2N}\rangle \ .
300: \end{equation}
301: Accordingly we define the reduced propagators :
302: \begin{eqnarray}
303: S^{(2N)}(k,t)&=&\int\!\! dk'
304:  \hat \C P^{(2N)}(k,k',t)~
305: S^{(2N)}(k',0) \ .
306: \label{defpropdmult}
307: \end{eqnarray}
308: 
309: In Figure \ref{fig2} we present the numerical
310: results for of the fourth and sixth order reduced propagators, at
311: the same ten different times. These objects display similar
312: qualitative behavior to that of the 2-point propagator. We will
313: learn in the next section how to think about the scaling
314: properties of these objects and how to re-plot the numerical
315: results in proper rescaled variables.
316: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
317: \begin{figure}
318: \centering
319: \psfrag{yl}{$\hat\C P^{(4)}(k,300,t)$}
320: \psfrag{xl}{$k$}
321: \includegraphics[width=.45\textwidth]{SPSfig2.eps}
322: \psfrag{yl}{$\hat\C P^{(6)}(k,300,t)$}
323: \psfrag{xl}{$k$}
324: \includegraphics[width=.45\textwidth]{SPSfig3.eps}
325: \caption{$\hat \C P^{(4)}(k,300,t)$ (upper panel) and $\hat \C
326:   P^{(6)}(k,300,t)$ (lower panel) for the same 10 different
327: times as in Fig. \ref{fig1}} \label{fig2}
328: \end{figure}
329: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
330: \section{An Exactly solvable case}
331: In order to motivate the data analysis presented in Sect. IV we
332: turn now to the Kraichnan model for passive scalar advection. The
333: model is exactly solvable, and examining the analytic forms of the
334: propagators offers clues to what should be expected in the
335: generic case. In principle we could perform this analysis in
336: terms of the Kraichnan model for a 2-dimensional passive scalar.
337: In fact, it is sufficient to consider the shell-model version;
338: the latter is very transparent to analytic manipulations, and for
339: our purposes the results throw equally useful light on how to
340: perform the analysis of the DNS results.
341: 
342: \subsection{The Kraichnan shell model}
343: The Kraichnan shell model for passive scalar advection \cite{97BBW,01ABCPV}, as all
344: other shell models of turbulent flows, is written in terms of
345: Fourier components of the field. The Fourier components are
346: restricted to shells denoted by the index $m$, and the equation
347: takes on the form
348: \begin{eqnarray}
349: {d\theta_m\over dt}&=&\C L_{m,n}\theta_n \label{ldef}\\
350: \C L_{m,n}&=&ik_{m+1}u_{m+1}\delta_{m+1,n}
351: +ik_m u^*_{m}\delta_{m-1,n}-\kappa k_m^2\delta_{m,n}\ .\nonumber
352: \label{pshell}
353: \end{eqnarray}
354: Here $k_m$ are the shell $k$ vector, $k_m=k_0 \lambda^m$ for some
355: $k_0$ and $\lambda$. The shell components of the velocity field,
356: $u_m(t)$, are Gaussian random variables, $\delta$-correlated in time,
357: satisfying:
358: \begin{eqnarray}\label{Vcort}
359: \langle u_n(t)u^*_m(t')
360: \rangle=c_0~\delta_{n,m}\delta(t-t')~\lambda^{-\xi m} \ .
361: \end{eqnarray}
362: Here $\xi$ is the scaling exponent of the $u$ field.
363: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
364: \subsection{The 2nd order propagator}
365: The second order correlator satisfies the explicit equation (see
366: \cite{02CGP}):
367: \begin{equation}\label{kri}
368: \frac{d}{dt}S^{(2)}_n(t) =\Big(M^{(2)}_{n,m} -\kappa
369: k_n^2\Big)S^{(2)}_m(t) \, ,
370: \end{equation}
371: $S^{(2)}_n(t)$ is the shell equivalent of the 2nd order structure
372: function in Eq. (\ref{defpropd2}),
373: \begin{equation}\label{kri1}
374: S^{(2)}_n(t)=\langle|\theta_n(t)|^2\rangle \ .
375: \end{equation}
376: The time evolution operator $\B M^{(2)}$ has the form
377: \begin{eqnarray}\label{krim1}
378: M^{(2)}_{n,m}&=& (\alpha_n+\alpha_{n+1})\delta_{n,m}
379: \\&-&\alpha_n\delta_{n-1,m}
380: -\alpha_{n+1}\delta_{n+1,m}\ ,\nonumber
381: \end{eqnarray}
382: where
383: \begin{equation}
384: \alpha_n\equiv -c_0k_0^2 \left(\frac{k_n}{k_0}\right)^{\zeta_2}
385: =-c_0k_0^2 \lambda^{\zeta_2 n} \ , \label{alphalam}
386: \end{equation}
387: and $\zeta_2=2-\xi$ is the dimensional scaling of the 2-point
388: correlation function. The operator $\B M^{(2)}$ has the following useful
389: scaling property:
390: \begin{equation}
391: \lambda^{-\zeta_2 p} M_{n+p,m+p}^{(2)}=M_{n,m}^{(2)} \ .
392: \label{propM}
393: \end{equation}
394: The second order propagator has, in the limit of vanishing
395: viscosity, the explicit form:
396: \begin{equation}\label{krip}
397: \C P^{(2)}_{n|m}(t)=\exp(t\B M^{(2)})\Big|_{n,m}\ .
398: \end{equation}
399: 
400: Note that the time evolution operator and the propagator are both
401: Hermitian, and thus admit an eigenvector decomposition. The time
402: evolution operator has two types of eigen-vectors $\psi^{(2,q)}_n$ and
403: $\tilde \psi^{(2,q)}_n$
404: which we can regard as slow modes and fast modes respectively. Here
405: the index $k$ stands for the eigen-mode index. The fast modes are
406: dominated by the viscous term, their support is in the viscous range,
407: and they can essentially be taken to be $\tilde \psi^{(2,q)}_n=\delta_{n,q}$
408: for some shell $q$  with the shell index $q > m_d$, above which viscosity
409: dominates Eq. (\ref{kri}). The transition shell $m_d$ is determined by the
410: condition $\kappa k^2_{m_{d}}= c_0k^{\zeta_2}_{m_{d}}$.
411: For the fast modes we have the following approximate Equation:
412: \begin{equation}
413: \frac{d}{dt}\tilde \psi^{(2,q)}_n = -\kappa k_n^2\tilde \psi^{(2,q)}_n \, ,
414: \end{equation}
415: therefore their eigen-values are $\beta_q =-\kappa k_q^2$.
416: 
417: For slow modes, $\psi^{(2,q)}_n$,  which have their support on
418: shells smaller than $m_d$, the dissipative term can be neglected
419: and they satisfy an eigen-function equation of the form:
420: \begin{equation}
421: \beta_q \psi^{(2,q)}_n=M^{(2)}_{n,m}\psi^{(2,q)}_m \ .
422: \end{equation}
423: 
424: For a sufficiently large inertial range we can use the scaling
425: property (\ref{propM}) to get:
426: \begin{equation}
427: \beta_q \psi^{(2,q)}_n=\lambda^{-\zeta_2p}
428: M^{(2)}_{n+p,m+p}\psi^{(2,q)}_{m} \ .
429: \end{equation}
430: Shifting indices we rewrite,
431: \begin{equation}
432: \beta_q \psi^{(2,q)}_{n-p}=\lambda^{-\zeta_2p}
433: M^{(2)}_{n,m}\psi^{(2,q)}_{m-p}\ .
434: \end{equation}
435: Define now a vector
436: \begin{eqnarray}\label{sl1}
437: \chi_{n}\equiv\psi^{(2,q)}_{n-p} \, ,
438: \end{eqnarray}
439: we have:
440: \begin{eqnarray}
441: \lambda^{\zeta_2 p}\beta_q \chi_{n}&=&M^{(2)}_{n,m}\chi_{m} \ ,
442: \end{eqnarray}
443: We can thus define $\psi^{(2,q+p)}_{n}\equiv \chi_n$,  an
444: eigen-function of the time evolution operator $\B M^{(2)}$ with
445: eigen-value $\lambda^{\zeta_2 p} \beta_q$.
446: 
447: We can thefore conclude that the eigen-vectors and eigen-values may
448: be obtained from each other by shift of the indices:
449: \begin{eqnarray}
450: \psi^{(2,q+p)}_{n+p}&=&\psi^{(2,q)}_{n} \label{psi_shift}\ ,\\
451: \beta_{q+p}&=&\lambda^{\zeta_2 p}\beta_{q}  \label{beta_shift}\ . 
452: \end{eqnarray}
453: 
454: Fig.~\ref{multi1} demonstrates that indeed two different
455: eigen-functions coincide once shifted with respect to each other.
456: It follows from (\ref{beta_shift}) that
457: \begin{eqnarray}\label{2pa}
458: \beta_{q}\propto \alpha_q.
459: \end{eqnarray}
460: 
461: In Fig.~\ref{multi2} the spectrum of the eigen-values is plotted,
462: showing the two expected regions: a region for which the scaling
463: of the eigen-values is $\propto k_n^{\zeta_2}$  for the slow
464: modes, and  $\propto k_n^{2}$ for the fast modes.
465: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
466: \begin{figure}
467: \centering
468: \psfrag{xl}{$n$}
469: \psfrag{yl}{$\psi^{(2,15)}_n,\,\psi^{(2,35)}_{n+20}$}
470: \includegraphics[width=.5\textwidth]{SPSfig4.eps}
471: \caption{A plot of the eigen-vectors
472:  $\psi^{2,15}_n$ (squares), and $\psi^{2,35}_{n+20}$ (triangles),
473:  for $n\in[0,30]$  }
474: \label{multi1}
475: \end{figure}
476: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: \begin{figure}
479: \centering
480: \psfrag{xl}{$n$}
481: \psfrag{yl}{$\alpha_n$}
482: \includegraphics[width=.5\textwidth]{SPSfig5.eps}
483: 
484: \caption{The eigen-values, plotted from small to large. The
485: eigen-values are well fitted by the analytic predictions
486: $k_m^{\zeta_2}$ (dashed line) for the slow modes, and by
487: $k_m^{2}$ (dotted line) for the fast modes. The transition to viscous
488: range occur at shell $m_d=80$ }
489: \label{multi2}
490: \end{figure}
491: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
492: The lowest eigenvalue is proportional to $\alpha_0$. The
493: eigen-function $\psi^{(2,0)}_n$ associated with it can be calculated
494: explicitly, and it exhibits a normal, dimensional, scaling
495: in the inertial range:
496: \begin{eqnarray}\label{zmod}
497: \psi^{(2,0)}_n \propto (k_n/k_0)^{-\zeta_2} \ .
498: \end{eqnarray}
499: Using  Eqs .(\ref{psi_shift}) and (\ref{zmod}), that 
500: imply that $\psi^{(2,q)}_n$ has a scailing 'tail' with exponent $\zeta_2$, 
501: starting after shell $q$ we establish:
502: \begin{eqnarray}
503: \psi^{(2,q)}_n\propto
504: (k_n/k_q)^{-\zeta_2}\propto\alpha_q/\alpha_n\, \quad q<n<m_d \ ,
505: \end{eqnarray}
506: 
507: We now use these results to learn how to re-scale our numerical
508: data. Suppose that we started with an initial condition in some
509: shell $n$ in the inertial range, $S^{(2)}_p(0)=\delta_{p,n}$. The
510: fast modes do not contribute significantly to the time
511: dependence, since they decay fast and anyhow have no support in
512: the inertial range. Thus, using only the slow modes and Eqs.
513: (\ref{krim1}),(\ref{krip}) , we have for $p<n$,
514: \begin{eqnarray} \label{phi1}
515: S^{(2)}_p(t)&=&\sum_{q=1}^{p}\psi^{(2,q)}_n
516: e^{\alpha_qt}\psi^{(2,q)}_p\nonumber\\&\propto& \frac{1}{\alpha_p
517: \alpha_n}\sum_{q=1}^{p}\alpha_q^2e^{\alpha_qt}\propto {\cal
518: P}^{(2)}_{p|n}(t) \ .
519: \end{eqnarray}
520: In going from the first to the second line we inserted the
521: $\delta$-function initial conditions, getting thus a column of the
522: propagator ${\cal P}^{(2)}_{p|n}(t)$. Remember that in our DNS
523: calculations we fixed the column index, and are interested in the
524: scaling behavior with respect to the row index and time. Therefore
525: we compute now (up to an overall dimensional constant),
526: \begin{eqnarray}\label{phi2}
527: &&\lambda^{\zeta_2 m} {\cal P}^{(2)}_{p-m|n}(\lambda^{\zeta_2 m}
528: t)= \frac{\lambda^{\zeta_2 m}
529: }{\alpha_{p-m}\alpha_n}\sum_{q=1}^{p-m}\alpha_q^2e^{\alpha_{q+m}t}\nonumber\\
530: &&=\frac{1}{\alpha_p\alpha_n}\sum_{q=1}^{p-m}\alpha_{q+m}^2e^{\alpha_{q+m}t}
531: =\frac{1}{\alpha_p\alpha_n}\sum_{q=m+1}^{p}\alpha_{q}^2e^{\alpha_{q}t}\nonumber
532: \\
533: &&= {\cal P}^{(2)}_{p|n}(t)-\frac{\alpha_m}{\alpha_p}{\cal
534: P}^{(2)}_{m|n}(t) \ ,
535: \end{eqnarray}
536: where we have used the explicit form of the propagator in Eq. (\ref{phi1})
537: and Eq. (\ref{alphalam}). For $m$ much smaller
538: than $p$ we can neglect the second term. Then we conclude that
539: the propagator is a homogeneous function of the variables $k_p$
540: and $t$. Explicitly, multiplying by $t$,
541: \begin{equation}\label{prp2}
542: t{\cal P}^{(2)}(k_p,k_n,t)=(\lambda^{-\zeta_2 m}t){\cal
543: P}^{(2)}(\lambda^m k_p,k_n,\lambda^{-\zeta_2 m} t)
544: \end{equation}
545: For fixed $k_n$ this is a homogeneous function of two arguments
546: which can be always written in the form
547: \begin{equation}
548: {\cal P}^{(2)}(k_p,k_n,t) = \frac{\text{Const}}{\alpha_n t}
549: \Lambda(\alpha_p t) \ ,
550: \end{equation}
551: for some function $\Lambda(x)$. The symmetry between $p$ and $n$
552: is restored by realizing that the asymptotic form of $\Lambda(x)$ is
553: $1/x$ for large $x$.
554: 
555: This scaling form was also found in the study of the generic shell
556: models of passive scalar advection (arbitrary time dependence in the
557: velocity correlations) \cite{02CGP} and we will show that the same scaling form
558: survives when we go back to the generic model studied by DNS.
559: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
560: \subsection{The multi point propagators}
561: The invariance of Eq. (\ref{pshell}) under a uniform phase change
562: $\theta_n\rightarrow e^{i\phi}\theta_n$ dictates that the only
563: non-zero correlation functions will have an equal number of
564: variables $\theta_n$ and conjugated variables $\theta^*_m$. We
565: define:
566: \begin{equation}\label{gkmult}
567: C^{(2N)}_{i_1,\dots,i_m,j_1,\dots,j_N}(t) =\langle
568: \theta_{i_1}(t)\dots \theta_{i_N}(t)\theta^*_{j_1}(t)\dots
569: \theta^*_{j_N}(t)\rangle
570: \end{equation}
571: As in the 2-point case we can define the respective multi-point
572: differential time derivative operator, analogous to Eq.
573: (\ref{kri}). In the limit of vanishing viscosity:
574: \begin{eqnarray}
575: \frac{d}{dt}C^{(2N)}_{\underline{\B i},\underline{\B
576:     j}}(t)=M^{(2N)}_{\underline{\B i},\underline{\B j}|\underline{\B
577:     i}',\underline{\B j}'}C^{(2N)}_{\underline{\B i}',\underline{\B
578:     j}'}(t)
579: \end{eqnarray}
580: The respective propagator is defined by:
581: \begin{eqnarray}
582: C^{(2N)}_{\underline{\B i},\underline{\B j}}(t) &=&\C
583: P^{(2N)}_{\underline{\B i},\underline{\B j}|\underline{\B
584:     i}',\underline{\B j}'}(t)
585: C^{(2N)}_{\underline{\B i}',\underline{\B j}'}(0)\label{prop2m}\\
586: \C P^{(2N)}_{\underline{\B i},\underline{\B j}|\underline{\B
587:     i'},\underline{\B j'}}&=&\exp(t \B M^{(2N)})\Big|_{\underline{\B
588:     i},\underline{\B j}|\underline{\B i}',\underline{\B j}'} \ .
589: \end{eqnarray}
590: 
591: Because of the time reversibility of the statistics of the $\B u$
592: fields, and the anti-hermiticity of the un-averaged differential
593: operator $\C L$ in Eq. (\ref{ldef}): $ \C L^*_{m,n}=-\C L_{n,m}$,
594: both the time derivative operator $\B M^{(2N)}$ and the
595: propagator itself are hermitian, and therefore admit eigenvalue
596: decomposition. Furthermore the operator $\B M^{(2N)}$ has the
597: following scaling property:
598: \begin{eqnarray}
599: M^{(2N)}_{\underline{\B i},\underline{\B j}|\underline{\B i'},\underline{\B
600:     j'}}=
601: \lambda^{-\zeta_2 p}M^{(2N)}_{\underline{\B i}+p,\underline{\B j}
602: +p|\underline{\B i'}+p,\underline{\B j'}+p} \ . \label{defM2m}
603: \end{eqnarray}
604: 
605: As in the two point case the dynamics within the scaling range is
606: determined by slow modes.
607: \begin{equation}\label{4eig}
608: \beta^{(2N,l)}_{k}\psi^{(2N,l,k)}_{\underline{\B i},\underline{\B
609: j }}=M^{(2N)}_{\underline{\B i},\underline{\B j}|\underline{\B
610: i}',\underline{\B j}'}\psi^{(2N,l,k)}_{\underline{\B
611: i}',\underline{\B j}'} \ .
612: \end{equation}
613: In this equation the index $l$ stands for a family of eigen-modes,
614: and $k$ for their index within the family. Each $l$ family can be
615: obtained from any one of its members by shifting the indices. The
616: eigen-values of the modes within a given family can be obtained
617: also by a shift :
618: \begin{eqnarray}
619: \psi^{(2N,l,k+p)}_{\underline{\B i}+p,\underline{\B
620: j}+p}&=&\psi^{(2N,l,k)}_{\underline{\B i}
621: ,\underline{\B j}}\label{gamshift}\\
622: \beta^{(2N,l)}_{k+p}&=&\lambda^{\zeta_2 p}\beta^{(2N,l)}_{k} \ .
623: \label{scalesim}
624: \end{eqnarray}
625: Note that in the 2-point case we had only one family of
626: eigen-modes. Here we added the index $l$ to the eigen-values and
627: eigen-modes to distinguish the different families. Note also that in
628: Eq. (\ref{scalesim}) the scaling exponent of the eigenvalues is
629: $\zeta_2$; this stems from the scaling properties of the
630: differential operator $\B M^{(2N)}$, cf. eq. (\ref{defM2m}).
631: However, the eigen-modes display in general anomalous scaling
632: which can be represented by
633: \begin{eqnarray}
634: \psi^{(2N,l,k)}_{\underline{\B i},\underline{\B j}}
635: =\lambda^{-p~\zeta_{2N,l}}~ \psi^{(2N,l,k)}_{\underline{\B
636: i}+p,\underline{\B j}+p} \ , \label{shift}
637: \end{eqnarray}
638: where $\zeta_{2N,l}$ is the (anomalous) scaling exponent of the
639: $l$th family of eigen-modes.
640: 
641: Since the DNS were analyzed in terms of fused objects, we focus
642: on initial conditions in Eq.(\ref{prop2m}) for which all the $2N$
643: indices are the same. We also measure the resulting structure
644: functions
645: \begin{equation}
646: S^{(2N)}_n=\langle |\theta_n|^{2N}\rangle \ . \label{S2m}
647: \end{equation}
648: This procedure will extract a fused propagator for which there
649: are only two indices, and we denote it bellow as ${\cal
650: P}^{(2N)}_{p|n}(t)$. This propagator is not Hermitian, it has in
651: general no eigen-function decomposition, but we can understand the
652: scaling form of any of its columns from the knowledge of the full
653: propagator and its eigen-functions.
654: 
655: 
656: The equivalent of Eq. (\ref{phi1}) for
657: $S_k^{(2N)}(t=0)=\delta_{k,n}$, $p<n$, summing over all families
658: of slow modes (i.e. over the index $l$) is:
659: \begin{eqnarray} \label{phi1m}
660: &&S^{(2N)}_p(t)=\sum_{l}\sum_{q=1}^{p}\psi^{(2N,l,q)}_n
661: e^{\alpha_qt}\psi^{(2N,l,q)}_p\nonumber\\
662: &&=\sum_{l}\frac{C_l}{(\alpha_p
663: \alpha_n)^{\frac{\zeta_{2N,l}}{\zeta_2}}}
664: \sum_{q=1}^{p}\alpha_q^\frac{2\zeta_{2N,l}}{\zeta_2}e^{\alpha_qt}={\cal P}^{(2N)}_{p|n}(t) \ .
665: \end{eqnarray}
666: In going from first to second line we have used the fact that
667: 
668: \begin{equation}
669: \psi_{i,i,\dots,i}^{2N,l,p} \propto
670: \left(\frac{\alpha_i}{\alpha_p}\right)^{-\zeta_{2N,l}/\zeta_2} \ .
671: \end{equation}
672: This follows directly from Eqs. (\ref{gamshift}) and
673: (\ref{shift}).
674: 
675: Using the same argumentation as in Eq. (\ref{phi2}) we get:
676: \begin{eqnarray}\label{phimult}
677: \C P^{(2N)}_{p|n}(t)=\sum_l \frac{C_l}{(\alpha_nt)^\frac{\zeta_{2N,l}}{\zeta_2}}
678: \Lambda^{(2N,l)}(\alpha_pt) \, ,
679: \end{eqnarray}
680: 
681: here again the $C_l$'s are dimensional constants and the functions
682: $\Lambda^{(2N,l)}(x)$ are some functions with the asymptotic form
683: $\Lambda^{(2N,l)}(x)\approx x^{-\frac{\zeta_{2N,l}}{\zeta_2}}$ for $x$
684: large. We note that for
685: sufficiently long times and if the scaling exponents are well
686: separated, we can expect the sum to be dominated by the leading
687: scaling exponent.
688: \section{Data Analysis}
689: The detailed analysis that was possible for the simple model of
690: Sect. III is not available for the generic model Eqs.(\ref{NS}),
691: (\ref{PS}). Our strategy is to assume that the scaling forms
692: derived in the last section are still valid for the generic
693: model, and to demonstrate this by replotting the data
694: accordingly. We will see that these predictions are born out by
695: the data.
696: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
697: \subsection{analysis of the 2-point porpagator}
698: In light of  Eq. (\ref{prp2}), we expect the 2nd order propagator
699: $\hat \C P^{(2)}(k,k_0,t)$ to be, for a fixed $k_0$, a
700: homogeneous function of the variable $kt^{1/\zeta_2}$, and to
701: decay as $1/t$:
702: \begin{equation}
703: \hat \C
704: P^{(2)}(k,k_0,t)\propto\frac{1}{t}H^{(2)}\Big(kt^{1/\zeta_2}\Big)
705: \ ,
706: \end{equation}
707: where $H^{(2)}$ is some function. We test the correctness of this
708: form in Fig. \ref{fig4-1}. To this aim we re-plot that data shown
709: in Fig. \ref{fig1} in different coordinates, multiplied by $t$ and
710: as a function of
711: \begin{equation} \hat k=kt^{1/\zeta_2}\ . \end{equation}
712: The quality of the data collapse appears to
713: strongly support the proposed scaling form. We note that the data
714: collapse is superior on the right of the maximum, and less
715: convincing at its left. We believe that this stems from two
716: reasons. First, there is better statistics for the right part of
717: the curve, simply because it belongs to larger $k$ vectors where
718: the angular average is more extensive. Second, the left part of
719: the curve is more sensitive to the finite size effects,
720: particularly to the fact that the driving velocity field loses its
721: scaling form close to $L$. With all the limitations of the
722: numerical simulations we consider the data collapse as very
723: satisfactory.
724: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
725: \begin{figure}
726: \centering
727: \psfrag{yl}{$t \hat\C P^{(2)}(k,300,t)$}
728: \psfrag{xl}{$k t^{1/\zeta_2}$}
729: \includegraphics[width=.45\textwidth]{SPSfig6.eps}
730: \caption{$t\hat \C P^{(2)}(k,300,t)$ for 10 different times
731: (light grey earliest dark grey last)} \label{fig4-1}
732: \end{figure}
733: %%%%%%%%%%%%%%%%%%%%%%%%
734: 
735: If the prediction for the 2-point function holds, then we will have for the
736: time dependent integral:
737: \begin{equation}\label{t2}
738: \int_0^\infty \hat \C P^{(2)}(k,k',t)dk \approx
739: \frac{1}{t^{1+1/\zeta_2}}\int_0^\infty  H^{(2)}(\hat
740: k )d\hat k\propto\frac{1}{t^{1+1/\zeta_2}}
741: \end{equation}
742: In Fig. \ref{fig4-2}, We show that indeed after an initial period
743: the integral settles on a scaling form consistent with Eq.
744: (\ref{t2}).
745: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
746: \begin{figure}
747: \centering
748: \psfrag{yl}{$\int_0^\infty \hat \C P^{(2)}(k,300,t)dk$}
749: \psfrag{xl}{$t$}
750: \includegraphics[width=.45\textwidth]{SPSfig7.eps}
751: \caption{The integral of the 2-point propagator, $\int_0^\infty \hat \C
752:   P^{(2)}(k,300,t)d^2k$, (solid line) as a function of time. The dotted line
753:   is the expected $t^{-(1+1/\zeta_2)}$. (time is
754:   measured in this figure, and in the ones that follow in units of $\tau_{30}$ the eddy turnover
755:   time of $k=30$)} \label{fig4-2}
756: \end{figure}
757: %%%%%%%%%%%%%%%%%%%%%%%%
758: 
759: Using the form of the propagator we can establish the existence
760: of a left zero mode of the 2-point propagator. Integrating over
761: the two sides of Eq. (\ref{zeromode}) we have a time independent
762: expression on the left hand side. We therefore expect that the
763: weighed integral of the propagator with the function $Z^{(2)}$
764: will be constant:
765: \begin{eqnarray}
766: &&\int  Z^{(2)}(\B k)\C P^{(2)}(\B k,\B k',t) d^2 \B k\\
767: &&\propto\int_0^\infty  k^{-2-\zeta_2}\frac{H^{(2)}(k t^{1/\zeta_2})}{t}k dk=\int_0^\infty \frac{H(\hat{k})}{\hat k^{1+\zeta_2}}
768: d\hat{k}=\text{const} \ ,\nonumber
769: \end{eqnarray}
770: where we have used the fact that in an isotropic 2-dimensional
771: systen we have $Z^{(2)}(\B k)\sim F^{(2)}(\B k) \propto
772: k^{-2-\zeta_2}$. We note that the constancy of this integral
773: should be judged on the background of the decaying function, as
774: done in  Fig. \ref{fig4-3}. We see that while the 2nd order
775: structure function decays over three orders of magnitude, the
776: ``constant" objects changes by a factor of 2. The lack of
777: constancy can be attributed to the sensitivity to the outer scale
778: as seen in the data collapse in Fig. (\ref{fig4-1}). If the
779: collapse on the left side of the curve were perfect, so would be
780: the constancy of the weighted integral. Note that in the
781: calculation we have employed $\zeta_2 = 0.67$ in agreement with
782: \cite{00CLMV}
783: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
784: \begin{figure}
785: \centering
786: \psfrag{yl}{$\int Z^{(2)}(k)\C P^{(2)}(k,300,t) d^2 \B k$}
787: \psfrag{xl}{$t$}
788: \includegraphics[width=.45\textwidth]{SPSfig8.eps}
789: \caption{The integral of the 2-point propagator $P^{(2)}(k,300,t)$ weighed by the left zero mode
790: $Z^{(2)}(k)$ (solid line),
791:   compared to the integral of the un-weighed second order structure function in dotted line.}
792: \label{fig4-3}
793: 
794: \end{figure}
795: %%%%%%%%%%%%%%%%%%%%%%%%
796: \subsection{analysis of the multi-point propagators}
797: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
798: \begin{figure}
799: \centering
800: \psfrag{yl}{$t^{\frac{\zeta_4}{\zeta_2}} \hat\C P^{(4)}(k,300,t)$}
801: \psfrag{xl}{$k t^{1/\zeta_2}$}
802: \includegraphics[width=.45\textwidth]{SPSfig9.eps}
803: \psfrag{yl}{$t^{\frac{\zeta_6}{\zeta_2}} \hat\C P^{(6)}(k,300,t)$}
804: \psfrag{xl}{$k t^{1/\zeta_2}$}
805: \includegraphics[width=.45\textwidth]{SPSfig10.eps}
806: \caption{$t^{\frac{\zeta_4}{\zeta_2}}\hat \C P^{(4)}(k,300,t)$ (upper panel) and
807: $t^{\frac{\zeta_6}{\zeta_2}}\hat \C
808:   P^{(6)}(k,300,t)$ (lower panel) for 10 different times (light grey earliest dark grey last)}
809: \label{fig4-4}
810: \end{figure}
811: %%%%%%%%%%%%%%%%%%%%%%%%
812: Examining Eq.(\ref{phimult}), we expect that for long times the
813: functional form of the propagator for the fused correlation
814: function defined in Eq.(\ref{defpropdmult}) is:
815: \begin{equation}
816: \hat \C P^{(2N)}(k,k',t)\propto\frac{1}{t^{\zeta_{2N}/\zeta_2}}
817: H^{(2N)}(kt^{1/\zeta_2}) \ ,
818: \end{equation}
819: where $\zeta_{2N}$ is the leading scaling exponent for the $2N$'th
820: correlation function. In Fig. \ref{fig4-4} we demonstrate the
821: data collapse obtained by assuming this form  for the 4'th and
822: 6'th order propagators. We notice the same excellent collapse on the right hand side of
823: the function as in the 2nd order propagator. Also the problems
824: with the outer scale show up in a similar manner, giving less
825: than impressive collapse of the left hand part of the function.
826: For the present data collapse we employed simple scaling
827: $\zeta_{2N} = N\zeta_2$; our data does not support strongly
828: anomalous exponents.
829: 
830: Using the form of the fused multi-point correlators we can again
831: predict the time behavior of their integral:
832: \begin{eqnarray}\label{tmult}
833: \int_0^\infty \hat \C P^{(2N)}(k,k',t)dk &\approx&
834: \frac{1}{t^{(1+\zeta_{2N})/\zeta_2}}\int_0^\infty  H(\hat
835: k)d\hat k \nonumber\\
836: &\propto&\frac{1}{t^{(1+\zeta_{2N})/\zeta_2}}\ .
837: \end{eqnarray}
838: In Fig. \ref{fig4-5} the time dependence of the integrals is
839: plotted, showing agreement of similar quality to Fig.
840: \ref{fig4-2}.
841: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
842: \begin{figure}
843: \centering
844: \psfrag{yl}{$\int_0^\infty \hat \C P^{(4)}(k,300,t)dk$}
845: \psfrag{xl}{$t$}
846: \includegraphics[width=.45\textwidth]{SPSfig11.eps}
847: \psfrag{yl}{$\int_0^\infty \hat \C P^{(6)}(k,300,t)dk$}
848: \psfrag{xl}{$t$}
849: \includegraphics[width=.45\textwidth]{SPSfig12.eps}
850: \caption{The integral of the 4-point (upper panel) and 6-point (lower
851:   panel) propagators in solid line, $\int_0^\infty \hat \C
852:   P^{(2m)}(k,300,t)d^2k$, compared with the expected
853:   $t^{-(1+\zeta_{2N})/\zeta_2}$, in the dotted line. } \label{fig4-5}
854: \end{figure}
855: %%%%%%%%%%%%%%%%%%%%%%%%
856: As in the 2-point case, weighing the fused multi-point functions
857: by the appropriate multi point fused steady state correlators
858: should yield a constant:
859: \begin{eqnarray}
860: &&\int  Z^{(2N)}(k)\hat{\C P}^{(2N)}( k,k',t) d^2 \B k
861: \\&&\propto\int_0^\infty  \frac{H^{(2N)}(kt^{1/\zeta_2})}
862: {k^{2+\zeta_{2N}} t^{\zeta_{2N}/\zeta_2}}k dk
863: =\int_0^\infty \frac{H^{(2N)}(\hat{k})}{\hat k^{1+\zeta_{2N}}}
864: d\hat{k}=\text{const} \ ,\nonumber
865: \end{eqnarray}
866: where $Z^{2N}(k)\propto k^{-2-\zeta_{2N}}$. The quality of the
867: constancy is demonstrated in Fig. \ref{fig4-6}. Again finite size
868: effects lead to some decrease in time of these weighted objects,
869: which nevertheless is very much reduced compared to the decaying
870: correlation functions.
871: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
872: \begin{figure}
873: \centering
874: \psfrag{yl}{$\int Z^{(4)}( k)\hat \C P^{(4)}( k,300,t) d^2 \B k$}
875: \psfrag{xl}{$t$}
876: \includegraphics[width=.45\textwidth]{SPSfig13.eps}
877: \psfrag{yl}{$\int Z^{(6)}( k)\hat \C P^{(6)}( k, 300,t) d^2 \B k$}
878: \psfrag{xl}{$t$}
879: \includegraphics[width=.45\textwidth]{SPSfig14.eps}
880: \caption{The integral of the 4-point propagator $P^{(4)}(k, 300,t)$ (upper
881:   panel) and the 6-point propagator $P^{(6)}(k,300,t)$ weighed by the left
882:   zero modes $Z^{(4)}( k)$ and   $Z^{(6)}(k)$ respectively in solid
883:   line, compared to the integral of the un-weighed object in dotted line}  \label{fig4-6}
884: \end{figure}
885: %%%%%%%%%%%%%%%%%%%%%%%%
886: \section{concluding remarks}
887: We have demonstrated in this paper that the generic advection of
888: a passive scalar by a velocity field that obeys the Navier-Stokes
889: equations can be discussed in terms of Eulerian Statistically
890: Preserved Structures. By initiating a decay with delta-function
891: initial conditions (concentrated on $k=300$) we have found
892: numerically the corresponding columns of the time-dependent
893: propagators for the 2nd, 4th and 6th order correlation functions
894: (where for the 4th and 6th order objects we considered partial
895: (``fused") information). Note that in contrast to the Lagrangian
896: formulation of Statistically Preserved Structures \cite{01CV},
897: for which there is no Preserved Structure corresponding to the
898: 2nd order correlation, in the Eulerian formulation such an object
899: exists and had been analyzed explicitly. We have used a simple
900: (non-generic) model of passive scalar advection to guess the
901: analytic scaling form of the propagators in the generic problem.
902: The test for the relevance of this guessed form is the data
903: collapse shown in Fig. 5 and 8. The guessed time dependence
904: appears to be in close correspondence with the data as shown in
905: Figs. 6 and 9.
906: 
907: The analytic forms of the propagators predict the existence of
908: eigen-modes with eigenvalue 1 (Statistically Preserved
909: Structures). We believe that this is the first demonstration of
910: Eulerian Statistically Preserved Structures in a generic flow. The
911: numerical evidence for the constancy of the latter is
912: encouraging, if not fully conclusive, as seen in Figs. 7 and 10.
913: We attributed the (relatively small) decrease in amplitude of the
914: putative Statistically Preserved Structures to the less than
915: perfect data collapse at the largest scales (smallest
916: $k$-vectors) that are seen in Figs. 5 and 8. These in turn stem
917: from the intervention of the outer scale in our scaling range, a
918: boundary effect that we did not succeed to eliminate in our
919: modest-size simulations. It would be interesting to see whether
920: larger 2-d simulations could remove this finite-size effect to
921: demonstrate conclusively the constancy of the Statistically
922: Preserved Structures.
923: 
924: In concluding we wish to point out an additional benefit to the
925: present formulation. Usually in modeling turbulent advection it
926: is customary to resort to dubious concepts such as `turbulent
927: diffusion' in order to write a diffusion equation for the
928: correlation functions. The present approach indicates a much
929: better procedure, i.e. to find, for a given turbulent field, the
930: form of the propagator which can be then used to provide analytic
931: predictions for any initial (t=0) correlation function. This
932: procedure may be quite attractive for more complex hydrodynamic
933: flows with unusual boundaries or coherent structures. We believe
934: that when the turbulence is sufficiently well developed scaling
935: forms for the propagator will exist, and once found can be used
936: to solve efficiently any statistical initial value problem.
937: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
938: \begin{acknowledgments}
939: We thank Antonio Celani for providing us with his code for DNS of
940: 2-dimensional turbulent advection. This work had been supported
941: in part by the European Commission under a TMR grant, the
942: Minerva Foundation, Munich, Germany, and the Naftali and Anna
943: Backenroth-Bronicki Fund for Research in Chaos and Complexity.
944: 
945: \end{acknowledgments}
946: \begin{thebibliography}{99}
947: \bibitem{79MY}
948: A.S. Monin and A.M. Yaglom, {\em Statistical Fluid Mechanics}, vol. 1, chapter 5,
949: (MIT, Cambridge 1979).
950: 
951: \bibitem{98BGK}
952: D. Bernard, K. Gawedzki, A. Kupiainen, J.Stat. Phys {\bf 90}, 519 (1998)
953: 
954: \bibitem{01CV}
955: A. Celani and M. Vergassola, Phys. Rev. Lett. {\bf 86}, 424
956: (2001).
957: 
958: \bibitem{01ABCPV}
959: I. Arad, L. Biferale, A. Celani, I. Procaccia and M. Vergassola,
960: Phys. Rev. Lett. {\bf 87}, 164502 (2001).
961: 
962: \bibitem{02CGP}
963: Y. Cohen, T. Gilbert and I. Procaccia, Phys. Rev. E, {\bf 65},
964: 026314 (2002).
965: 
966: \bibitem{97BBW}
967: R. Benzi, L. Biferale and A. Wirth, Phys. Rev. Lett. {\bf 78}, 4926 (1997)
968: 
969: \bibitem{00CLMV}
970: A. Celani,  A. Lanotte, A. Mazzino and M. Vergassola, Phys. Rev. Lett. {\bf 84}, 2385
971: (2000).
972: 
973: \end{thebibliography}
974: 
975: \end{document}
976: