5
Exact Relation for Correlation Functions in Compressible Isothermal Turbulence Se ´bastien Galtier 1,2 and Supratik Banerjee 1 1 Universite ´ Paris-Sud, Institut d’Astrophysique Spatiale, UMR 8617, ba ˆtiment 121, F-91405 Orsay, France 2 Institut Universitaire de France, 103, boulevard Saint-Michel, 75005 Paris, France (Received 24 March 2011; published 23 September 2011) Compressible isothermal turbulence is analyzed under the assumption of homogeneity and in the asymptotic limit of a high Reynolds number. An exact relation is derived for some two-point correlation functions which reveals a fundamental difference with the incompressible case. The main difference resides in the presence of a new type of term which acts on the inertial range similarly as a source or a sink for the mean energy transfer rate. When isotropy is assumed, compressible turbulence may be described by the relation 2 3 " eff r ¼ F r ðrÞ, where F r is the radial component of the two-point correlation functions and " eff is an effective mean total energy injection rate. By dimensional arguments, we predict that a spectrum in k 5=3 may still be preserved at small scales if the density-weighted fluid velocity & 1=3 u is used. DOI: 10.1103/PhysRevLett.107.134501 PACS numbers: 47.27.eb, 47.27.Gs, 47.40.x Introduction.—Fully developed turbulence is often seen as the last great unsolved problem in classical physics which has evaded physical understanding for many deca- des. Although significant advances have been made in the regime of wave turbulence for which a systematic analysis is possible [1], the regime of strong turbulence—the sub- ject of this Letter—continues to resist modern efforts at a solution; for that reason, any exact result is of great im- portance. In his third 1941 turbulence paper, Kolmogorov derived an exact relation for incompressible isotropic hy- drodynamics in terms of third-order longitudinal structure function and in the asymptotic limit of a high Reynolds number (Re) [2]. Because of the rarity of such results, the Kolmogorov’s universal four-fifths law has a cornerstone role in the analysis of turbulence [3]. Few extensions of such results to other fluids have been made; it concerns, for example, a scalar passively advected such as the tempera- ture or a pollutant in the atmosphere, quasigeostrophic flows, or astrophysical magnetized fluids described in the framework of (Hall) MHD [4]. It is only recently that an attempt to generalize such laws to axisymmetric turbulence has been made, but an additional assumption is made about the foliation of the correlation space [5]. The previous results are found for incompressible fluids, and to our knowledge no universal law has been derived for compressible turbulence (except for the wave turbulence regime [6]), which is far more difficult to analyze. The lack of knowledge is such that even basic statements about turbulence like the presence of a cascade, an inertial range, and constant flux energy spectra are not well documented [7]. That is in contrast with the domain of application of compressible turbulence which ranges from aeronautical engineering to astrophysics [810]. In the latter case it is believed that highly compressible turbulence controls star formation in interstellar clouds [11], whereas in the former case Re is relatively smaller. In that context, the pressureless hydrodynamics is an interesting model to investigate the limit of high Mach number compressible turbulence whose simplest form is the one-dimension Burgers equation, which has been the subject of many investigations [12]. Among the large number of results, we may note that with exact field- theoretical methods it is possible to find explicit forms of some probability distributions [13]; it is also possible to derive the corresponding exact Kolmogorov law for the third-order structure function [3]. In the general case, our knowledge of compressible hy- drodynamic turbulence is mainly limited to direct numeri- cal simulations [14]. The most recent results for supersonic isothermal turbulence with a grid resolution up to 2048 3 [15] reveal that the inertial range velocity scaling deviates substantially from the incompressible Kolmogorov spec- trum with a slope of the velocity power spectrum close to 2 and an exponent of the third-order velocity structure function of about 1.3. Surprisingly, the incompressible pre- dictions are shown to be restored if the density-weighted fluid velocity & 1=3 u is used instead of simply the velocity u. Although a 2 spectrum may be associated with shocks— like in one dimension—it seems that their contribution in three dimensions (3D) is more subtle. Generally speaking, it is fundamental to establish the equivalent of the 4=5 law for compressible turbulence before going to the more difficult problem of intermittency [16]. In this Letter, compressible isothermal hydrodynamic turbulence is analyzed in the limit of high Re. We shall investigate the nature of such a compressible turbulence through an analysis in the physical space in terms of two- point correlation functions. In particular, the discussion is focused on the isotropic case for which a simple exact relation emerges. The theoretical predictions illuminate some recent high-resolution direct numerical simulations made in the astrophysical context. PRL 107, 134501 (2011) PHYSICAL REVIEW LETTERS week ending 23 SEPTEMBER 2011 0031-9007= 11=107(13)=134501(5) 134501-1 Ó 2011 American Physical Society

Exact Relation for Correlation Functions in Compressible Isothermal Turbulence

Embed Size (px)

Citation preview

Exact Relation for Correlation Functions in Compressible Isothermal Turbulence

Sebastien Galtier1,2 and Supratik Banerjee1

1Universite Paris-Sud, Institut d’Astrophysique Spatiale, UMR 8617, batiment 121, F-91405 Orsay, France2Institut Universitaire de France, 103, boulevard Saint-Michel, 75005 Paris, France

(Received 24 March 2011; published 23 September 2011)

Compressible isothermal turbulence is analyzed under the assumption of homogeneity and in the

asymptotic limit of a high Reynolds number. An exact relation is derived for some two-point correlation

functions which reveals a fundamental difference with the incompressible case. The main difference

resides in the presence of a new type of term which acts on the inertial range similarly as a source or a sink

for the mean energy transfer rate. When isotropy is assumed, compressible turbulence may be described

by the relation� 23"effr ¼ F rðrÞ, where F r is the radial component of the two-point correlation functions

and "eff is an effective mean total energy injection rate. By dimensional arguments, we predict that a

spectrum in k�5=3 may still be preserved at small scales if the density-weighted fluid velocity �1=3u is

used.

DOI: 10.1103/PhysRevLett.107.134501 PACS numbers: 47.27.eb, 47.27.Gs, 47.40.�x

Introduction.—Fully developed turbulence is often seenas the last great unsolved problem in classical physicswhich has evaded physical understanding for many deca-des. Although significant advances have been made in theregime of wave turbulence for which a systematic analysisis possible [1], the regime of strong turbulence—the sub-ject of this Letter—continues to resist modern efforts at asolution; for that reason, any exact result is of great im-portance. In his third 1941 turbulence paper, Kolmogorovderived an exact relation for incompressible isotropic hy-drodynamics in terms of third-order longitudinal structurefunction and in the asymptotic limit of a high Reynoldsnumber (Re) [2]. Because of the rarity of such results, theKolmogorov’s universal four-fifths law has a cornerstonerole in the analysis of turbulence [3]. Few extensions ofsuch results to other fluids have been made; it concerns, forexample, a scalar passively advected such as the tempera-ture or a pollutant in the atmosphere, quasigeostrophicflows, or astrophysical magnetized fluids described in theframework of (Hall) MHD [4]. It is only recently that anattempt to generalize such laws to axisymmetric turbulencehas been made, but an additional assumption is made aboutthe foliation of the correlation space [5].

The previous results are found for incompressible fluids,and to our knowledge no universal law has been derived forcompressible turbulence (except for the wave turbulenceregime [6]), which is far more difficult to analyze. The lackof knowledge is such that even basic statements aboutturbulence like the presence of a cascade, an inertial range,and constant flux energy spectra are not well documented[7]. That is in contrast with the domain of application ofcompressible turbulence which ranges from aeronauticalengineering to astrophysics [8–10]. In the latter case it isbelieved that highly compressible turbulence controls starformation in interstellar clouds [11], whereas in the formercase Re is relatively smaller.

In that context, the pressureless hydrodynamics is aninteresting model to investigate the limit of high Machnumber compressible turbulence whose simplest form isthe one-dimension Burgers equation, which has been thesubject of many investigations [12]. Among the largenumber of results, we may note that with exact field-theoretical methods it is possible to find explicit forms ofsome probability distributions [13]; it is also possible toderive the corresponding exact Kolmogorov law for thethird-order structure function [3].In the general case, our knowledge of compressible hy-

drodynamic turbulence is mainly limited to direct numeri-cal simulations [14]. The most recent results for supersonicisothermal turbulence with a grid resolution up to 20483

[15] reveal that the inertial range velocity scaling deviatessubstantially from the incompressible Kolmogorov spec-trum with a slope of the velocity power spectrum close to�2 and an exponent of the third-order velocity structurefunction of about 1.3. Surprisingly, the incompressible pre-dictions are shown to be restored if the density-weighted

fluid velocity�1=3u is used instead of simply the velocityu.Although a�2 spectrum may be associated with shocks—like in one dimension—it seems that their contribution inthree dimensions (3D) ismore subtle. Generally speaking, itis fundamental to establish the equivalent of the 4=5 law forcompressible turbulence before going to the more difficultproblem of intermittency [16].In this Letter, compressible isothermal hydrodynamic

turbulence is analyzed in the limit of high Re. We shallinvestigate the nature of such a compressible turbulencethrough an analysis in the physical space in terms of two-point correlation functions. In particular, the discussion isfocused on the isotropic case for which a simple exactrelation emerges. The theoretical predictions illuminatesome recent high-resolution direct numerical simulationsmade in the astrophysical context.

PRL 107, 134501 (2011) P HY S I CA L R EV I EW LE T T E R Sweek ending

23 SEPTEMBER 2011

0031-9007=11=107(13)=134501(5) 134501-1 � 2011 American Physical Society

Homogeneous compressible turbulence.—We startour analysis with the following 3D compressible equa-tions [17]:

@t�þ r � ð�uÞ ¼ 0;

@tð�uÞ þ r � ð�uuÞ ¼ �rPþ��uþ�

3rðr � uÞ þ f;

where � is the density, u the velocity, P the pressure,� thecoefficient of viscosity, and f a stationary homogeneousexternal force acting at large scales. The system is closedwith the isothermal equation P ¼ C2

s�, where Cs is thespeed of sound. The energy equation takes the form

@thEi ¼ ��hðr � uÞ2i � 43�hðr � uÞ2i þ F; (1)

with h i an ensemble average (which is equivalent toa spatial average in homogeneous turbulence), E ¼�u2=2þ �e the total energy, e ¼ C2

s lnð�=�0Þ (�0 is aconstant density introduced for dimensional reasons), andF the energy injected.

The relevant two-point correlation functions associatedwith the total energy may be obtained by noting that forhomogeneous turbulence

h�ð�uÞ � �ui ¼ 2h�u2i � hð�þ �0Þu � u0i; (2)

h���ei ¼ 2h�ei � h�e0 þ �0ei; (3)

where for any variable �, ����ðxþrÞ��ðxÞ��0 ��.Then, we find

RðrÞ þRð�rÞ2

¼ hEi � 1

4h�ð�uÞ � �ui � 1

2h���ei;

(4)

where RðrÞ � h�u � u0=2þ �e0i � hRi and Rð�rÞ �h�0u0 � u=2þ �0ei � h ~Ri. Note that for homogeneouscompressible turbulence the relationRðrÞ ¼ Rð�rÞ holdsonly when isotropy is assumed, whereas it is always validin the incompressible limit for which R is reduced to asecond-order velocity correlation function [18]. As we willsee below, relation (4) is very helpful for deriving an exactrelation for some two-point correlation functions. In prac-tice, we shall derive a dynamical equation for RðrÞ; first,we have to compute

@th�u � u0i ¼ h�u � @tu0 þ u0 � @tð�uÞi¼

��u �

��u0 � r0u0 � 1

�0 r0P0��

þ hu0 � ½�r � ð�uuÞ � rP�i þ 2Dþ 2F ;

(5)

where for simplicity 2D and 2F denote, respectively, thecontributions to the correlation of the viscous and forcingterms. By remarking that

��

�0u �r0P0�¼

��C2

su‘@0‘�

0

�0

�¼ h�u‘@0‘e0i ¼ hr0 � ð�e0uÞi

and that

hu0 � r0ð�u �u0Þi ¼ hr0 � ½�ðu �u0Þu0���ðu �u0Þðr0 �u0Þi;we can rewrite (5) in the following way:

@th�u � u0i ¼ h�u0 � r0ð�u � u0Þ � r0 � ð�e0uÞi� hr � ½�ðu � u0Þuþ Pu0�i þ 2Dþ 2F

¼ rr � h��ðu � u0Þ�uþ Pu0 � �e0uiþ h�ðu � u0Þðr0 � u0Þi þ 2Dþ 2F : (6)

Second, we have to complete the computation with

@th�e0i ¼ h�@te0 þ e0@t�i ¼�C2s

�0 @t�0 þ e0@t�

¼��C2

s

�0 r0 � ð�0u0Þ � e0r � ð�uÞ�

¼ �hr0 � ðC2s�u

0Þi þ��0u0 � r0

�C2s

�0

��

� hr � ð�e0uÞi: (7)

By noting that��0u0‘@

0‘

�C2s

�0

��¼ �hr0 � ð�e0u0Þi þ he0r0 � ð�u0Þi;

we obtain after simplification

@th�e0i ¼ rr � h��e0�u� Pu0i þ h�e0ðr0 � u0Þi: (8)

The combination of (6) and (8) leads to

@tRðrÞ ¼ hðr0 � u0ÞRi þDþF

þrr � h�R�u� 12Pu

0 � 12�e

0ui: (9)

The same type of analysis may be performed for Rð�rÞwhich eventually leads to the dynamical equation

@t

�RðrÞ þRð�rÞ

2

�¼ 1

2hðr0 � u0ÞRi þ 1

2hðr � uÞ ~Ri

þ 1

2ðDþ ~DþF þ ~F Þ

þ 1

2rr �

��ðRþ ~RÞ�u

� 1

2ðPu0 � �e0u� P0uþ �0eu0Þ

�;

(10)

where ~D and ~F denote, respectively, the additional con-tribution of the viscous and forcing terms.Local turbulence.—For the final step of the derivation,

we shall introduce the usual assumption specific to 3Dfully developed turbulence with a direct energycascade [3,17]. In particular, we suppose the existence of

PRL 107, 134501 (2011) P HY S I CA L R EV I EW LE T T E R Sweek ending

23 SEPTEMBER 2011

134501-2

a statistical steady state in the infinite Reynolds numberlimit with a balance between forcing and dissipation. Werecall that the dissipation is a sink for the total energy andacts mainly at the smallest scales of the system. Then, far inthe inertial range we may neglect the contributions of Dand ~D in Eq. (10) [19]. The introduction of structurefunctions leads to the final form

�2" ¼ hðr0 � u0ÞðR� EÞi þ hðr � uÞð ~R� E0Þi

þ rr ���

�ð�uÞ � �u2

þ ���e� C2s���

��u

þ ��e�ð�uÞ�; (11)

where ��X � ðX þ X0Þ=2 and " is the mean total energyinjection rate [which is equal to the mean total energydissipation rate; see relation (1)]. Note that at relativelysmall r the function R� E (and ~R� E0) and its derivativeare negative since the correlation between two points ismaximum if the points are the same.

It is straightforward to show that in the limit of incom-pressible turbulence we recover the well-known expres-sion; indeed, we obtain (with � ! �0 ¼ 1)

� 4" ¼ rr � hð�uÞ2�ui; (12)

which is the primitive form of Kolmogorov’s law. Anintegration over a ball of radius r leads to the well-knownexpression [20] �ð4=3Þ"r ¼ hð�uÞ2�uri, where r meansthe radial (often called longitudinal) component, i.e., theone along the direction r.

Isotropic turbulence.—Expression (11) is the main resultof the Letter. It is an exact relation for some two-pointcorrelation functions when fully developed turbulence isassumed. It is valid for homogeneous—not necessarilyisotropic—3D compressible isothermal turbulence. Notethat the pressure contribution appears through the termC2s��� and is therefore negligible in the large Mach number

limit (Cs ! 0). When isotropy is additionally assumed,this relation can be written symbolically as

� 2" ¼ SðrÞ þ 1

r2@rðr2F rÞ; (13)

where F r is the radial component of the isotropic energyflux vector. In comparison with the incompressible case(12), expression (13) reveals the presence of a new type ofterm S which is by nature compressible since it is propor-tional to the dilatation (i.e., the divergence of the velocity).This term has a major impact on the nature of compressibleturbulence since as we will see it acts like a source or a sinkfor the mean energy transfer rate. Note that S consists oftwo terms which account for the two-point measurementapproach.

Discussion.—We may further reduce Eq. (13) by per-forming an integration over a ball of radius r. After sim-plification, we find the exact relation

� 2

3"r ¼ 1

r2

Z r

0SðrÞr2drþF rðrÞ: (14)

We start the discussion by looking at the small scalelimit of the previous relation which means that the scalesare assumed to be small enough to perform a Taylorexpansion but not too small to be still in the inertialrange. We obtain SðrÞ¼Sð0Þþr@rSð0Þ¼ r@rSð0Þ, whichleads to

� 23½"þ 3

8r@rSð0Þ�r � �23"effr ¼ F rðrÞ: (15)

Note that we do not assume the cancellation of the firstderivative of S at r ¼ 0, although the function R� Ereaches an extremum at r ¼ 0; the reason is that thisfunction is weighted by the dilatation function, whichmay have a nontrivial form. We see that at the leadingorder the main contribution of SðrÞ is to modify " forgiving an effective mean total energy injection rate "eff .Then, the physical interpretation of (15) is the following.When the flow is mainly in a phase of dilatation (positivevelocity divergence), the additional term is negative and"eff is smaller than ". On the contrary, in a phase ofcompression, @rSð0Þ is positive and "eff is larger than ".An illustration of dilation and compression effects in thespace correlation is given in Fig. 1. In both cases, the fluxvector F (dashed arrows) is oriented towards the center ofthe sphere (r ¼ 0) since a direct cascade is expected.Dilatation and compression act additionally (solid arrows):In the first case, the effect is similar to a decrease of thelocal mean total energy transfer rate, whereas in the secondcase it is similar to an increase of the local mean totalenergy transfer rate.The discussion may be extended to the entire inertial

range (i.e., for larger values of r) when the (turbulent)Mach number is relatively high. In this case, the analysisis focused on expression (14) for which we have alreadynoted that a term like R� E is mainly negative. It isinteresting to note that SðrÞ is composed of two types ofterm which are different by nature. First, there is the

FIG. 1. Dilatation (left) and compression (right) phases inspace correlation for isotropic turbulence. In a direct cascadescenario, the flux vectors (dashed arrows) are oriented towardsthe center of the sphere. Dilatation and compression (solidarrows) are additional effects which act, respectively, in theopposite or in the same direction as the flux vectors.

PRL 107, 134501 (2011) P HY S I CA L R EV I EW LE T T E R Sweek ending

23 SEPTEMBER 2011

134501-3

dilatation dominated by the smallest scales in the flow—the shocklets—which mainly give a negative contributionwith a fast variation [21]. Second, there is the correlationR� E, which derives most of its contribution from rela-tively larger scales with a slower variation. This remarkmay lead to the assumption that both terms are relativelydecorrelated [19]. Then SðrÞ may be simplified as [byusing relation (4)]

S ðrÞ ’ �h ��ðr � uÞ½14�ð�uÞ � �uþ 12���e�i: (16)

The previous expression is not derived rigorously, but itmay give us some intuition about its contribution. Forexample, we may expect a power law dependence close

to r2=3 for the structure functions. Direct numerical simu-lations have never shown a scale dependence for the dila-tation, and we may expect that it behaves like a relativelysmall factor. Then SðrÞ will still modify " as explained inthe discussion above; however, the power law dependencein r would be slightly different now. In conclusion andaccording to this simple analysis, we see that compressioneffects (through the dilatation) will mainly impact thescaling law at the largest scales.

Compressible spectrum.—Wemay try to predict a powerlaw spectrum for compressible turbulence. First, we notethat several predictions have been made for the kineticenergy spectrum and also for the spectra associated withthe solenoidal or the compressible part of the velocity [22].We recall that, although these decompositions are conve-nient for analytical developments, the associated energiesare not inviscid invariants and the predictions are heuristic.For incompressible turbulence, the situation is different

because a prediction in k�5=3 for the kinetic energy spec-trum may be proposed by applying a dimensional analysisdirectly on the 4=5 law [3]. Although it is not an exactprediction, the 4=5 law gives a stronger foundation to theenergy spectrum for which a constant flux is expected. Thisremark was already noted, in particular, in recent 3D directnumerical simulations of isothermal turbulence where it isobserved that the Kolmogorov scaling is not preserved forthe spectra based only on the velocity fluctuations [15].

We shall derive a power law spectrum for compressibleturbulence by applying a dimensional analysis on Eq. (11).Dimensionally, we may find "effr� �u3. By introducing

the density-weighted fluid velocity v � �1=3u and follow-

ing Kolmogorov, we obtain EvðkÞ � "2=3eff k�5=3, where

EvðkÞ is the spectrum associated to the variable v. Ourprediction is compatible with the measurements recentlymade by direct numerical simulations [15], where theauthors have noted that the exponent of the third-ordervelocity structure function is close to 1 if the field used isv instead of u. (Note that two other scaling relationsmay be predicted like for the pressure term.) As explainedby several authors [22], in compressible turbulence we donot expect a constant flux in the inertial range. Here, thesame conclusion is reached since we are dealing with

an effective mean energy transfer rate. More precisely,if we expect a power law dependence in k for the effectivetransfer rate, we arrive at the conclusion that asteeper power law spectrum may happen at the largestscales. According to relation (16) and the simple estimate

�v2 � r2=3, we could have EvðkÞ � k�19=9. This predictionmeans that for a small prefactor in (16) one needs anextended inertial range to feel the compressible effectson the power spectrum. The scale at which the transitionhappens between �19=9 and �5=3 may be the sonicscale ks as proposed in Ref. [23], where such powerlaws were detected; in our case, a rough estimate givesks � hðr � uÞ=�ui.Conclusion.—The present work opens important per-

spectives to further understand the nature of compressibleturbulence in the asymptotic limit of large Reynolds num-bers with the possibility to extend the analysis to magne-tized fluids with possibly other types of closures (e.g.,polytropic gas) or to improve intermittency models byusing the new relation—obtained by a statistical analysisat low order—as pivotal for a heuristic extension to statis-tical laws at higher order. We believe that astrophysics(e.g., interstellar turbulence) is one of the most importantdomains of application of the present work [10].We acknowledge S. Boldyrev, A. Kritsuk, and T. Passot

for useful discussions.

[1] A. C. Newell and B. Rumpf, Annu. Rev. Fluid Mech. 43,59 (2011); S. Galtier, Nonlinear Proc. Geophys. 16, 83(2009).

[2] A. N. Kolmogorov, Dokl. Akad. Nauk SSSR 32, 16(1941); R.A. Antonia and P. Burattini, J. Fluid Mech.550, 175 (2006).

[3] U. Frisch, Turbulence: The Legacy of A. N. Kolmogorov(Cambridge University Press, Cambridge, England, 1995).

[4] A.M. Yaglom, Dokl. Akad. Nauk SSSR 69, 743 (1949);H. Politano and A. Pouquet, Phys. Rev. E 57, R21 (1998);S. Galtier, Phys. Rev. E 77, R015302 (2008); J. Geophys.Res. 113, A01102 (2008).

[5] S. Galtier, C.R. Physique 12, 151 (2011).[6] V. E. Zakharov and R. Z. Sagdeev, Sov. Phys. Dokl. 15,

439 (1970).[7] S. Chandrasekhar, Proc. R. Soc. A 210, 18 (1951).[8] O. Zeman, Phys. Fluids A 2, 178 (1990).[9] P. Sagaut and C. Cambon, Homogeneous Turbulence

Dynamics (Cambridge University Press, Cambridge,England, 2008).

[10] T. Passot and E. Vazquez-Semadeni, Phys. Rev. E 58, 4501(1998); B. G. Elmegreen and J. Scalo, Annu. Rev. Astron.Astrophys. 42, 211 (2004).

[11] T. Passot, A. Pouquet, and P. Woodward, Astron.Astrophys. 197, 228 (1988); S. Boldyrev, A. Nordlund,and P. Padoan, Phys. Rev. Lett. 89, 031102 (2002).

[12] J. Bec and K. Khanin, Phys. Rep. 447, 1 (2007); J.-P.Bouchaud, M. Mezard, and G. Parisi, Phys. Rev. E 52,3656 (1995).

PRL 107, 134501 (2011) P HY S I CA L R EV I EW LE T T E R Sweek ending

23 SEPTEMBER 2011

134501-4

[13] A.M. Polyakov, Phys. Rev. E 52, 6183 (1995).[14] S. Lee, S. K. Lele, and P. Moin, Phys. Fluids A 3, 657

(1991); D.H. Porter, A. Pouquet, and P. R. Woodward,Phys. Rev. Lett. 68, 3156 (1992); W. Schmidt, C.Federrath, and R. Klessen, Phys. Rev. Lett. 101, 194505(2008).

[15] A. G. Kritsuk, M. L. Norman, P. Padoan, and R. Wagner,Astrophys. J. 665, 416 (2007).

[16] D. Porter, A. Pouquet, and P. R. Woodward, Phys. Rev. E66, 026301 (2002).

[17] L. Landau and E. Lifchitz, Mecanique Des Fluides (Mir,Moscow, 1989), 2nd ed.

[18] G. K. Batchelor, The Theory of Homogeneous Turbulence(Cambridge University Press, Cambridge, England, 1953).

[19] H. Aluie, Phys. Rev. Lett. 106, 174502 (2011).[20] R. A. Antonia, M. Ould-Rouis, F. Anselmet, and Y. Zhu, J.

Fluid Mech. 332, 395 (1997).[21] M.D. Smith, M.-M. Mac Low, and J.M. Zuev, Astron.

Astrophys. 356, 287 (2000).[22] J. E. Moyal, Proc. Cambridge Philos. Soc. 48, 329 (1952);

B. B. Kadomtsev and V. I. Petviashvili, Sov. Phys. Dokl.18, 115 (1973); S. S. Moiseev, V. I. Petviashvili, A. V. Toor,and V.V. Yanovsky, Physica (Amsterdam) 2D, 218 (1981).

[23] C. Federrath et al., Astron. Astrophys. 512, A81 (2010).

PRL 107, 134501 (2011) P HY S I CA L R EV I EW LE T T E R Sweek ending

23 SEPTEMBER 2011

134501-5