+ All Categories
Home > Documents > Large-scale instabilities in a nonrotating turbulent convection

Large-scale instabilities in a nonrotating turbulent convection

Date post: 05-Oct-2016
Category:
Upload: igor
View: 213 times
Download: 0 times
Share this document with a friend
12
Large-scale instabilities in a nonrotating turbulent convection Tov Elperin, Ilia Golubev, Nathan Kleeorin, and Igor Rogachevskii Citation: Phys. Fluids 18, 126601 (2006); doi: 10.1063/1.2401223 View online: http://dx.doi.org/10.1063/1.2401223 View Table of Contents: http://pof.aip.org/resource/1/PHFLE6/v18/i12 Published by the American Institute of Physics. Related Articles Heat transfer and friction characteristics of impinging jet solar air heater J. Renewable Sustainable Energy 4, 043121 (2012) Temporal evolution, morphology, and settling of the sediment plume in a model estuary Phys. Fluids 24, 086601 (2012) Rectification of evanescent heat transfer between dielectric-coated and uncoated silicon carbide plates J. Appl. Phys. 112, 024304 (2012) Reduction of mean-square advection in turbulent passive scalar mixing Phys. Fluids 24, 075104 (2012) Scaling range of velocity and passive scalar spectra in grid turbulence Phys. Fluids 24, 075101 (2012) Additional information on Phys. Fluids Journal Homepage: http://pof.aip.org/ Journal Information: http://pof.aip.org/about/about_the_journal Top downloads: http://pof.aip.org/features/most_downloaded Information for Authors: http://pof.aip.org/authors Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions
Transcript
Page 1: Large-scale instabilities in a nonrotating turbulent convection

Large-scale instabilities in a nonrotating turbulent convectionTov Elperin, Ilia Golubev, Nathan Kleeorin, and Igor Rogachevskii Citation: Phys. Fluids 18, 126601 (2006); doi: 10.1063/1.2401223 View online: http://dx.doi.org/10.1063/1.2401223 View Table of Contents: http://pof.aip.org/resource/1/PHFLE6/v18/i12 Published by the American Institute of Physics. Related ArticlesHeat transfer and friction characteristics of impinging jet solar air heater J. Renewable Sustainable Energy 4, 043121 (2012) Temporal evolution, morphology, and settling of the sediment plume in a model estuary Phys. Fluids 24, 086601 (2012) Rectification of evanescent heat transfer between dielectric-coated and uncoated silicon carbide plates J. Appl. Phys. 112, 024304 (2012) Reduction of mean-square advection in turbulent passive scalar mixing Phys. Fluids 24, 075104 (2012) Scaling range of velocity and passive scalar spectra in grid turbulence Phys. Fluids 24, 075101 (2012) Additional information on Phys. FluidsJournal Homepage: http://pof.aip.org/ Journal Information: http://pof.aip.org/about/about_the_journal Top downloads: http://pof.aip.org/features/most_downloaded Information for Authors: http://pof.aip.org/authors

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 2: Large-scale instabilities in a nonrotating turbulent convection

Large-scale instabilities in a nonrotating turbulent convectionTov Elperin,a� Ilia Golubev,b� Nathan Kleeorin,c� and Igor Rogachevskiid�

Pearlstone Center for Aeronautical Engineering Studies, Department of Mechanical Engineering,Ben-Gurion University of the Negev, POB 653, Beer-Sheva 84105, Israel

�Received 23 May 2006; accepted 29 October 2006; published online 12 December 2006�

A theoretical approach proposed by Elperin et al. �Phys. Rev. E 66, 066305 �2002�� is developedfurther to investigate formation of large-scale coherent structures in a nonrotating turbulentconvection via excitation of a large-scale instability. In particular, the convective-wind instabilitythat causes formation of large-scale coherent motions in the form of cells, can be excited in ashear-free regime. It was shown that the redistribution of the turbulent heat flux due to nonuniformlarge-scale motions plays a crucial role in the formation of the coherent large-scale structures in theturbulent convection. The modification of the turbulent heat flux results in strong reduction of thecritical Rayleigh number �based on the eddy viscosity and turbulent temperature diffusivity�required for the excitation of the convective-wind instability. The large-scale convective-shearinstability that results in the formation of the large-scale coherent motions in the form of rollsstretched along imposed large-scale velocity, can be excited in the sheared turbulent convection.This instability causes the generation of convective-shear waves propagating perpendicular to theconvective rolls. The mean-field equations that describe the convective-wind and convective-shearinstabilities, are solved numerically. We determine the key parameters that affect formation of thelarge-scale coherent structures in the turbulent convection. In particular, the degree of thermalanisotropy and the lateral background heat flux strongly modify the growth rates of the large-scaleconvective-shear instability, the frequencies of the generated convective-shear waves, and changethe thresholds required for the excitation of the large-scale instabilities. This study elucidates theorigin of the large-scale circulations and rolls observed in the atmospheric convective boundarylayers. © 2006 American Institute of Physics. �DOI: 10.1063/1.2401223�

I. INTRODUCTION

Large-scale coherent structures in a nonrotating turbu-lent convection at very large Rayleigh numbers are observedin the atmospheric convective boundary layers,1–12 in numer-ous laboratory experiments in the Rayleigh-Bénardapparatus,13–23 and in direct numerical simulations.24,25 Spa-tial scales of the large-scale coherent structures in a turbulentconvection are much larger than turbulent scales and theirlifetimes are larger than the largest time scales of turbulence.In the atmospheric shear-free convection, the structures�cloud cells� represent large, three-dimensional, long-livedBénard-type cells composed of narrow uprising plumes andwide downdrafts. They usually embrace the entire convectiveboundary layer �of the order of 1–3 km in height� and in-clude pronounced convergence flow patterns close to the sur-face. In the sheared convective flows, the structures representlarge-scale rolls �cloud streets� stretched along the meanwind.1,2,12

Coherent structures in convective turbulent flows werecomprehensively studied theoretically, experimentally, and innumerical simulations.1–28 However, some aspects related tothe origin of large-scale coherent structures in nonrotatingturbulent convection are not completely understood. Hartlep

et al. �2003� noted that there are two points of view on theorigin of large-scale circulation in turbulent convection.24

“According to one point of view, the rolls which develop atlow Rayleigh numbers near the onset of convection continu-ally increase their size as Rayleigh number is increased andcontinue to exist in an average sense at even the highestRayleigh numbers reached in the experiments.29 Anotherhypothesis holds that the large-scale circulation is a genuinehigh Rayleigh number effect.”13

Recently, a new mean-field theory of nonrotating turbu-lent convection has been developed.30,31 This theory predictsthe convective-wind instability in the shear-free turbulentconvection that results in the formation of large-scale mo-tions in the form of cells. In the sheared convection, thelarge-scale instability causes generation of convective-shearwaves. The dominant coherent structures in this case arerolls. It was demonstrated30,31 that a redistribution of the tur-bulent heat flux due to nonuniform large-scale motions playsa crucial role in the formation of the large-scale coherentstructures in turbulent convection.

In this study, a theoretical approach30,31 is developed fur-ther to investigate the formation of the coherent structures inthe nonrotating turbulent convection. In particular, we inves-tigated how the modification of the turbulent heat flux due tononuniform large-scale motions affects the critical Rayleighnumber �based on the eddy viscosity and turbulent thermalconductivity� required for the excitation of the convective-wind instability. We performed a numerical study of the

a�Electronic address: [email protected]�Electronic address: [email protected]�Electronic address: [email protected]�Electronic address: [email protected]

PHYSICS OF FLUIDS 18, 126601 �2006�

1070-6631/2006/18�12�/126601/11/$23.00 © 2006 American Institute of Physics18, 126601-1

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 3: Large-scale instabilities in a nonrotating turbulent convection

convective-wind and convective-shear instabilities in orderto determine key parameters that affect formation of thelarge-scale coherent structures in the turbulent convection.

The paper is organized as follows. In Sec. II we discussthe physics of the formation of the large-scale coherent struc-tures and formulate the mean-field equations that describethe formation of the coherent structures. In Sec. III we de-termine the critical Rayleigh number required for the excita-tion of the convective-wind instability in a shear-free turbu-lent convection. In Sec. IV we study numerically theconvective-shear instability in a sheared convection. Finally,conclusions are drawn in Sec. V.

II. TURBULENT HEAT FLUX AND MEAN-FIELDEQUATIONS

In this section we discuss a redistribution of the turbu-lent heat flux due to the nonuniform large-scale motions as akey mechanism for the formation of the large-scale coherentstructures in turbulent convection. Here we also formulatethe mean-field equations that describe the formation of thecoherent structures. Traditional theoretical models of theboundary-layer turbulence, such as the Kolmogorov-type lo-cal closures, imply the following assumptions. Fluid flowsare decomposed into organized mean motions and turbulentflow. Turbulent fluxes are assumed to be proportional to thelocal mean gradients, whereas the proportionality coeffi-cients �eddy viscosity, turbulent temperature diffusivity� areuniquely determined by local turbulent parameters. Forexample,32 the turbulent heat flux reads F��su�=−�T�S,where �T is the turbulent temperature diffusivity, S is themean entropy, and u and s are fluctuations of the velocityand entropy, respectively. This turbulent heat flux F does notinclude the contribution from anisotropic velocity fluctua-tions.

Actually, the mean-velocity gradients can directly affectthe turbulent heat flux. The reason is that additional essen-tially anisotropic velocity fluctuations are generated by tan-gling the mean-velocity gradients with the Kolmogorov-typeturbulence due to the influence of the inertial forces duringthe life time of large turbulent eddies. The Kolmogorov-typeturbulence supplies energy to the anisotropic velocity fluc-tuations that cause formation of coherent structures due tothe excitation of a large-scale instability.30,31 Anisotropic ve-locity fluctuations are characterized by a steeper spectrumthan the Kolmogorov-type turbulence.30,33–36

The theoretical model30 of the anisotropic velocity fluc-tuations and their effect on the turbulent heat flux includesthe following steps in the derivations: applying the spectralclosure, solving the equations for the second moments in thek-space, and returning to the physical space to obtain formu-las for the Reynolds stresses and the turbulent heat flux. Thederivation are based on the Navier-Stokes equation and theentropy evolution equation formulated in the Boussinesq ap-proximation. This derivation30 yields the following expres-sion for the turbulent heat flux F��su�:

F = F* − �0��Fz* div U� −

1

5� +

3

2�W � Fz

*�

−1

2�Wz � F*�� , �1�

where �0 is the correlation time of turbulent velocity corre-sponding to the maximum scale of turbulent motions,W=��U is the mean vorticity, U=U�+Uz is the meanvelocity with the horizontal U� and vertical Uz components,� is the degree of thermal anisotropy, and

Fi* = − �ij� jS − �0Fz

*�zUi�0��z� �2�

is the background turbulent heat flux that is the sum of thecontribution due to the Kolmogorov-type turbulence �de-scribed by the first term in Eq. �2�� and a contribution of theanisotropic turbulence caused by the shear of the imposedlarge-scale mean velocity U�0��z� �the so-called counter-windheat flux described by the second term in Eq. �2��,

�ij = �T��ij + beiej� �3�

is a generalized anisotropic turbulent temperature diffusivitytensor. For turbulent convection, b= �3/2��2+ ��, � is the ra-tio of specific heats �e.g., �=7/5 for the air flow� and e is thevertical unit vector. The equation for the tensor �ij was de-rived in Appendix A in Ref. 30 using the budget equationsfor the turbulent kinetic energy, fluctuations of the entropyand the turbulent heat flux. The anisotropic part of the tensor�ij �described by the second term in the square brackets ofEq. �3��, is caused by a modification of the turbulent heatflux by the buoyancy effects. Note that for a laminar convec-tion b is set to zero and the coefficient of the turbulent tem-perature diffusivity �T should be replaced by the coefficientof the molecular temperature diffusivity. The parameter � inEq. �1� is given by

� =1 + 4�

1 + �/3, � = l�

lz2/3

− 1, �4�

where l� and lz are the horizontal and vertical scales, respec-tively, in which the background turbulent heat flux Fz

*�r�= �s�x�uz�x+r�� tends to zero. The parameter � describes thedegree of thermal anisotropy. In particular, in isotropic casewhen l�= lz the parameter �=0 and �=1. For l�� lz, theparameter �=−1 and �=−9/2. The maximum value �max ofthe parameter � is given by �max=2/3 for �=3. The upperlimit for the parameter � arises because the function Fz

*�r�has a global maximum at r=0. Depending on the parameter�, the small-scale thermal structures in the background tur-bulent convection have the form of columns or pancakes�sometimes they are called small-scale thermal plumes�. For�1, the small-scale thermal structures have the form ofcolumns �l� lz�, and for �1 there exist pancake thermalstructures �l� lz� in the background turbulent convection�i.e., a turbulent convection with zero gradients of the meanvelocity�.

The terms in the square brackets in the right-hand side ofEq. �1� result from the anisotropic turbulence and depend onthe �mean� �including coherent� velocity gradients. Theseterms lead to the excitation of large-scale instability and for-

126601-2 Elperin et al. Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 4: Large-scale instabilities in a nonrotating turbulent convection

mation of coherent structures. In Eq. �1�, the terms with zerodivergence are omitted, because only div F contributes to themean-field dynamics. Neglecting the anisotropic turbulenceterm in Eq. �1� recovers the traditional equation for the tur-bulent heat flux.

The physical meaning of Eq. �1� is the following. The

first term �−�0�Fz* div U� in square brackets in Eq. �1� de-

scribes the redistribution of the vertical background turbulentheat flux Fz

* by the perturbations of the convergent �or diver-

gent� horizontal mean flows U�. This redistribution of thevertical turbulent heat flux occurs during the lifetime of tur-

bulent eddies. The second term ��0��+3/2��W�Fz*� in

square brackets in Eq. �1� determines the formation of thehorizontal turbulent heat flux due to �rotation� of the verticalbackground turbulent heat flux Fz

* by the perturbations of the

horizontal mean vorticity W�. The last term ��0�Wz�F*� insquare brackets in Eq. �1� describes the formation of thehorizontal heat flux through the “rotation” of the horizontalbackground heat flux F�

* �the “counter-wind” heat flux in Eq.�2�� by the perturbations of the vertical component of the

mean vorticity Wz. These three effects are determined by thelocal inertial forces in inhomogeneous mean flows. A moredetailed discussion of Eq. �1� is given in Secs. III and IV.

The counter-wind turbulent heat flux �in the directionopposite to the mean wind� is well known in the atmosphericphysics and arises due to the following reason. In the shearedturbulent convection an ascending fluid element has largertemperature then that of surrounding fluid and smaller hori-zontal fluid velocity, while a descending fluid element hassmaller temperature and larger horizontal fluid velocity. Thiscauses the background turbulent heat flux F�

* =−�0Fz*�zU�

�0�

in the direction opposite to the background horizontal meansheared fluid velocity U

�0��z�.We use the mean-field approach whereby the small-scale

turbulent convection is parametrized. This is the reason wedo not explicitly use thermal plumes in the consideration.The main reason for the appearance of the large-scale coher-ent structures is related to the modification of the heat flux bythe nonuniform mean flows. The thermal plumes contributeto the modification of the turbulent heat flux. To some extent,the redistribution of the turbulent heat flux can be interpretedas a redistribution of the thermal plumes.

In order to study the formation of the large-scale coher-ent structures in a small-scale nonrotating turbulent convec-tion we used the mean-field Navier-Stokes equation and themean entropy evolution equation �with the turbulent heatflux �1�� formulated in the Boussinesq approximation. Thesemean-field equations yield the following linearized equations

for the small perturbations from the equilibrium, U=Uz

−Uz�0�, W=Wz−Wz

�0�, and S=S−S�0�:

�t+ Uy

�0��y − �T U = g �S , �5�

�t+ Uy

�0��y − �T W = − ��xU , �6�

�t+ Uy

�0��yS = − �� · F� − ��zS�0��U , �7�

where �T is the eddy viscosity, �= −�z2, and

� · F = −4�0

45��F* · e��10� � − �8� − 3� �U

+ 6��F* � e� · ��W − �T� + b�z2�S , �8�

�T is the turbulent temperature diffusivity. In order to deriveEq. �5�, the pressure term was excluded by calculating thecurl of the momentum equation. Equations �5�–�7� allow usto study the linear stage of the large-scale instabilities. The

variables U, W, and S describe the large-scale coherent struc-tures. In Sec. III we study a shear-free convection withU�0�=0, and in Sec. IV we investigate turbulent convectionwith the background �equilibrium� large-scale velocity shearU�0��z�=�zey and the background mean vorticity W�0�=��U�0�=−�ex.

III. SHEAR-FREE CONVECTION

Let us consider a shear-free convection �U�0�=0�. In theshear-free regime, the large-scale instability is related to the

first term �−�0�Fz* div U� in square brackets in Eq. �1� for

the turbulent heat flux.30,31 When �Uz /�z0, perturbations

of the vertical velocity Uz cause negative divergence of the

horizontal velocity, div U�0 �provided that div U=0�.This strengthens the local vertical turbulent heat flux andcauses increase of perturbations of the local mean entropyand buoyancy. The latter enhances perturbations of the local

mean vertical velocity Uz, and by this means, the convective-wind instability is excited. Similar reasoning is valid when

�Uz /�z0, whereas div U�0. Negative perturbations ofthe vertical flux of entropy then lead to a decrease of pertur-bations of the mean entropy and buoyancy, which enhancesthe downward flow and once again excites the convective-

wind instability. Therefore, nonzero div U� causes redistri-bution of the vertical turbulent heat flux and formation ofregions with large values of this flux. These regions �where

div U�0� alternate with the low heat flux regions �where

div U�0�. This process results in formation of the large-scale coherent structures.

The role of the second term ��0��+3/2��W�Fz*� in

square brackets in Eq. �1� is to decrease the growth rate ofthe large-scale instability for �−3/2. Indeed, the interac-tion of perturbations of the mean vorticity with the verticalbackground turbulent heat flux Fz

* produces the horizontalturbulent heat flux. The latter decreases �increases� the meanentropy in the regions with upward �downward� local flows,thus diminishing the buoyancy forces and reducing the mean

vertical velocity Uz and the mean vorticity W. This mecha-nism dampens the convective-wind instability for �−3/2.The above two competitive effects determine the growth rateof the convective-wind instability. A solution of Eqs. �5� and�7� in the shear-free convection regime yields the followingexpression for the growth rate � of long-wave perturbations:

126601-3 Large-scale instabilities in a nonrotating Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 5: Large-scale instabilities in a nonrotating turbulent convection

� � gFz*�0

2K2���sin ���� −3

8−

5�

4sin2 ��1/2

, �9�

where the parameter �= �l0K�−2�1, l0 is the maximum scaleof turbulent motions, � is the angle between the vertical unitvector e and the wave vector K of small perturbations, and gis acceleration of gravity. The analysis of the convective-wind instability was performed in Refs. 30 and 31 only for asmall square of Brunt-Väisälä frequency. In particular, Eq.�9� was derived in Refs. 30 and 31 for the case �N2��gFz

*�0K2, where N2=−�g ·��S�0�0 is the square of Brunt-Väisälä frequency.

In the present study we consider arbitrary values of theBrunt-Väisälä frequency, and we investigate the effect of themodification of the turbulent heat flux �due to nonuniformlarge-scale motions� on the critical effective Rayleigh num-ber required for the excitation of the convective-wind insta-bility. We also study here the effect of the anisotropy ofturbulent thermal diffusivity �caused by the buoyancy� on thecritical effective Rayleigh number. To this end, we rewriteEqs. �5� and �7� in a nondimensional form:

�t− V = Ra �S , �10�

PrT�S

�t− � + b�z

2�S = V + �PrT

Ra1/3

��10� � − �8� − 3� �V , �11�

where V=PrT ��U is dimensionless velocity, the length ismeasured in the units of the total vertical size Lz of the sys-tem, the parameter b=3�2+ �� /2 describes the anisotropy ofturbulent thermal diffusivity caused by the buoyancy effect�see Eqs. �3��, Ra=6�3�PrT is effective Rayleigh numberbased on the turbulent viscosity, �T, and the turbulent tem-perature diffusivity, �T, �=Lz / l0, PrT=�T /�T is the turbulentPrandtl number, the parameter � is given by

� =4a*

15 6

�21/3

, � =6gl0

u02

�T

T0,

�T is the mean temperature difference between bottom andupper boundaries of the turbulent convection, the parametera*=2g�0Fz

* /u02, and T0 is the reference mean temperature.

The last term �� in the right-hand side of Eq. �11� deter-mines the modification of the turbulent heat flux due to thenonuniform large-scale motions, and the parameter

�PrT

Ra1/3

=4g�0Fz

*

�N2�Lz2

has the meaning of the normalized heat flux, where �N2�=g�T / �T0Lz� and Ra= �N2�Lz

4 / ��T�T�.

A. Solution for two free boundaries

Let us consider the solution of Eqs. �10� and �11� for twofree boundaries, using the following boundary conditions

V = �z2V = S = 0 for z = 0;1. �12�

We seek for a solution of Eqs. �10� and �11� in the form

V, S � sin��nz�exp��t − iK� · r� ,

where n is the integer number and K� is the horizontal com-ponent of the wave vector. The critical effective Rayleighnumber �at �=0� is determined by the equation

�K�2 + �b + 1��2n2��K�

2 + �2n2�2

= K�2 �Rac − ��PrTRac

2�1/3

���2� + 3�K�2 − �8� − 3��2n2� ,

where the critical effective Rayleigh number Rac is based onthe turbulent viscosity and the turbulent temperature diffu-sivity.

In the case of �=0 �i.e., there is no modification of theturbulent heat flux due to the nonuniform large-scale mo-tions�, the critical effective Rayleigh number is given by

Rac =�K�

2 + �2n2�2�K�2 + �b + 1��2n2�

K�2 . �13�

The minimum value of the critical effective Rayleigh numberfor the first mode �n=1� for b=0 is Rac�657.5. This is theclassical Rayleigh solution for the laminar convection withtwo free boundaries. The critical effective Rayleigh numberincreases with the increase of the anisotropy of turbulenttemperature diffusivity �see Eqs. �3�� described by the pa-rameter b. Indeed, for b=3.9, the critical effective Rayleighnumber is Rac�2247, and for b=5.1 it is Rac�2722.

The modification of the turbulent heat flux due to non-uniform large-scale motions strongly decreases the criticaleffective Rayleigh number. Indeed, Fig. 1 shows the effectiveRayleigh number versus the aspect ratio Lz /L��K� /Kz

=tan � of the perturbations for different values of parameter�. The increase of the parameter � causes strong reductionof the critical Rayleigh number �see Table I�.

FIG. 1. Effective Rayleigh number versus the aspect ratio Lz /L� of theperturbations for two free boundaries and different values of the parameter�: �=0 �dashed line�; for �=0.7 �dotted line�; �=5 �dashed-dotted line�.Here, �=1 and b=5.1. The classical Rayleigh solution for the laminar con-vection �b=0� with the two free boundaries is shown by solid curve.

126601-4 Elperin et al. Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 6: Large-scale instabilities in a nonrotating turbulent convection

B. Solution for two rigid boundaries

Now let us consider the solution of Eqs. �10� and �11� fortwo rigid boundaries. In view of the symmetry of this prob-lem with respect to two bounding planes, it is convenient totranslate the origin of z to be midway between the twoplanes. Fluid is then confined between two planes z= ±1/2,and we seek for a solution of Eqs. �10� and �11� satisfyingthe following boundary conditions

V = �zV = S = 0 for z = ± 12 . �14�

We seek for the solution of Eqs. �10� and �11� in the form V,

S�exp��t+qz− iK� ·r�, where the critical effective Rayleighnumber is determined by the equation

�K�2 − �1 + b�q2��K�

2 − q2�2

= K�2 �Rac − ��PrTRac

2�1/3

���2� + 3�K�2 + �8� − 3�q2� . �15�

The problem is symmetric with respect to the two boundariesso the eigenfunctions fall into two distinct classes: the evenmode with vertical velocity symmetry with respect to themidplane and the odd mode with vertical velocity asymme-try. Following the procedure described in Refs. 37–39, weadopted the even solution that has minimum critical effectiveRayleigh number. Our numerical analysis showed that theanisotropy of turbulent temperature diffusivity described bythe parameter b increases the critical effective Rayleighnumber. In particular, for b=0 the critical effective Rayleighnumber is Rac�1707.8. This is classical Rayleigh solutionfor the laminar convection with the two rigid boundaries. For�=0 and b=3.9, the critical effective Rayleigh number isRac�4683, and for b=5.1 it is Rac�5547.

The effective Rayleigh number versus the aspect ratioLz /L� of the perturbations for two rigid boundaries is plottedin Fig. 2. Increasing of the parameter � decreases both thecritical effective Rayleigh number and the aspect ratio Lz /L�

of perturbations �see Table I�. If ��5, the behavior of theeffective Rayleigh number drastically changes; i.e., there aretwo local minima for the effective Rayleigh number.

C. Solution for one rigid and one free boundaries

Solution for one rigid and one free boundary can beobtained from the results for two rigid boundaries using theodd mode. We use the domain from z=0 �the free boundary�to z=1/2 �the rigid boundary�. The anisotropy of turbulenttemperature diffusivity described by the parameter b in-creases the critical effective Rayleigh number. Indeed, forb=0, the critical effective Rayleigh number is Rac�1101�the classical Rayleigh solution for the laminar convection�.For �=0 and b=3.9, the critical effective Rayleigh numberis Rac�3359, and for b=5.1 it is Rac�4023. The effectiveRayleigh number versus the aspect ratio Lz /L� of the pertur-bations for the one rigid and one free boundary is plotted inFig. 3. Increasing the parameter � decreases the critical ef-fective Rayleigh number and reduces the aspect ratio Lz /L�

of the perturbations �see Table I�.Therefore, for these three types of boundaries the modi-

fication of the turbulent heat flux due to the nonuniformlarge-scale motions strongly reduces the critical effectiveRayleigh number �based on the eddy viscosity and turbulenttemperature diffusivity� required for the excitation of theconvective-wind instability. We summarized the final resultsfor the above three types of the boundary conditions in Table

TABLE I. Critical effective Rayleigh numbers for different types of the boundaries. Here for the turbulent flow�=1 and b=5.1. The case of laminar convection is presented in Table I only for comparison with the resultsobtained for the turbulent convection.

Case

Boundaries

Two free One free and one rigid Two rigid

Lz /L� Rac Lz /L� Rac Lz /L� Rac

Laminar flow 0.707 657.5 0.854 1101 0.994 1708

Turbulent flow

�=0 0.891 2722 1.096 4023 1.280 5547

�=0.7 0.613 1076 0.697 1328 0.754 4218

�=2.0 0.578 344 0.645 420 0.688 1743

�=5.0 0.568 98 0.628 120 0.662 549

FIG. 2. Effective Rayleigh number versus the aspect ratio Lz /L� of theperturbations for two rigid boundaries and different values of the parameter�: �=0 �dashed line�; �=0.7 �dotted line�; �=5 �dashed-dotted line�. Here,�=1 and b=5.1. The classical Rayleigh solution for the laminar convection�b=0� with the two rigid boundaries is shown by the solid curve.

126601-5 Large-scale instabilities in a nonrotating Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 7: Large-scale instabilities in a nonrotating turbulent convection

I. The case of laminar convection is presented in Table I onlyfor comparison with the results obtained for the turbulentconvection.

IV. SHEARED TURBULENT CONVECTION

In this section we consider turbulent convection with alarge-scale linear velocity shear U�0��z�=�zey. In a shearedturbulent convection the mechanism of the convective-shear

instability30,31 is related to the last term ��0�Wz�F*� insquare brackets in Eq. �1�. The generation of the potentialtemperature perturbations by vorticity perturbations plays thekey role in this mechanism. Indeed, in two adjacent vortices

with the opposite directions of the vertical vorticity Wz, theturbulent fluxes of potential temperature are directed towardsthe boundary between the vortices. This increases perturba-tions of the mean potential temperature and the buoyancy,and generates the upward flow between the vortices. Thesevertical flows excite vorticity perturbations, and theconvective-shear instability mechanism is sustained.

Let us consider an evolution of perturbations with zero

y-derivatives of the fields U, W, and S. We seek for a solu-tion of Eqs. �5�–�7� in the form �exp��t− iK ·r�. The growthrate of the convective-shear instability of long-wave pertur-bations is given by30,31

� � gFz*�0

2K2��� sin2 ��2/3, �16�

where �=��0 is the shear parameter, the parameter���l0K�−2�1 and K=�Kx

2+Kz2. The convective-shear insta-

bility causes formation of large-scale coherent fluid motionsin the form of rolls �see Fig. 4� aligned along the imposedmean velocity U�0�. The instability can also result in genera-tion of the convective-shear waves with the frequency

� � �3gFz*�0

2K2��� sin2 ��2/3, �17�

which implies the wave-number dependence ��K2/3. Theconvective-shear waves propagate perpendicular to convec-tive rolls �see Fig. 4�. The analysis of the convective-shearinstability was performed in Refs. 30 and 31 only for a small

square of Brunt-Väisälä frequency and zero y-derivatives of

the fields U, W, and S. This corresponds to the convective-shear instability for a very small component of the wavenumber along the imposed mean shear �i.e., uniform pertur-bations along the large-scale shear velocity�. In this case thegrowth rate of the convective-shear instability is maximum.

In the present study we consider arbitrary values of theBrunt-Väisälä frequency and perform the numerical analysisof the convective-shear instability for nonzero y-derivatives

of the fields U, W, and S. We consider the eigenvalue prob-lem with boundary conditions. We seek for a solution of Eqs.�5�–�7� in the form ���z�exp��t− iK� ·r�, where the eigen-function ��z� and the growth rate � of the convective-shearinstability are determined by Eqs. �5�–�7�. The system of theordinary differential equations for the eigenvalue problem issolved numerically with the following boundary conditions:

U= U�= UIV=W= S=0 at z=0, and U�=W�= S�=0 at z=1,where f�=df /dz. We also take into account that for a turbu-lent convection, the turbulent Prandtl number can be esti-mated as PrT

−1�4/ �1+Pr��2.34 with Pr=0.71 �for air flow�.The latter estimate follows from the balance equations forthe turbulent heat flux, the entropy fluctuations and the tur-bulent kinetic energy.30

Let us consider the thermally isotropic ��=1� tur-bulent convection. Figure 5 shows the range of parameters

FIG. 3. Effective Rayleigh number versus the aspect ratio Lz /L� of theperturbations for one rigid and one free boundary and different values of theparameter �: �=0 �dashed line�; �=0.7 �dotted line�; �=5 �dashed-dottedline�. Here, �=1 and b=5.1. The classical Rayleigh solution for the laminarconvection �b=0� with one rigid and one free boundary is shown by solidcurve.

FIG. 4. Large-scale coherent rolls formed due to the convective-shear insta-bility and aligned along the sheared mean velocity U�0��z�. The instabilityresults in generation of the convective-shear waves which propagate perpen-dicular to the convective rolls.

FIG. 5. Range of parameters �Lz /L�; L / l0� for which the convective-shearinstability occurs, for different values of the shear parameter�=0; 0.05; 0.1; 0.2. Here, �=1.

126601-6 Elperin et al. Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 8: Large-scale instabilities in a nonrotating turbulent convection

�Lz /L�; L / l0�, for which the convective-shear instability oc-curs for different values of the shear parameter �. Here,L�1/�Lz

−2+L�−2, and we assumed that a*=1. The case

�=0 in Fig. 5 corresponds to the convective-wind instability�the shear-free turbulent convection�. Inspection of Fig. 5shows that the increase of shear is favorable for the excita-tion of the convective-shear instability. In Fig. 6 we plottedthe range of parameters �Lz /L�; L / l0� for the convective-shear instability for different values of the angle � betweenthe horizontal wave vector K� and the x axis. Increasing theangle � prevents from the excitation of the convective-shearinstability �i.e., reduces the range of parameters for which theinstability occurs�. In Fig. 7 we plotted the growth rates ofthe convective-shear instability and the frequencies of thegenerated convective-shear waves versus Lz /L� and L / l0.The curves in Fig. 7 have a point L* whereby the first de-rivative d� /dK has a singularity, which is indicative of bi-furcation. The growth rate of the convective-shear instability

for very small y-derivatives of the fields U, W, and S, isdetermined by cubic algebraic equation.30 Below the bifur-cation point, the cubic equation has three real roots �whichcorresponds to aperiodic instability without generation ofwaves�. Above the bifurcation point, the cubic equation hasone real and two complex conjugate roots. In this case theconvective-shear waves are generated. The source of energyfor these waves is the turbulence energy.

Now we perform the detailed numerical analysis of theconvective-shear instability in order to determine the keyparameters that affect this instability. First, we study the ef-fect of the thermal anisotropy � on the convective-shear in-stability. Figure 8 shows the range of parameters �Lz /L�;L / l0� for which the convective-shear instability occurs, fordifferent values of the thermal anisotropy �. In Fig. 9 weplotted the growth rates of the convective-shear instabilityand the frequencies of the generated convective-shear wavesfor different values of �. The decrease of the degree of ther-mal anisotropy � increases the threshold in the parameterL / l0 required for the excitation of the convective-shear insta-

bility. Figure 10 shows the growth rates of the convective-shear instability versus the angle � between the horizontalwave vector K� and the x axis for different values of �.Here, the values Lz /L� and L / l0 correspond to the maximumgrowth rates of the instability. For �0.7 the growth rate ofthe convective-shear instability attains the maximum for�m0°. An increase of the degree of thermal anisotropy �increases the angle �m. In the thermally isotropic ��=1� tur-bulent convection the angle �m=18°, while for �=0.8 �i.e.,��0.92�, the angle �m�10°. Note that according to the at-mospheric observations, the observed angle between thecloud streets and direction of the wind is of the order of10°–14°. The calculated angle �m is in compliance with theseobservations. Note that the convective rolls are stretched inthe horizontal plane in the direction perpendicular to K� andthe shear velocity is directed along the y axis. Inspection ofFigs. 9 and 10 shows that decrease of the parameter � re-duces the growth rates of the convective-shear instability. InFigs. 8–10 we considered the case ��1, which is of interestin view of the atmospheric applications.

Next, we study the effect of the lateral background heat

flux �determined by the third term ���F*�e� ·��W in

FIG. 6. Range of parameters �Lz /L�; L / l0� for which the convective-shearinstability occurs, for different values of the angle � between the horizontalwave vector and the x axis: �=0° �solid line�; �=18° �dashed line�;�=30° �dotted line�; �=90° �dashed-dotted line�. Here, �=1 and �=0.1.

FIG. 7. Growth rates of the convective-shear instability and the frequenciesof the generated convective-shear waves versus: Lz /L� and L / l0. Corre-sponding dependencies on the parameters L / l0 are given for different Lz /L�

and vice versa. Here, �=1 and �=0.1.

126601-7 Large-scale instabilities in a nonrotating Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 9: Large-scale instabilities in a nonrotating turbulent convection

the right-hand side of Eq. �8� , on the convective-shear in-stability. We introduce the angle � between the horizontalcomponent F�

* of the background turbulent heat flux andx axes, where the total background heat flux isF*= �F�

* cos � ,F�* sin � ,Fz

*�. The angle � is determined bythe boundary conditions in the horizontal plane �e.g., by thetemperature gradient in the horizontal plane�. Figure 11shows the range of parameters �Lz /L�; L / l0� for which theconvective-shear instability occurs, for different directions �of the lateral background heat flux F�

* . In Figs. 12 and 13 weplotted the growth rates of the convective-shear instabilityand the frequencies of the generated convective-shear wavesfor this case, where F�

* /Fz*=0.5. Note that the background

mean vorticity due to the imposed large-scale shear is W�0�

=�ÃU�0�=−�ex. This is the reason there is no symmetrywith respect to the Y -Z plane of the large-scale shear; i.e.,the contributions to the convective-shear instability caused

by the positive and negative angles � of the lateral back-ground heat flux are different. In particular, the range of theconvective-shear instability in the presence the lateral back-ground heat flux with the positive angles � is wider than thatfor the negative angles � �see Fig. 11�. On the other hand,even for the negative angles � the range of the convective-shear instability is wider than that in the absence of the lat-eral background heat flux. Note also that in the presence ofthe lateral background heat flux with the positive angles �,the convective-shear waves are not generated. This is reasonwe plotted in Figs. 12�c� and 13�c� the frequencies of thegenerated convective-shear waves only for ��0.

Note that there are three groups of parameters in thisstudy of the large-scale coherent structures formed in a tur-bulent convection:

�i� the external parameters: the value of shear � and thebackground heat flux F*= �F�

* cos � ,F�* sin � ,Fz

*�;�ii� the parameters that determine the background turbu-

lent convection: the degree of thermal anisotropy �, the cor-relation time �0= l0 /u0, and the parameter a*=2g�0Fz

* /u02;

FIG. 8. Range of parameters �Lz /L�; L / l0� for which the convective-shearinstability occurs, for �=0.1 for different values of the degree of thermalanisotropy �: �=1 �solid line�; �=0.9 �dashed line�; �=0.8 �dotted line�;�=0.7 �dashed-dotted line�.

FIG. 9. Growth rates of the convective-shear instability and the frequenciesof the generated convective-shear waves for different values of the degree ofthermal anisotropy �: �=1 �solid line�; �=0.9 �dashed line�; �=0.8 �dottedline�; �=0.7 �dashed-dotted line�. Here �=0.1 and L / l0=23.

FIG. 10. Growth rates of the convective-shear instability versus the angle �between the horizontal wave vector and the x axis for different values of thedegree of thermal anisotropy �: �=1 �solid line�; �=0.9 �dashed line�;�=0.8 �dotted line�; �=0.7 �dashed-dotted line�. Here, �=0.1; the valuesLz /L� and L / l0 correspond to maximum growth rates of the instability.

FIG. 11. Range of parameters �Lz /L�; L / l0� for which the convective-shearinstability occurs, for different directions � of the lateral background heatflux: �=−45°; �=0°; �=45°. Here, �=0.1.

126601-8 Elperin et al. Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 10: Large-scale instabilities in a nonrotating turbulent convection

�iii� the parameters related to the characteristics of thelarge-scale coherent structures: the aspect ratio of the struc-ture L� /Lz, the minimum size L of the structure,Lx=L� cos �, the characteristic time of the formation of thelarge-scale coherent structures ��−1, and the frequency � ofthe generated convective-shear waves.

The parameters related to the characteristics of the large-scale coherent structures are determined in this study. Theexternal parameters and the parameters that determine thebackground turbulent convection at the present level ofanalysis are treated as free parameters. The external param-eters are determined by the boundary conditions. The degreeof thermal anisotropy � can be determined by the budgetequation for the two-point correlation function for thevelocity-entropy fluctuations. This parameter has been re-cently measured in a laboratory experiment in turbulent con-vection in air-flow.23 In the range of the Rayleigh numbers107–108 �based on the kinematic viscosity and moleculardiffusivity� this parameter varies within the range from 0.5 to2. The parameter a* and the correlation time �0= l0 /u0 can bedetermined from the budget equations for the turbulent ki-netic energy and vertical turbulent fluxes of momentum andthe entropy. The turbulent correlation time �0 and correlationlength l0 are measured in laboratory convection �see, e.g.,Ref. 23�.

V. DISCUSSION

In the present study we investigated formation of large-scale coherent structures in a nonrotating turbulent convec-tion due to an excitation of large-scale instabilities. In theshear-free turbulent convection, the cell-like structures areformed due to the convective-wind instability. The redistri-bution of the turbulent heat flux due to the nonuniform large-scale motions causes strong reduction of the critical effectiveRayleigh number required for the excitation of theconvective-wind instability. The effective Rayleigh numberis based on the eddy viscosity and turbulent thermal conduc-tivity. We also found that the critical effective Rayleigh num-ber increases with the increase of the anisotropy of turbulenttemperature diffusivity caused by the buoyancy effects.

In the sheared turbulent convection, the roll-like struc-tures stretched along the imposed large-scale sheared veloc-ity are formed due to the large-scale convective-shear insta-bility. This instability produces the convective-shear wavespropagating perpendicular to the convective rolls. We studiednumerically the convective-shear instability and determinedthe key parameters that affect the formation of the large-scalecoherent structures in the turbulent convection. In particular,we found that the degree of thermal anisotropy and the lat-eral background heat flux strongly modify the growth ratesof the large-scale convective-shear instability, the frequen-

FIG. 12. Growth rates of the convective-shear instability and the frequen-cies of the generated convective-shear waves versus Lz /L� for differentdirections � of the lateral background heat flux: �=−45°; �=0°; �=45°. �a�The growth rates of the instability for L / l0=10. �b� The growth rates of theinstability for L / l0=20. �c� The frequencies of the generated waves forL / l0=10 �solid line� and L / l0=20 �dashed line�. Here, �=0.1.

FIG. 13. Growth rates of the convective-shear instability and the frequen-cies of the generated convective-shear waves versus L / l0 for different direc-tions � of the lateral background heat flux: �=−45°; �=0°; �=45°. �a� Thegrowth rates of the instability for Lz /L�=0.5. �b� The growth rates of theinstability for Lz /L�=2. �c� The frequencies of the generated waves forLz /L�=0.5 �solid line� and Lz /L�=2 �dashed line�. Here, �=0.1.

126601-9 Large-scale instabilities in a nonrotating Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 11: Large-scale instabilities in a nonrotating turbulent convection

cies of the generated convective-shear waves and change theinstability thresholds.

The results described in this study are based on the lin-earized mean-field equations, and therefore, they cannot de-scribe detail features of the turbulent convection observed inthe numerous laboratory experiments13–22 and in direct nu-merical simulations.24,25 In particular, we made the followingassumptions about the turbulent convection. We considered ahomogeneous, incompressible background turbulent convec-tion �i.e., the turbulent convection without mean-velocitygradients�. The nonuniform mean velocity affects the back-ground turbulent convection; i.e., it causes generation of theadditional strongly anisotropic velocity fluctuations by tan-gling of the mean-velocity gradients with the backgroundturbulent convection. We assumed that the generated aniso-tropic fluctuations do not affect the background turbulentconvection. This implies that we considered a one-way cou-pling due to a weak inhomogeneity of the large-scale veloc-ity. Thus, we studied simple physical mechanisms to describean initial stage of the formation of large-scale coherent struc-tures in a nonrotating turbulent convection. The simplemodel considered in our paper can only mimic the real flowsassociated with laboratory turbulent convection. Clearly, thecomprehensive theoretical and numerical studies are requiredfor quantitative description of the laboratory turbulent con-vection.

Although this model is very simple, it reproduces someproperties of the semi-organized structures observed in theatmospheric turbulent flows.31 The semi-organized structuresare observed in the form of rolls �cloud streets� or three-dimensional convective cells �cloud cells�. The observedangle between the cloud streets and the mean horizontalwind of the sheared turbulent convection is about 10°–14°,the lengths of the cloud streets vary from 20 to 200 km, thewidths from 2 to 10 km, and convective depths from2 to 3 km. The ratio of the minimal size of the structure tothe maximum scale of turbulent motions L / l0=10–100. Thecharacteristic lifetime of rolls varies from 1 to 72 h. Rollsmay occur over water surface or land surfaces.1,2 Our studyyield the following parameters of the convective rolls: L / l0

=10–100, the characteristic time of formation of the rolls�1/� varies from 1 to 3 h. The lifetime of the convectiverolls is determined by a nonlinear evolution of theconvective-shear instability. The latter is a subject of a sepa-rate ongoing study. We have shown that the maximumgrowth rate of the convective-shear instability is attainedwhen the angle between the cloud streets and the mean hori-zontal wind of the convective layer is about 10°–17° inagreement with observations. We also found an excitation ofthe convective-shear waves propagating perpendicular toconvective rolls. This finding is in agreement with observa-tions in the atmospheric convective boundary layer, wherebythe waves propagating perpendicular to cloud streets havebeen detected.11 In addition, the motions in the convectiverolls have a nonzero helicity in agreement with predictionsmade in Ref. 40.

There are two types of cloud cells in the atmosphericshear-free turbulent convection: open and closed. Open-cellcirculation has downward motion and clear sky in the cell

center, surrounded by cloud associated with upward motion.Closed cells have the opposite circulation.2 Both types ofcells have diameters ranging from 10 to 40 km, they occur ina convective layer with a depth of about 1 to 3 km, and thecharacteristic lifetime of cloud cells is about several hours.Our analysis shows that the minimum threshold value of theeffective Rayleigh number required for the excitation of thelarge-scale instability is attained at L� /Lz=2 �see Figs. 1–3,dotted and dashed-dotted curves�, is in agreement withnumerous observations. The ratio of the minimum size ofthe structure to maximum scale of turbulent motionsL / l0=5–15. The characteristic time of formation of the con-vective cells �1/� varies from 1 to 3 h. Therefore, the pre-dictions of the developed theory are in agreement with ob-servations of the semi-organized structures in theatmospheric convective boundary layer. The typical temporaland spatial scales of structures are always much larger thenthe turbulence scales. This justifies the separation of scales,which was assumed in the suggested in the theory. Note thatthe applicability of the mean-field equations for study of tur-bulent convection was discussed in Ref. 41.

In our study we consider nonrotating turbulent convec-tion and apply our results to the atmospheric convectiveboundary layers, where the shear is usually caused by wind.The rotation of the Earth usually affects the hight of theatmospheric convective boundary layer. The rotation can alsoaffect the longitudinal spatial structure of the cloud streets.Note that in astrophysical applications the shear �or differen-tial rotation� can in general be a consequence of anisotropiesin rotating systems. Our study can be also useful for under-standing the origin of formation of the mesogranular struc-tures in the solar convection �see Ref. 42�.

ACKNOWLEDGMENTS

The authors benefited from stimulating discussions withF. H. Busse, D. Etling, H. J. S. Fernando, R. Foster, A.Tsinober, and S. Zilitinkevich. This work was partially sup-ported by the Israel Science Foundation governed by theIsraeli Academy of Science and the Israeli UniversitiesBudget Planning Committee �VATAT� and Israel Atomic En-ergy Commission.

1D. Etling and R. A. Brown, “Roll vortices in the planetary boundary layer:a review,” Boundary-Layer Meteorol. 65, 215 �1993�.

2B. W. Atkinson and J. Wu Zhang, “Mesoscale shallow convection in theatmosphere,” Rev. Geophys. 34, 403 �1996�.

3D. H. Lenschow and P. L. Stephens, “The role of thermals in the convec-tive boundary layer,” Boundary-Layer Meteorol. 19, 509 �1980�.

4J. C. R. Hunt, “Turbulence structure in thermal convection and shear-freeboundary layers,” J. Fluid Mech. 138, 161 �1984�.

5J. C. Wyngaard, “A physical mechanism for the asymmetry in top-downand bottom-up diffusion,” J. Atmos. Sci. 44, 1083 �1987�.

6J. C. R. Hunt, J. C. Kaimal, and J. I. Gaynor, “Eddy structure in theconvective boundary layer-new measurements and new concepts,” Q. J. R.Meteorol. Soc. 114, 837 �1988�.

7H. Schmidt and U. Schumann, “Coherent structure in the convectiveboundary layer derived from large-eddy simulations,” J. Fluid Mech. 200,511 �1989�.

8S. S. Zilitinkevich, Turbulent Penetrative Convection �Avebury Technical,Aldershot, 1991�, and references therein.

9A. G. Williams, H. Kraus, and J. M. Hacker, “Transport processes in thetropical warm pool boundary layer, Part I: Spectral composition of fluxes,”J. Fluid Mech. 53, 1187 �1996�.

126601-10 Elperin et al. Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions

Page 12: Large-scale instabilities in a nonrotating turbulent convection

10S. S. Zilitinkevich, A. Grachev, and J. C. R. Hunt, “Surface frictionalprocesses and non-local heat/mass transfer in the shear-free convectiveboundary layer,” in Buoyant Convection in Geophysical Flows, edited byE. J. Plate, E. E. Fedorovich, and D. X. Viegas �Kluwer Academic, Dor-drecht, 1998�, pp. 83–113.

11B. Brümmer, “Roll and cell convection in winter-time arctic cold-air out-breaks,” J. Atmos. Sci. 56, 2613 �1999�.

12G. S. Young, D. A. R. Kristovich, M. R. Hjelmfelt, and R. C. Foster,“Rolls, streets, waves and more,” BAMS, July 2002, ES54.

13R. Krishnamurti and L. N. Howard, “Large-scale flow generation in tur-bulent convection,” Proc. Natl. Acad. Sci. U.S.A. 78, 1981 �1981�.

14M. Sano, X. Z. Wu, and A. Libchaber, “Turbulence in helium-gas freeconvection,” Phys. Rev. A 40, 6421 �1989�.

15S. Ciliberto, S. Cioni, and C. Laroche, “Large-scale flow properties ofturbulent thermal convection,” Phys. Rev. E 54, R5901 �1996�.

16L. P. Kadanoff, “Turbulent heat flow: structures and scaling,” Phys. Today54, 34 �2001�.

17J. J. Niemela, L. Skrbek, K. R. Sreenivasan, and R. J. Donnelly, “Thewind in confined thermal convection,” J. Fluid Mech. 449, 169 �2001�.

18J. J. Niemela and K. R. Sreenivasan, “Rayleigh-number evolution of large-scale coherent motion in turbulent convection,” Europhys. Lett. 62, 829�2003�.

19U. Burr, W. Kinzelbach, and A. Tsinober, “Is the turbulent wind in con-vective flows driven by fluctuations?,” Phys. Fluids 15, 2313 �2003�.

20H. D. Xi, S. Lam, and X. Q. Xia, “From laminar plumes to organizedflows: the onset of large-scale circulation in turbulent thermal convection,”J. Fluid Mech. 503, 47 �2004�.

21X. D. Shang, X. L. Qiu, P. Tong, and X. Q. Xia, “Measurements of thelocal convective heat flux in turbulent Rayleigh-Benard convection,” Phys.Rev. E 70, 026308 �2004�.

22E. Brown, A. Nikolaenko, and G. Ahlers, “Orientation changes of thelarge-scale circulation in turbulent Rayleigh-Bénard convection,” Phys.Rev. Lett. 95, 084503 �2005�.

23A. Eidelman, T. Elperin, N. Kleeorin, A. Markovich, and I. Rogachevskii,“Hysteresis phenomenon in turbulent convection,” Exp. Fluids 40, 723�2006�.

24T. Hartlep, A. Tilgner, and F. H. Busse, “Large-scale structures inRayleigh-Benard convection at high Rayleigh numbers,” Phys. Rev. Lett.91, 064501 �2003�.

25A. Parodi, J. von Hardenberg, G. Passoni, A. Provenzale, and E. A.

Spiegel, “Clustering of plumes in turbulent convection,” Phys. Rev. Lett.92, 194503 �2004�.

26F. H. Busse and J. A. Whitehead, “Instabilities of convective rolls in ahigh Prandtl number fluid,” J. Fluid Mech. 47, 305 �1971�.

27F. H. Busse and J. A. Whitehead, “Oscillatory and collective instabilitiesrolls in large Reynolds number convection,” J. Fluid Mech. 66, 67 �1974�.

28F. H. Busse, “Generation of mean flows by thermal convection,” PhysicaD 9, 287 �1983�.

29D. E. Fitzjarrald, “An experimental study of turbulent convection in air,”J. Fluid Mech. 73, 693 �1976�.

30T. Elperin, N. Kleeorin, I. Rogachevskii, and S. S. Zilitinkevich, “Forma-tion of large-scale semiorganized structures in turbulent convection,”Phys. Rev. E 66, 066305 �2002�.

31T. Elperin, N. Kleeorin, I. Rogachevskii, and S. S. Zilitinkevich, “Tanglingturbulence and semi-organized structures in convective boundary layers,”Boundary-Layer Meteorol. 119, 449 �2006�.

32A. S. Monin and A. M. Yaglom, Statistical Fluid Mechanics �MIT Press,Cambridge, MA, 1975�, and references therein.

33J. L. Lumley, “Rational approach to relations between motions of differentscales in turbulent flows,” Phys. Fluids 10, 1405 �1967�.

34J. C. Wyngaard and O. R. Cote, “Cospectral similarity in the atmosphericsurface layer,” Q. J. R. Meteorol. Soc. 98, 590 �1972�.

35S. G. Saddoughi and S. V. Veeravalli, “Local isotropy in turbulent bound-ary layers at high Reynolds number,” J. Fluid Mech. 268, 333 �1994�.

36T. Ishihara, K. Yoshida, and Y. Kaneda, “Anisotropic velocity correlationspectrum at small scales in a homogeneous turbulent shear flow,” Phys.Rev. Lett. 88, 154501 �2002�.

37W. H. Reid and D. L. Harris, “Some further results on the Bénard Prob-lem,” Phys. Fluids 1, 102 �1958�.

38S. Chandrasekhar, Hydrodynamic and Hydromagnetic Stability �OxfordUniv. Press, Oxford, 1961�.

39P. G. Drazin, Introduction to Hydrodynamic Stability �Cambridge Univ.Press, Cambridge, 2002�.

40D. Etling, “Some aspects on helicity in atmospheric flows,” Contrib. At-mos. Phys. 58, 88 �1985�.

41I. Tuominen, A. Brandenburg, D. Moss, and M. Rieutord, “Does solardifferential rotation arise from a large scale instability,” Astron. Astro-phys. 284, 259 �1994�.

42F. Cattaneo, D. Lenz, and N. Weiss, “On the origin of the solar me-sogranulation,” Astrophys. J. 563, L91 �2001�.

126601-11 Large-scale instabilities in a nonrotating Phys. Fluids 18, 126601 �2006�

Downloaded 27 Aug 2012 to 128.248.155.225. Redistribution subject to AIP license or copyright; see http://pof.aip.org/about/rights_and_permissions


Recommended