• Aucun résultat trouvé

Computation of the dimension of a model of fully developed turbulence

N/A
N/A
Protected

Academic year: 2021

Partager "Computation of the dimension of a model of fully developed turbulence"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: jpa-00210300

https://hal.archives-ouvertes.fr/jpa-00210300

Submitted on 1 Jan 1986

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Computation of the dimension of a model of fully developed turbulence

R. Grappin, J. Léorat, Annick Pouquet

To cite this version:

R. Grappin, J. Léorat, Annick Pouquet. Computation of the dimension of a model of fully developed turbulence. Journal de Physique, 1986, 47 (7), pp.1127-1136. �10.1051/jphys:019860047070112700�.

�jpa-00210300�

(2)

1127

Computation of the dimension of a model of fully developed turbulence

R. Grappin (1), J. Léorat (1) and A. Pouquet (2)

(1) CNRS Observatoire de Meudon, DAF, F-92190 Meudon, France (2) CNRS Observatoire de Nice, BP252, F-06007 Nice Cedex, France

(Reçu le 27 décembre 1985, accepté le 24

mars

1986)

Résumé. 2014 On calcule la dimension de l’attracteur d’un modèle scalaire de turbulence MHD développée selon la

méthode des exposants de Lyapounov d’une part, et d’autre part à l’aide d’un algorithme de comptage. On obtient numériquement que la dimension de Lyapounov est à peu près égale

au

nombre total de modes présents dans la

zone

inertielle, et que l’exposant maximal, c’est-à-dire le taux de divergence des trajectoires proches, est donné

par l’inverse du temps caractéristique le plus court présent dans la zone inertielle. La dimension de corrélation,

d’autre part, est égale à environ un tiers du nombre de modes présents dans la zone inertielle. Les deux estimations de la dimension

ne

coincident qu’à faible nombre de Reynolds, c’est-à-dire à basse dimension (égale à environ 3).

Extrapolés à la turbulence réelle à 3 dimensions (Navier-Stockes), nos résultats numériques sont compatibles

avec

les deux lois d’échelle : 1) l’exposant de Lyapounov maximal varie

comme

Re1/2, 2) la dimension de Lya-

pounov et l’entropie de Kolmogorov varient comme Re9/4 (pas de saturation de la dimension de l’attracteur

turbulent).

Abstract

2014

We compute numerically the attractor dimension of

a

model of fully developed MHD turbulence both

using Lyapounov exponents, and via

a

counting algorithm. The Lyapounov dimension is found to be essentially given by the total number of modes which lie in the inertial range; the maximal exponent, i.e. the divergence rate

of nearby trajectories, is given by the inverse of the shortest time scale available in the inertial range. The correlation

dimension,

on

the other hand, is found to be given by roughly

one

third of the modes in the inertial range. Both evaluations of dimension coincide only at very low Reynolds, i.e. at low dimension (equal to about 3). If extra- polated to real 3-dimensional Navier-Stokes turbulence,

our

results

are

consistent with the two scaling laws : 1) the maximal Lyapounov exponent scales

as

Re1/2, 2) the Lyapounov dimension and Kolmogorov entropy scale

as

Re9/4 (no saturation of the dimension of the turbulent attractor).

J. Physique 47 (1986) 1127-1136 JUILLET 1986,

Classification Physics Abstracts

03.20 - 47.25

1. Introduction

The description of homogeneous turbulent flows at

high Reynolds number requires a great number of variables, which may exceed the capacity of even the biggest computers : there is no hope in the near

future to simulate the smallest scales of most geophy-

sical or astrophysical flows. Turbulence modelling

is a way to circumvent this difficulty, by reducing

the number of degrees of freedom taken into account in the calculations. The minimum number of inde-

pendent variables which is necessary to represent a turbulent flow depends on the Reynolds number,

but also on the method used to model turbulence.

For instance, direct numerical integration of the primitive equations at a given Reynolds number requires more variables than the integration of spectral equations obtained by an isotropic closure model, with geometric discretisation of wavenumbers.

Can we justify any eventual reduction of degrees

of freedom in homogeneous turbulence modelling ?

One way would be to show that the dimension of the turbulent attractor is smaller than the total dimension of phase space. Laminar flows, which alone are

accessible to detailed analysis, have a small number of effective degrees of freedom; new excited modes appear during the transition to turbulence, the precise scenario depending on the flow (B6nard, Couette, etc...). Review of work in that line is given

in Abraham et al. [1].

Since Lorenz, several models have allowed to better understand the first bifurcations from laminar to turbulent (o chaotic ») states, but their extension to

fully developed turbulence seems difficult (see for

example Francheschini and Tebaldi [2]). On another

side, using a rigorous analytical study of the Navier- Stokes equations, Constantin et al. [3a, 3b] have recently found an upper bound for the number of

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:019860047070112700

(3)

pertinent modes. We propose here to compute nume-

rically the dimension of a model of fully developed

turbulence.

The model we shall study is a generalization to

MHD of a hydrodynamical model introduced by Desniansky and Novikov [4, 5] : in its simplest for- mulation, it involves N

=

2 S real variables Xn (S kinetic modes and S magnetic modes). Spatial

scales are discretized by octaves and appear in the

equations via the wavenumbers k,, - 2n -1. The name

of the model stems from the fact that all physical quantities are scaled, using the association of the n-th mode with wavenumber kn. The equations read formally (see Gloaguen et al. [6] for details) :

In equation (1) the constant coefficients Aij couple only neighbouring scales (i, j

=

n, n ± 1) and are

chosen so as to conserve the two inviscid invariants of MHD equations, total energy and velocity-magnetic

fied correlation. This constraint reduces the number of free coupling parameters to two coupling constants

a and 3 (actually only the ratio 0153lP matters). The

coefficient Q stands for a dissipation coefficient,

either the viscosity v (for kinetic modes), or the magnetic diffusiviiy 11 (for magnetic modes). The

third term is a forcing (constant acceleration) which only acts on the first (large scale) kinetic mode.

The complete equations are given in the Appendix.

Let us summarize briefly the properties of the

model (see Gloaguen et al. [6] for details). If one neglects the dissipation and forcing terms, it is a

conservative system, i.e. it conserves an N-volume in phase space. Gibbs statistical equilibrium obtains,

with equipartition of energy between all modes.

If on the other hand we put non zero dissipation and forcing, then if N is high enough, and with suitable values of coupling constants a and fl (we shall come

back to this point in the next Section), two regions

appear in the wavenumber space. The large scales (o inertial » zone) show chaotic but persistent fluc- tuations ; long-time averages of energy distribution

reproduce very nearly the Kolmogorov power-law spectrum. Small scales (dissipative range) are on the

other hand characterized by excitation appearing

very intermittently, and an energy spectrum decaying

much more quickly (quasi-exponentially) with wave-

number. The boundary between both zones goes towards small scales when the viscosity v and magnetic diffusivity decrease.

Note that the coexistence of spatial scaling laws

and temporal chaos is an original property of this model; in particular, the hydrodynamical version

of this model (Desnyansky and Novikov [4, 5]) has

the Kolmogorov spectrum as a stable fixed point,

and shows thus no chaotic behaviour.

Since the scalar model verifies several characteristic basic properties of fully developed turbulence, it is

a good candidate to investigate how its dimension scales with the Reynolds number. There are several methods to measure the dimension of an attractor

numerically. The simplest is probably the «natural»

(or «pointwise») measure which amounts to count the number of points of the trajectory contained in a

disk of radius R around a given centre (average is eventually made later on the position of the centre :

see Grassberger and Procaccia [7]). If the number of

points grows as RD, then D is the dimension of the

attractor. A detailed account of results obtained via this method (of which a preliminary presentations has

been previously published in Pouquet et al. [8])

will be given in section 4.

Another method, which will be presented now, brings more information than a mere evaluation of dimension. It consists in computing the Lyapounov exponents (Benettin et al. [10]). (A first account of

some results obtained on the scalar model is given

in Grappin et al. [9].) These exponents generalize

to nonlinear dynamics the concept of eigenvalues of a

linear system with constant coefficients. Let the initial state X ° (which is a vector with N coordinates)

in equation (1) be replaced by X ° + 6X ’; linearizing

about X’, we have

where A (X’) is an N x N matrix. The trace of matrix A is a (negative) constant :

which gives the rate of contraction of a parallelepiped

with N orthogonal sides 6X’,..., 03B4Xtn. Lyapounov exponents Ai are time-averages of the eigenvalues of

matrix A, defined as :

where the 6 Yj are defined from the 6X! by an ortho- gonalization procedure (Benettin et al. [10]), 11 Y 11 being the norm of Y. An ordered spectrum of expo- nents is obtained in this way : ;’1 > Å2 > ... > AN,

the first exponent being positive (if the system is

chaotic) and measuring the average divergence of

two nearby trajectories, and at least the last exponent being negative, in order to satisfy equation (3). If we

start from a segment and build parallelepipeds with

a growing number of dimensions n, we see from

equation (3) that there must be a critical number Nr

such that the parallelepiped with Nc dimensions is still expanding, i.e. Jl( N c) = ;’1 + ... + ÅN c >

=

0,

while the one with N c + 1 dimensions shrinks with

time, i.e. Ii(Nc + 1) = A 1 + ... + ÅNc+ 1 0. The Lya-

pounov dimension of the attractor is defined by

Kaplan and Yorke [11] as a linear interpolation

(4)

1129

between these two successive integers :

Figure 1 illustrates this definition; it shows, for two

different Reynolds numbers, the curve of cumulated exponents (M(n) as a function of n) obtained by inte- grating system (1) : the Lyapounov dimension is

given by the abcissa of the intersection of the curve

p(n) with the horizontal axis.

Note that relation (5) implies that the conservative

dynamical system which obtains when forcing and dissipation are absent (v = 11

=

0) has dimension

DL

=

N. At first sight, this is a paradoxical result,

since system (1) possesses two invariants in that case, hence it should have N - 2 degrees of freedom left,

not N. However, the paradox disappears if we recall

that the Lyapounov exponents explore the dimensions of the linearized system (2), which has no reason to

have any invariant. Thus, in the conservative case, different Lyapounov exponents could result from

calculating N nearby nonlinear trajectories from

system (1), without any linearization. Remark however that this problem disappears in the dissipative case,

since there are no invariants any more.

2. Identification of attractors at large Reynolds.

In order that solutions of system (1) exhibit both chaos and power-laws for spectra averaged with time,

several conditions must be met. First, as in simula-

tions, we want the truncation kS 1 to be smaller than the dissipation scale. This is obtained by imposing

a viscosity v and magnetic diffusivity 11 large enough,

so that at the smallest scale, the dissipation terms are

greater than the nonlinear terms : and

where Xs is the (kinetic or magnetic) excitation at the smallest available scale. Otherwise, excitation will try to flow into scales which are not present in the truncated N-system, energy will accumulate at the smallest scale and the system will resemble a conser-

vative system. Assuming Kolmogorov scaling Xn N kn 1/3, the above condition (6) leads to

Second, substantial magnetic excitation must be present, since the hydrodynamic version of system (1) (Desnyansky and Novikov, [4, 5]) is not turbulent.

Last, kinetic and magnetic excitation should not be

completely correlated, (i.e. we should not have Un =bn

for all n or un = - bn for all n), otherwise nonlinear terms in equation (1) will vanish : this property is shared with the primitive MHD equations. These last

two conditions are governed by the choice of two parameters a and f3 which weight two groups of nonlinear terms in the evolution equations (1) (see

Fig. 1.

-

Cumulated Lyapounov exponents. The ordinate

n

is the

sum

of the first n Lyapounov exponents p,(n) = L 03BBi,

which measures the growth rate

a

parallelepiped in phase

space,

as

a function of its dimension

n.

The Lyapounov

dimension is given by the intersection of the

curve

with the horizontal axis. Two

runs are

shown, with parameters

a =

10- 2, fl

=

l, viscosities v

=

l OT 4 and v =6.25 x 10- 6,

and respectively N

=

18 and 24 modes. Integration times

are

up to T

=

1 024 and T

=

512, respectively. Note the large plateau, which grows with Reynolds number, of quasi-

null Lyapounov exponents in both

runs.

Appendix). The properties of the model have been

investigated in detail at small N(N

=

4 to 6) by (Gloaguen et al. [6]); these authors have also looked at one special value of the ratio of coupling parameters

«(Xlf3

=

0.01) at N

=

18. Analysis of the bifurcation

properties of the model when (XI f3 varies and N is

large is difficult, so that numerical integration is

needed.

A predictor-corrector numerical scheme of order four was used to integrate the main trajectory (Eq. (1)),

while the linear system of N equations (2) was inte- grated by calculating a fourth-order Taylor expansion

of the propagator exp(A(Xt) H), where H is the time step used to integrate equation (1). (Since the matrix A is tridiagonal, the computation of the first powers of matrix A is performed rapidly). Time-step is chosen

to be small enough, so that equality (3) is satisfied with

a relative precision of 10- 5. The method was tested

on the Lorenz system. Calculations up to T =1024 with N

=

18 modes (ks/k1

=

256), v

=

10- 3, 11 = 1.2 x 10- 3, time step H

=

2 x 10- 3, took a little

less than 10 min of CPU time on a Cray 1 computer.

For most of the runs, initial conditions were as

follows. Magnetic excitation was small at all scales

(5)

(bn

=

10-4), and kinetic excitation was substantial

only at large scale (exponential spectrum, u,,

=

exp( - kn)).

Rapid exploration of system (1) with these values of parameters, varying the value of the ratio of cou-

pling constants a/P, was undertaken. Integration was stopped at time T

=

64, or in some cases at time T

=

128. Such an integration time proved to be large enough to reveal unambiguously the nature of

the attractor. We found only three different types depending on the value of ratio a/# (see Table I). For oc/# smaller than about 0.3, a high dimension for the attractor is the rule, with a large inertial zone, large fluctuations, and equipartition of kinetic and magnetic

energy. For a/fl between 0.5 and 1, the attractor is a non-magnetic, stable fixed point (non-turbulent, cons-

tant Kolmogorov kinetic spectrum). For all larger

than 2, the correlation coefficient p

=

2 E v. bn/

E(V2 + bn) is seen to grow rapidly at almost maximal

values, with dimension decreasing accordingly with

time : the energy spectra are decreasing much more rapidly than a power-law, and less and less small scales are excited with time. We interpret these results

as indicating that the attractor is a fixed point with

maximal correlation (p

=

1 or p

= -

1), and thus

its Lyapounov dimension is zero. Note that, since

nonlinear terms vanish when I p I

=

1, time scales become infinite as [ p approaches its maximal value : hence the slow convergence in the above data.

It thus seems (see Table I) that the picture of bifur-

cations (in a/fl space) is much simpler at large N than

at small N, showing one stable fixed point at large all (magnetic at sufficiently large a/fl, non-magnetic at

moderate a/fl) and, at sufficiently low ratio alP, one high dimensional turbulent attractor. No attempt has been made to have a complete view of the bifurcation

landscape which has been skeched above (in particular,

we have kept fixed initial conditions).

We shall now concentrate on the magnetic turbu-

lent attractor properties, and study in detail the case

{3

=

1 and a

=

0.01, a case already considered in

(Gloaguen et al. [6]). Note that the a and fl coupling parameters represent respectively two groups of

interacting wavenumber triads (k, p, q) (with k

=

p + q) : the a parameter corresponds to (almost)

«flat triangles (i.e. quasi-parallel wavenumbers), whereas fl represents triangles such that k ;= p = q.

Thus the choice of a/fl

=

0.01 is consistent with

incompressible turbulence, for which strictly flat triangles do not contribute to nonlinear terms.

3. Spectral properties and Lyapounov exponents of the

turbulent attractor.

To obtain reasonably converging values of Lyapounov

exponents, we integrated system (1) at long times (up to T

=

1024),.with a

=

10-2, #

=

1 (see Fig. 2).

Table I.

-

Short time exploration of correlation coefficients p and Lyapounov dimension DL with various values of the coupling parameters ratio alP. The correlation coefficients p is 2 Ev,, b,,l E(V2 + b 2). Its initial value is 0.001 in all runs. The’Lyapounov dimension DL is given by equation (5). a and fl are two coupling parameters in system (1) (see Appendix). Number of modes is N

=

18, viscosity is v

=

10-3, magnetic diff usivity is ~

=

1.2 x 10- 3. The

---

correspond to non-calculated data. Note the three regions. First, almost maximal correlation obtains for alp>

=

2,

and dimension decreasing with time (with probable 0 limit, see text). Second, the stable fixed point for a/fl between 0.5

and 1 (there is vanishing magnetic energy in this case). Third, large values of dimension, which grows slowly with

time, for small values of a/fl (see text). Case a/03B2

=

0.01 is studied extensively in section 3.

(6)

1131

Fig. 2.

-

Convergence of Lyapounov dimension with time.

Note the growth of dimension with time, corresponding

to more and

more

modes becoming turbulent. Turn-over time (characteristic time of largest scale) is of order 1.

The large time scale in system (1) (which would be the

«turn-over time » of the larger eddies in a classical

turbulence description) is given by 1 /(k 1 u 1 )

=

I /u 1,

which was close to unity in all runs. We thus integrated

about 1 000 turn-over times, which seems necessary to completely explore the attractor. Note however that the time-step is not determined by the large scale

characteristic time, but by the small scale time, which depends on the viscosity, and may be much smaller.

Four values of viscosities, ranging from 6.25 x 10- 6

up to 10- 2, have been considered. In all runs, magnetic diffusivity and viscosity were taken with comparable

values (see Table II). The total number of modes

was 18 in all but one case, where it was 24.

Figure 1 gives the cumulated Lyapounov exponents distribution u(n) at time T

=

1 024, for the two higher Reynolds numbers. Note the large plateau, due to a

majority of exponents being very close to zero, which leads to a high Lyapounov dimension (Eq. (5)).

Figure 2 gives the Lyapounov dimension obtained

as a function of time, for different values of dissipation

coefficient. Note the slow rise of dimension with time, showing that a growing number of degrees of freedom

come into play : it may be a consequence of intermit- tency in the smallest scales.

Time-averaged spectra show large inertial ranges with Kolmogorov-like power-laws (Fig. 3), both for

kinetic and magnetic energy. A strong non-linear dynamo is at work : one observes equipartition

between both field at all but the largest scales, where magnetic energy remains almost absent. Note that initial magnetic energy is rather small ( 10- 4 at all modes) and that we have only kinetic forcing, (see Eq. (1)).

Inspection of figure 3 shows that decreasing the viscosity lengthens, as expected, the inertial range.

Let us recall the standard Kolmogorov phenome- nological prediction. The largest « inertial » wave-

number k* is determined by approximate balance

Fig. 3.

-

Time-average of kinetic and magnetic spectra.

Abcissa : n

=

log2(kn) ; ordinate : kinetic energy un /2 and magnetic energy b;/2 (right). Straight line : Kolmogorov

law k; 2/3. Note equipartition between kinetic and magnetic spectra, at all but the largest magnetic scales, and the slight departure from Kolmogorov law. All

runs are

with 18 modes, except with the smallest viscosity.

Table II.

-

Maximal Lyapounov exponents and Lyapounov dimension at long times, for case a

=

10- 2, 03B2

=

1.

The number of modes in the inertial range is found by visual inspection of figure 3 : one counts the modes which

follow a power-law( forinstance, in the run with viscosity 6.25 x 10-6, all modes are kept, except the two large scale

magnetic modes).

(7)

between linear dissipation terms and nonlinear terms in equation (1) (cf. [6]). This leads to :

and, as a corollary, we see that the shortest time scale in the inertial range scales as T * -(I lvk *2) - v 1/2.

A corresponding scaling is observed for both the maximum Lyapounov exponent and the Kolomo- gorov entropy 8 (sum over positive Lyapounov exponents) : 03BB1- E v-ll2 (see Fig. 4) suggesting

that the divergence of nearby states is governed by the

time scale of the smallest scale available in inertial range. Note however that the divergence rate at large scale, (which is related to the predictability problem

and is outside the scope of this paper), needs not be

related in any simple way to this time scale.

The fact that smaller scales do not contribute to the

divergence of trajectories may be viewed as due to the dominance of viscous damping at such scales, for

most of the time : these scales are only intermittently excited, when nearby (larger) scales have their maxi-

mum amplitude, which gives rise to transient nonlinear bursts. Figure 5 shows the spectrum of the flatness factor Fn = XI >/ X; )2 for kinetic and magnetic

modes. We make two remarks. First, the flatness

« spectrum >> is flat (about 10, a value substantially higher than the Gaussian value 3) in a large range of wavenumbers. Second, this flat range corresponds reasonably well to the inertial range, i.e. to the range where the corresponding energy spectrum follows a power-law. At all« non-inertial » scales, which are the

Fig. 4.

-

Maximum Lyapounov exponent and Kolmo- gorov entropy

versus

viscosity. Straight line is the curve

1 /. (see text).

Fig. 5.

-

Kinetic and magnetic flatness factor. Abcissa :

n =

log2(kn); ordinate : Fn = Xn ) j( XI >2, where Xn

is the kinetic (left)

or

magnetic (right) fluctuation. Note that the flat part of the spectrum corresponds to the inertial ranges of figure 3.

dissipation scales and the large magnetic scales, the intermittency is significantly higher.

Developed turbulence may thus be viewed as being

made of pertinent modes (those in the inertial range),

and non-pertinent, intermittent modes. More precisely,

we mean that the detailed number of non-inertial modes represented in the system does not change the

dimension : only the number of «inertial modes matters. To show this, we plot the Lyapounov dimen-

sion DL as a function of viscosity; results agree well with the curve :

(See Fig. 6). This may be interpreted as follows. The number of kinetic (or magnetic) inertial modes is

approximately given (see Eq. (7)) by :

(recall kn ’" 2n-l); the total number N* of inertial modes is (forgetting the one or two large scale inter- mittent magnetic modes) twice that number :

Hence equation (8) indicates that

which is what we wanted to show (see Table II)

4. Dimension obtained by counting algorithms.

A direct evaluation of dimension is provided by simply

sampling a set of N points of the trajectory and deter-

mining, as has been proposed by Grassberger and

(8)

1133

Fig. 6.

-

Lyapounov and correlation dimensions

versus

viscosity. Abcissa : n

=

log2(kn). Straight line : - 3/2 log2(viscosity) (see text). The good parallelism between this line and the Lyapounov dimension DL shows that DL grows

as

the number of modes following

a

Kolmogorov power- law (see text). Note that Lyapounov and correlation dimen- sions coincide only at low dimension (about 3).

Procaccia [7], the correlation function :

where H(x) is 0 when x 0 and 1 when x >

=

0.

Assuming C(r) - rD, D is the dimension (henceforth

called correlation dimension Dc) of the attractor (see also Mandelbrot [12]). The point-wide dimension

is obtained in the same way, but keeping one of the points (say X

=

Xo) fixed (with normalization factor

equal to I/N). The point-wise dimension looks at how the number of points in balls of increasing radii r

centred on X o augments with r. It naturally requires N

times less operations than (11) in principle (see below)

but is a local defmition and its value may vary from

point to point.

Numerically, the algorithm requires the storage of the relative distance in phase space between the N

points; this array is of order N 2 and brings an im-

mediate limitation to the algorithm. In pratice, one

stores the coordinates of the points on the attractor,

with a sampling time which is of the order (but smaller than) the large-scale eddy turn-over time.

The code was checked on the Lorenz system : the dynamical dimension calculated with 104 points was

found to agree with the Lyapounov dimension, for the standard values of parameters. We also checked that,

for parameter values in the scalar model, such that

either a fixed point or a limit cycle obtains, the corre-

lation dimension found numerically is as expected (0 and 1).

An important feature of these counting algorithms

is that it is very easy to obtain, on a log-log plot, a straight line, even when convergence is not established

Figure 7a shows the correlation functions (11) obtained

on the 6-mode system with cx = 0.01, # = 1, v = tj = 0.1769, varying the integration time (T

=

1000, 2 000, 20 000) with fixed sampling time interval (At

=

0.2).

This low-dimensional system has for these values of parameters a stable fixed point (see Gloaguen [13]).

However, with a given value of the time-step, the

numerical solution shows persistent oscillations (whose amplitude diminish when the time-step is reduced)

around the fixed point, hence the non-zero slopes of

the correlation functions observed on the figure at

small scales.

The resulting dimension appears to converge from below with the integration time (or number of sam- pling points). Figure 7b shows the case of the 8-modes system with v = q

=

0.0102 and respectively T

=

3 000, 4 000, 5 000. The resulting dimensions are

summarized in table III. These data show that a small

sample strongly underestimates the actual dimension of the attractor but one needs not double the number of points, since a small increase suffices to indicate the rate of convergence of the algorithm.

We now come to the evaluation of dimension at

high Reynolds numbers, with a

=

0.01, P

=

1 as in

the calculations of the preceeding section. As the number of modes is increased, the Reynolds number

is increased, by decreasing viscosity and magnetic diffusivity according to inequality (6). Table IV and figure 6 show the variation of both Lyapounov and

correlation dimensions with viscosity, with for corre-

lation dimension a number of modes between 6 and 18, and a viscosity between 2.5 x 10-2 and 2.5 x 10-4.

Table III.

-

Dimension calculated by counting algo-

rithms : variation with number of sampling points for

low-dimensional systems. All dimensions have been obtained via the correlation function (11), except the

fourth result of run A, which is a point-wise dimension (see text) with averaging on 2fl points and calculating

on 105 sampling points. Run A : 6-modes system;

run B : 8-modes system (see text). Note that reasonable

convergence obtains at about 30 000 points.

(9)

Fig. 7.

-

Correlation functions versus number of sampling points. Abcissa : log, (distance); ordinate : log. (correlation function). The correlation function is defined in equation (11); it is obtained by counting

a

subset of points

on

the computed trajectory. Its slope gives the

«

correlation dimension » (see text).

a) 6-mode system. A best fit on small scales give the following dimensions : number of points : 5 000 10 000 100 000

correlation dimension : 1.68 2.22 2.56

b) 8-mode system. A best fit gives :

number of points : 15 000 20 000 25 000

correlation dimension : 2.86 3.93 4.06

Note in all cases the straight lines in log-log plot,

even

at the lowest resolutions.

Table IV.

-

Dimension calculated by counting algorithm : variation with Reynolds number. All dimensions have been obtained via the correlation function (11). Viscosity and diffusivity are equal. Reynolds is about I /viscosity.

The correlation dimension varies from 3.3 to 8.3, but it must be stressed that, as the number of modes is increased, the time needed to achieve reasonable convergence on the computer becomes large (of the

order of one hour on a Cray 1 A for the last case).

5. Discussion

We have studied in this paper a model of fully deve- loped MHD turbulence which shares many structural

properties with real fluid equations. This model

exhibits «intrinsic stochasticity », i.e. does not rely

on external forcing to reach a turbulent regime. Note

that the constant acceleration on large kinetic scale

in system (1) was imposed only in order to obtain

statistical stationarity.

Computation of Lyapounov dimension has also

’ been done recently for another model of developed turbulence, governed by the Kuramoto-Sivashinsky equation (see Manneville [14]). However, the results

so obtained cannot be easily extrapolated to fluid turbulence, since there is no other common properties

between this model and real fluid than the Galilean invariance and the possibility to control the number of

«turbulent degrees of freedom by fixing an external parameter.

On the other hand, direct study of the Navier- Stokes or MHD equations are time-consuming, and

it seems to be impractical to calculate more than the

first Lyapounov exponents on the primitive equations.

Legras et al. (1985, private communication) have

recently calculated the first two Lyapounov exponents

(10)

1135

of 2-dimensional Navier-Stokes flow. Their results

seem to be compatible with the relation that is shown in figure 4, namely Ål being comparable to the inverse of the shortest time-scale in the inertial range.

Let us summarize our results. A rapid exploration, varying the coupling parameters indicated that for a large number of degrees of freedom (N

=

18, corres- ponding to k max/k min

=

256), system (1) has three

attractors : one hydrodynamical (non-magnetic) fixed point, a maximal velocity-magnetic correlation attrac- tor which is probably a fixed point, and a turbulent

attractor.

A detailed study of this turbulent attractor, varying

the viscosity and using both the Lyapounov and

correlation algorithms, has shown that its dimension is growing with Reynolds number with apparently no saturation, thus confirming the indication given in Pouquet et al. [8]. However, the resulting dimension

estimations differ greatly at large Reynolds, depen- ding on which technique is employed For the Lya-

pounov dimension, all « inertial modes » are relevant, while for the correlation dimension, only a fraction (around one third) is relevant. Actually, extrapolating

the results plotted in figure 6 shows that both esti- mations probably coincide only at very low Reynolds,

for an attractor’s dimension equal to about 3. A

comparison between different methods of measuring

dimensions of dynamical systems has been given by

Termonia and Alexandrowicz [ 15] : these authors also find that the Lyapounov dimension exceeds the correlation dimension for dimensions larger than

about 3 (they find Dc - 7.5, while DL 10 in the

case of a delay differential Eq.). This discrepancy may be explained by arguments such as those developed

in Atten et al. [16], that the counting algorithms

underestimate the dimension more and more when the dimension grows, due to the simple geometric effect that, in a phase space with large number of dimen- sions, almost all couples of points accumulate at or near to the maximum distance available (a border effect).

The results obtained by computing Lyapounov exponents may be summarized as follows (see Figs. 4

and 6) :

(i) first Lyapounov exponent and Kolmogorov entropy scale as 10,

(ii) dimension DL is close to the number N * of modes in the inertial range,

(iii) dimension DL scales as - 3 j2 log2 ( v).

It has been shown in section 3 that points (ii) and (iii) were.consistent with standard Kolmogorov pheno- menology, which predicts for the logarithmic discre-

tization of modes used in the scalar model that N* = - 3/2 log2(v). (Note that Kolmogorov phe- nomenology «works » in the framework of our

model (which is a model of MHD turbulence), while

we would perhaps expect the MHD phenomenology

of Kraichnan [17] to work here. This is due to the lack,

in our model, of interactions between widely separated

modes, which suppresses the possibility of any Alfv6n effect (see Gloaguen et al. [6D).

These results are in line with those of Ruelle [18],

who found by studying the Navier-Stokes equation, using analytical and phenomenological arguments, that intermittent flows should have an infinite number of null characteristic (Lyapounov) exponents per unit volume (see the large plateau in Fig. 1). Equivalently,

we may say that all degrees of freedom contribute

to the dimension, except the intermittent modes

(which are essentially in the dissipation range at small scales).

Constantin et al. [2] have proposed, as an upper

bound, that the dimension of the Navier-Stokes attractor would scale as Re . It is easy to see that relation (ii) leads to the same result, when extrapolated

to 3-dimensional Navier-Stokes turbulence. Indeed,

the number N* of inertial modes is then (k*)3, or

v- 9/4 (see Eq. (7)).

We are thus tempted to generalize our results and to

conjecture that for turbulence with Kolmogorov spectra :

(1) the first Lyapounov, and Kolmogorov entropy, scale as Re1/2

(2) the Lyapounov dimension scales as Re9/4.

That the dimension of the scalar model is so large (indeed, the largest possible) raises some questions.

First, the model presents temporal intermittency,

both at large and small scales (see Fig. 5), and this does not produce any noticeable reduction of the dimension

(see the arguments of Kraichnan [19] in the context

of spatial intermittency, in favour of such a reduction).

Second, a dissipative turbulent system and a conser- vative system with the same number of excited modes may thus have almost the same dimension, although

their statistical properties are qualitatively different.

This limits the usefulness of the concept of dimension.

Finally, if the dimension of real Navier-Stokes tur- bulence is indeed maximal, as conjectured above, all

modes are pertinent to describe the turbulent attractor

(except the most « intermittent » modes which lie in the dissipative range plus the largest scales in the MHD case), and there would thus be no hope for any substantial reduction of the number of degrees of

freedom at large Reynolds.

We are of the opinion that the importance of

« pertinent >> modes, as measured by the Lyapounov dimension, must not be exaggerated Even if the turbulent attractor asks for an infinite number of

degrees of freedom, its spatio-temporal properties

may be correctly described with a reduced number of

degrees of freedom with proper (yet to be found) averaging. Solitons provide an example of such a

reduction in the context of integrable systems. One of the remaining most interesting questions on the scalar

model is the possibility (under study) to model struc-

tures like random bursts of excitation propagating

towards both ends of the spectrum, which could build

power spectra in the long term.

(11)

Acknowledgments.

We want to thank J. Curry, U. Frisch, C. Froeschl6,

C. Gloaguen and D. Russel for useful discussions at

early stages of this work. We are also grateful to

P. Manneville for pointing to us the work of Ruelle

[18] and Atten et al. [16], and for his constructive

criticism of an early version of this paper. Part of the

computational facilities were provided by the National Center for Atmospheric Research (Boulder), and part

by the Scientific Commitee of the Centre de Calcul Vectoriel pour la Recherche (CCVR), which are both

thanked

Appendix.

EQUATIONS OF THE SCALAR MODEL (SEE GLOAGUEN et al. [6]).

-

We give here the details of equation (1), which

describe the one-dimensional model (o scalar model ») of the MHD equations studied in this paper. The equa- tions for the kinetic modes read :

The equations for the magnetic modes read :

The coefficients v and q are respectively the viscosity and magnetic diffusivity. Note that a and P are free para- meters ; actually, only the ratio al P matters, since, given (XI/3, choosing a amounts to choose a time unit.

References

[1] ABRAHAM, N., GOLLUB, J. P., SWINNEY, H. L., Physica

11D (1984) 252.

[2] FRANCHESCHINI, V. and TEBALDI, C., J. Stat. Phys. 25 (1981) 397.

[3a] CONSTANTIN, P., FOIAS, C., MANLEY, O. P., TEMAM, R., J. Fluid Mechanics 150 (1985) 427.

[3b] CONSTANTIN, P., FOIAS, C., TEMAM, R., Memoirs of

AMS 53 (1985) 314.

[4] DESNIANSKI, V. N., NOVIKOV, E. A., Prikl. Mat. Mech.

38 (1974) 507.

[5] DESNIANSKI, V. N., NOVIKOV, E. A., Izvestya Atm.

Oceanic Phys., 10 (1974) 127.

[6] GLOAGUEN, C., LÉORAT, J., POUQUET, A., GRAPPIN, R., Physica 17D (1985) 154.

[7] GRASSBERGER, P. and PROCACCIA, I., Phys. Rev. Lett.

83 (1983) 346.

[8] POUQUET, A., GLOAGUEN, C., LÉORAT, J., GRAPPIN, R.,

A scalar model of MHD turbulence, in Chaos and Statistical Methods, ed. Y. Kuramoto (Springer, Berlin) 1984.

[9] GRAPPIN, R., POUQUET, A., LÉORAT, J., Computation

of a dimension of

a

model of developed turbulence,

in Macroscopic modeling of turbulent flows, ed.

U. Frisch, J. B. Keller, G. Papanicolaou, O.

Pironneau (Springer, Berlin) 1985.

[10] BENETTIN, G., GALGANI, L., GIORGILLI, A., STRELCYN,

J. M., Meccanica 15 (1980) 10 and 21.

[11] KAPLAN, J. L., YORKE, J. A., Lecture notes in Mathe- matics, 730, eds. Dietten and Walther (Springer)

1979, p. 228.

[12] MANDELBROT, B., J. Fluid Mech. 62 (1974) 381.

[13] GLOAGUEN, C., Un modèle scalaire de turbulence MHD,

Thèse de troisième cycle, Université Paris 7 (1983).

[14] MANNEVILLE, P., Lyapounov exponents for the Kura-

moto-Sivashinsky model, in Macroscopic model- ing of turbulent flows, ed. U. Frisch, J. B. Keller,

G. Papanicolaou, O. Pironneau (Springer, Berlin)

1985.

[15] TERMONIA, Y., ALEXANDROWICZ, Z., Phys. Rev. Letters 51 (1983) 1265.

[16] ATTEN, P., CAPUTO, J. G., MALRAISON, B., GAGNE, Y.,

Determination of attractor dimension of various flows (Laboratoire d’électrostatique, CNRS, 166X

38042 Grenoble, France), preprint (1984).

[17] KRAICHNAN, R. H., Phys. Fluids 8 (1965) 1385.

[18] RUELLE, D., Commun. Math. Phys. 87 (1982) 287.

[19] KRAICHNAN, R. H., Phys. Fluids 28 (1985) 10.

Références

Documents relatifs

In fact, the exceptional leaf just constructed has a fundamental group which is not finitely generated.. To see this, let yi be, as above, a generator of the fundamental group of

As it has been recently pointed out by Alexander Premet, Remark 3.12 in [M08] which claims that ”in the classical case, the dimension of a sheet containing a given nilpotent orbit

To get the lower bound, the strategy is to couple our process with a family of other stable-like processes whose Hausdorff dimension is known and then compare the Hausdorff dimension

In the next section, we only state that in case of trees, the Boolean dimension, geometric dimension and the minimum rank coincide, thus the formula given in Theorem 2.9 below for

Our aim in this article is to present the fractal nature of X(E) through its Hausdorff dimension, with the assumption that X belongs to the class of multistable Lévy processes,

For symmetric distributions satisfying some exponential moments condition, we proved that the sum S n of the random variables behaves as in the typical critical generalized

Then the closed ideals of L correspond exactly, via the dimension map, to polar subsets of the Dedekind-complete lattice- ordered group K 0 (L). This leads to methods which can

Actually, the Broad Economic Policy Guidelines (BEPG) had been synchronized, with the policy cycle of the European Employment Strategy (EES) already in 2003, in order