Modeling return to isotropy using kinetic equations

0 downloads 0 Views 270KB Size Report
return-to-isotropy are investigated. Anisotropic turbulent decay is a parametrically simple but theoretically complex turbulent flow that is dominated by nonlinear ...
PHYSICS OF FLUIDS 17, 035101 共2005兲

Modeling return to isotropy using kinetic equations Blair Perota兲 and Chris Chartrand Department of Mechanical Engineering, University of Massachusetts, Amherst, Massachusetts 01003

共Received 7 July 2004; accepted 27 October 2004; published online 28 January 2005兲 Kinetic equations modeling the behavior of the velocity probability density function 共PDF兲 in homogeneous anisotropic decaying turbulence are hypothesized and their implications for return-to-isotropy are investigated. Anisotropic turbulent decay is a parametrically simple but theoretically complex turbulent flow that is dominated by nonlinear interactions. The physical implications of the Bhatnagar–Gross–Krook model, a relaxation model, and the Fokker–Planck model for the “collision” term in the PDF evolution equation are analyzed in detail. Using fairly general assumptions about the physics, three different parameter-free return-to-isotropy models are proposed. These models are compared with experimental data, classical models, and analytical limits. The final model expression is particularly interesting, and can easily be implemented in classic Reynolds stress transport models. © 2005 American Institute of Physics. 关DOI: 10.1063/1.1839153兴 I. INTRODUCTION

In most turbulent flows of interest the turbulent velocity fluctuations are anisotropic, that is, they differ in magnitude depending on their orientation. One aspect of Reynolds stress transport models 共and other more advanced models兲 that distinguishes them from simple two-equation transport models like k-␧ is their ability to more accurately model turbulence anisotropy. The degree of anisotropy is important because it can directly impact how turbulence affects the mean flow. In the absence of any driving mechanism, anisotropic turbulent flows tend to return to an isotropic state 共the state of least order兲. This nonlinear process is often called returnto-isotropy. It was identified early on in the development of Reynolds stress transport models and first modeled by Rotta.1 Since that time, the return-to-isotropy process has been extensively investigated and modeled.2–11 The return-to-isotropy problem is of significant theoretical interest in the theory of turbulence because it is entirely due to nonlinear interactions. In theory 共if the Navier–Stokes equations are solved兲, the return process is reversible. However, averaging processes 共such as the ensemble averages used in RANS models兲 lead to irreversibility. A similar effect happens in thermodynamics—molecular collisions are completely reversible, but their thermodynamic average behavior is not. This means that at the RANS modeling level, return to isotropy is an irreversible process and should be modeled as such. Existing models for return-to-isotropy tend to make extensive use of mathematical concepts, such as the Cayley– Hamilton theorem, realizability, Taylor series expansions, and fixed-point analysis. The resulting models invariably have at least one model “constant” that must be set via experiments. In this work, we are interested in deriving models for the return process 共or setting the unknown constants in existing models兲 based on physical ideas as well as mathematical a兲

Electronic mail: [email protected]

1070-6631/2005/17共3兲/035101/18/$22.50

tools. We make the assumption that turbulence behaves as a kinetic process, and that kinetic models of turbulence may lead to some useful insights about the return process. The advantage of this approach is that by assuming some very general physical conditions, the resulting models can be made to be free of any tunable model constants. In Sec. II, the classic Reynolds stress transport equation approach to modeling return-to-isotropy is briefly reviewed. We use these classic results as a reference since this is the approach that is most widely understood by most readers. In Sec. III we consider return-to-isotropy from the perspective of the Bhatnagar–Gross–Krook 共BGK兲12 approximation to the Boltzmann equation. Classic linear return models result from this kinetic equation. The deficiencies of the BGK approach are largely solved by two parameter-free relaxation collision models developed and tested in Sec. IV. Section V investigates the predictive performance of these models for five different experimental cases. The relaxation model is extended in Sec. VI to enable any desired Reynolds stress return behavior, and another parameter-free model is proposed that has some unique properties and better agreement with experimental data. Section VII explores the implications and connections to the Fokker–Planck collision model, and the results are discussed in Sec. VIII, where some speculation is presented as to what these kinetic models imply about turbulent eddy interactions. II. REYNOLDS STRESS TRANSPORT MODELS

In the absence of any mean flow the evolution of the Reynolds stress tensor Rij in homogeneous but anisotropic turbulence evolves according to the equation

⳵ Rij ⬘ u⬘j,k + p共u⬘i,j + u⬘j,i兲. = − 2v ui,k ⳵t

共1兲

The first term on the right-hand side is the dissipation rate tensor and the second term is the slow pressure strain. The pressure strain is considered “slow” in this situation because

17, 035101-1

© 2005 American Institute of Physics

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-2

Phys. Fluids 17, 035101 共2005兲

B. Perot and C. Chartrand

the pressure in this term depends only on the turbulence not on the “rapid” mean flow velocity gradients 共since there are none in this situation兲. Both terms require modeling. However, one half the trace of the dissipation tensor is the dissipation rate, ␧ = v ui,k ⬘ u⬘j,k, which is assumed to be known 共given by another transport equation model兲, and the trace of the pressure-strain term is zero in incompressible flows. The most common modeling approach is to assume that the dissipation tensor is close to isotropic. If small anisotropy in the dissipation tensor exists then it is included with the pressure strain model. The slow pressure strain and anisotropic dissipation are then collectively modeled as a “returnto-isotropy” term. There are reasons to suggest that modeling dissipation anisotropy and slow pressure strain separately is also advantageous,13,14 but for simplicity we retain the “collective” approach described above. The simplest model 共due to Rotta兲 is that the return-to-isotropy term is proportional to the Reynolds stress anisotropy. This gives a Reynolds stress transport model of the form





2 ⳵ Rij ˆ ␧ Rij − 2 ␦ . = − ␧␦ij − C R ij 3 ⳵t K 3

共2兲

The return-to-isotropy term will tend to drive the Reynolds stress tensor towards an isotropic state as time proceeds. The rate at which this happens is governed by the Rotta constant ˆ . This return model is the simplest possible one, and is C R linear in the Reynolds stress anisotropy, aij = 关共Rij / K兲 − 32 ␦ij兴. Equation 共2兲 appears to imply return-to-isotropy for any ˆ must be positive value of CˆR. In fact this is not the case, C R greater than 1. To see this we look at the evolution equation for Rij / K which should tend towards 32 ␦ij,





⳵ 共Rij/K兲 1 ⳵ Rij Rij ⳵ K ˆ − 1兲 ␧ Rij − 2 ␦ . 共3兲 − 2 = = − 共C R ij ⳵t K ⳵t K ⳵t K K 3 The isotropic dissipation actually causes the Reynolds stress tensor to move away from isotropy which must be counterˆ is actually a parameter, not a acted by the return term. C R strict constant, which can be 共and often is兲 a function of the Reynolds stress invariants and turbulent Reynolds number. Due to the strict requirement described above the splitting ˆ = 1 + C is useful. This gives a model equation of the form C R R





⳵ Rij Rij Rij 2 = − ␧ − C R␧ − ␦ij , K ⳵t K 3

⳵ Rij ␧ =− 共⌸imRmj + ⌸ jmRmi兲 , ⳵t 2K

where the dimensionless ⌸ij is some as yet unspecified ˆ gives model. Expanding this model as ⌸ij = ␦ij + ⌸ ij

⳵ Rij ␧ ␧ ˆ 共⌸ imRmj + ⌸ˆ jmRmi兲 , = − Rij − ⳵t K 2K

共6兲

ˆ is the return part of the model. The trace so it is clear that ⌸ ij of the last term should be zero, so we have a single constraint ˆ be symˆ R = 0. It is not necessary that ⌸ on the model, ⌸ ij ji ij metric. The explicit inclusion of the Reynolds stress in Eq. 共5兲 means that this general model can be strongly realizable ˆ is finite. If one component of 共Schumann,18 Lumley2兲 if ⌸ ij the turbulence goes to zero then this model will also make the time derivative of that component go to zero. However, ˆ becomes singular 共goes in the unusual circumstance that ⌸ ij to infinity兲 this model can potentially violate strong realizability. The classic linear return model described above is ˆ = C 共␦ − 2 KR−1兲. This model becomes singular given by ⌸ ij R ij 3 ij in the two-component limit 共because of the Reynolds stress inverse兲. The classic linear model satisfies weak realizability19 if CR ⬎ 0, but for the linear model to satisfy strong realizability CR must go to zero in the two-component limit. Slightly more complex nonlinear models for return-toisotropy have the general form





␦ij K ⳵ aij = − CR共aij兲 + CN aikakj − ankakn . 3 ␧ ⳵t

共7兲

Cubic and higher order nonlinear models can also be represented by this quadratically nonlinear model due to the Cayley-Hamilton theorem. Sarkar and Speziale4 suggest values of CR = 0.7 and CN = 1.05. The realizability conditions are clearer when this model is written in terms of the Reynolds stresses,

再 冋

⳵ Rij RnkRkn 4 ␧ = − Rij − CR − CN − 3 ⳵t K 2K2 ⫻

共4兲

where CR ⬎ 0. Typical values for CR lie between 0.5 and 1.0 共Durbin兲.15 Launder, Reece, and Rodi16 suggest a value of 0.8. No return to isotropy is the case of CR = 0. Physically, the no-return limit appears to occur at low Reynolds numbers. In addition, the no-return limit is often enforced in the twocomponent limit 共where one of the Reynolds stress diagonals goes to zero faster than the others, such as near walls兲. For this reason CR is often not a constant but is actually a parameter that depends on the turbulent Reynolds number and Reynolds stress invariants.7,17 It is helpful to propose a general model for the Reynolds stress evolution,

共5兲







册冎



2 ␧ ␧ RnkRkn Rij . Rij − K␦ij + CN 2 RikRkj − 3 K K 2K 共8兲

Pre- and postmultiplying this expression by the eigenvector matrix Q diagonalizes the Reynolds stress tensor 共QTRQ = D兲, so QT

再 冋

⳵R ␧ RnkRkn 4 Q = − D − CR − CN − 3 ⳵t K 2K2 ⫻







册冎



2 ␧ ␧ RnkRkn D , D − KI + CN 2 DD − 3 K K 2K 共9兲

since

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-3

QT

Phys. Fluids 17, 035101 共2005兲

Modeling return to isotropy



冊冉



⳵R ⳵D ⳵ QT ⳵ QT Q= +D Q − Q D ⳵t ⳵t ⳵t ⳵t

the off diagonal evolution is trivial, and the diagonal components individually satisfy the right-hand side of Eq. 共9兲. Weak realizability is satisfied as long as CR − CN关共RnkRkn / 2K2兲 − 34 兴 艌 0. Strong realizability requires equality on this previous expression and 1 + CN关RnkRkn / 2K2兴 艌 0. The quantity RnkRkn / 2K2 appears frequently and is related to the second invariant of the anisotropy tensor via RnkRkn / 2K2 = 32 + 21 ankakn. The model expression for the nonlinear return model is

再 冋

⌸ij = ␦ij + CR − CN − CN



RnkRkn 4 − 3 2K2



册冎冉

2 3

␦ij − KR−1 ij

Rij RnkRkn − ␦ij . K 2K2



共10兲

The singularity due to R−1 ij is weakly realizable as long as the leading coefficient is positive. It is strongly realizable if this leading coefficient is zero in the two-component limit and the coefficient of ␦ij is positive.

In homogeneous turbulence in the absence of mean accelerations or mean pressure gradients the evolution equation for the velocity probability density function 共PDF兲 is

冏 冏

.

共11兲

collisions

This equation governs the decay of anisotropic homogeneous turbulence, which is the focus of this work. One of the simplest collision models is a relaxation of the PDF to some known equilibrium form

⳵f ␧ = − CBGK共f − f eq兲, ⳵t K

共12兲

where the constant CBGK共x , t兲 might be a function of position and time but is not a function of the velocity. This model is similar to the BGK approximation for collisions used in Lattice-Boltzmann methods. It is also similar to the IEM models used in scalar mixing. In this particular context there are no theoretical justifications for this model 共such as an H theorem兲. As the simplest possible collision model it is informative to explore its attributes. The constant CBGK should always be greater than zero for a well-posed method. Unlike molecules, turbulence particles do not conserve kinetic energy when they collide, so the form of f eq, the equilibrium target distribution, must be slightly different from classical theory. If we take the target distribution to be ˆ 兲 = 共 4 ␲K ˆ 兲−3/2e−3vn⬘vn⬘/4Kˆ , f eq共K 3

⳵f ␧ ˆ 共f − 共 34 ␲Kˆ兲−3/2e−3vn⬘vn⬘/4K兲. =− ˆ ⳵t 共K − K兲

共14兲

This is a model in which the PDF relaxes towards a spherical Gaussian PDF with less turbulent kinetic energy 共see Fig. 1兲. Those portions of the PDF which are farthest from the target spherical distribution decay faster than those portions of the PDF which are closer to the target. The equivalent Reynolds stress transport equation is obtained by multiplying by vi⬘v⬘j and integrating over all velocities. This is shown in Appendix A, and results in the following equation:

⳵ Rij ␧ =− 共Rij − 32 Kˆ␦ij兲 ⳵t 共K − Kˆ兲

III. BHATNAGAR–GROSS–KROOK COLLISION MODELS

⳵f df = ⳵t dt

FIG. 1. BGK relaxation model. Solid line represents an isocontour for an anisotropic PDF. Dashed line is the spherical target distribution with less energy, which causes both dissipation and return-to-isotropy.

共13兲

ˆ ⬍ K, then 共as shown in Appendix A兲 mass and where 0 ⬍ K momentum are conserved and turbulent kinetic energy obeys ˆ 兲. This implies that the equation ⳵K / ⳵t = −共␧ / K兲CBGK共K − K ˆ / K兲, and the dissipating collision model is CBGK = 1 / 共1 − K

=−

1 ␧ ␧ Rij − 共Rij − 32 K␦ij兲. K K 共K/Kˆ − 1兲

共15兲

ˆ , this model is ⌸ ˆ = 共1 / 关K / 共Kˆ − 1兲兴兲共␦ In terms of ⌸ ij ij 兲 − 32 KR−1 , which is identical to the classic return model if ij CR = 1 / 关K / 共Kˆ − 1兲兴, or equivalently Kˆ = K关CR / 共1 + CR兲兴. This ˆ between the BGK reimplies the relation CBGK = 1 + CR = C R laxation constant and the Rotta constant. From this analysis it can be seen that there is no return to isotropy if CR = 0 共or Kˆ = 0兲. Under the condition Kˆ = 0, f eq becomes a delta function. This observation suggests an alternative model of the form

⳵f ␧ ␧ = − (f − ␦共v⬘兲) − CR(f − f eq共K兲). ⳵t k k

共16兲

The first term 共involving a delta function兲 produces pure decay and the second produces return to isotropy with no decay 共relaxation to a spherical PDF of the same energy兲. This two-part model has been proposed by Degond and Lemou.11 While both 共14兲 and 共16兲 result in an identical equation for the Reynolds stress evolution 共the classic linear Rotta model兲, the models themselves are not identical. Differences exist in the evolution of the higher turbulence moments. The model given by Eq. 共16兲 will tend to produce a spike in the PDF around its mean value 共due to the delta function兲. Equation 共14兲 has a smoother influence on the PDF in general but will also produce a spike if CR goes to zero 共in the twocomponent or low Reynolds number limits兲. Neither model has the ellipsoidal 关Eq. 共23兲兴 or spherical 关Eq. 共18兲兴 Gaussian as a solution. This implies that even if the turbulence starts with a Gaussian PDF it does not stay Gaussian. It is not a strict fact that turbulence should be Gaussian. Certainly under the influence of inhomogeneity we

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-4

Phys. Fluids 17, 035101 共2005兲

B. Perot and C. Chartrand

know it is not Gaussian at all. Even in homogeneous turbulence the tails of the PDF are not expected to be Gaussian. However, statistical arguments based on the central limit theorem would suggest that decaying homogeneous turbulence ought to be close to Gaussian or at least evolve in that direction for most of the core portion of the PDF. Experiments 共Tavoularis and Corrsin兲20 of homogeneous turbulent shear flows support the hypothesis that homogeneous turbulence 共even when sheared兲 has a central core that is closely approximated by an elliptical Gaussian PDF 共sometimes called a trivariate normal distribution兲. IV. RELAXATION COLLISION MODELS

A more general form than the BGK model 关Eq. 共12兲兴 for collisions is the linear relaxation model

⳵f = g共v兲 − h共v兲f , ⳵t

共17兲

where g共v兲 ⬎ 0 and h共v兲 ⬎ 0 are some positive functions of the velocity 共and possibly position and time as well兲. The positivity requirements keep the governing equation stable and the probability always greater than zero. In addition, the model should conserve the total probability 共or mass兲, so that 兰g共v兲dv = 兰h共v兲fdv, and it should not cause any mean flow to be created, implying 兰vn⬘关g − hf兴dv = 0. Finally the model should dissipate energy at the correct rate, 兰共vn⬘vn⬘ / 2兲关hf − g兴dv = ␧. One way to determine a suitable choice for the model functions is to insert a desired solution for the PDF function f and then derive the parameters from Eq. 共17兲. In isotropic decaying turbulence there is evidence that the core of the PDF is very close to a Gaussian and retains this shape during the decay process 共Yeung and Pope兲.21 If we assume the PDF equation 共17兲 has a Gaussian solution, f共v,t兲 =

共 34 ␲K兲−3/2e−3vn⬘vn⬘/4K ,

共18兲

where vn⬘ = vn − un and un is the mean velocity, then taking the time derivative gives

冉 冊

⳵f 4 = ␲K 3 ⳵t

−3/2



e−3vn⬘vn⬘/4K 1 −



vn⬘vn⬘ 3 ␧ . 2K 2 K

共19兲

Comparing with Eq. 共17兲 suggests that a suitable choice for the model functions is g共v兲 = f eq共v兲共3␧ / 2K兲 and h共v兲 = 共3␧ / 2K兲共vn⬘vn⬘ / 2K兲. Actually, these functions do not conserve momentum or dissipate energy at the correct rate. They must be generalized slightly to g共v兲 = CM

冉 冊

3␧ 4 ˆ ␲K 2K 3

−3/2

e−3vˆ n⬘vˆ n⬘/4K ,

˜vn⬘˜vn⬘ 3␧ , h共v兲 = C M 2K 关2K + 共u − ˜u兲2兴

ˆ

共20兲

ˆ → K, vˆ ⬘ → ˜v⬘ → v⬘ when the where we expect C M → 1, K i i i PDF approaches a spherical Gaussian 关Eq. 共18兲兴. Conservation of mass is already satisfied. Conservation of momentum implies a relationship exists between the hat and tilde velocities 共see Appendix B兲,

共uˆ p − u p兲关2K + 共u − ˜u兲2兴 = 2Rip共ui − ˜ui兲 +

冕 ⬘⬘⬘

v pvi vi fdv.

共21兲 This implies that either uˆ p or ˜u p can be specified arbitrarily and then the other determined by Eq. 共21兲. The two simplest choices are uˆ p = u p which implies ˜ui = ui + 共R−1 u p = u p which implies uˆi = ui ip / 2兲兰v⬘pv⬘ n vn⬘ fdv, and ˜ + 共1 / 2K兲兰v⬘i vn⬘vn⬘ fdv. In either case, if the PDF is symmetric then the odd order integral is zero and ˜u p = uˆ p = u p. Since by definition vi = uˆi + vˆ ⬘i = ˜ui + ˜v⬘i = ui + v⬘i , this also implies vˆ ⬘i = ˜v⬘i = v⬘i as well. Therefore the hat and tilde quantities in Eq. 共20兲 can be viewed as a small perturbation imposed when the PDF is skewed 共not symmetric兲, and are largely a formal technicality to enforce conservation of momentum. Conservation of energy imposes a relation between CM ˆ / K 共Appendix B兲, and K





1 1 2 Kˆ + 共uˆ − u兲2 + 关2K + 共u − ˜u兲2兴 K 2K CM 3 =

1 2K

冕 ⬘⬘⬘⬘

v pv pvi vi fdv +

共u p − ˜u p兲 K

冕 ⬘⬘⬘

v pvi vi fdv

+ 共u − ˜u兲2 .

共22兲

If f is symmetric this simplifies considerably to ˆ 1 2 1 K + = K C M 3 4K2

冕 ⬘⬘⬘⬘

v pv pvi vi fdv.

If f is an elliptic Gaussian given by f = 关共2␲兲3det共Rnm兲兴−1/2e−1/2Rnmvn⬘vm⬘ −1

共23兲

then the integral can be evaluated and is 4K2 + 2RnmRmn 共Appendix C兲. Then Kˆ / K = 1 + 共RnmRnm / 2K2兲 − 32 共1 / C M 兲 or perhaps even more informatively



3 RnmRmn Kˆ CM = 1 + − 2 2K2 K



−1

.

The relaxation model therefore has one free parameter 共either C M or Kˆ / K兲. Both of these parameters should go to 1 when the turbulence is isotropic 共i.e., when f is a spherical Gaussian兲. Since RnmRnm / 2K2 → 32 in isotropic turbulence, forcing one of these conditions is sufficient to guarantee the other. The derivation of the equivalent Reynolds stress equation is given in Appendix B. The result is that Eq. 共20兲 is equivalent to

⳵ Rij =− ⳵t



␧ 1+

RnmRmn 2K2

Kˆ − K





ˆ Rij RinRnj 2 K + − ␦ij 3K K K2



共24兲 if an elliptic Gaussian is assumed for the PDF shape. Equation 共24兲 in turn implies the return parameters

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-5

CR =

冋 冋

Phys. Fluids 17, 035101 共2005兲

Modeling return to isotropy

ˆ 4 RmnRnm K − 2 + 3 2K K ˆ RmnRnm K 1+ 2 − 2K K

册 册

and

CN =



−1 1+

RmnRnm Kˆ − 2K2 K



共25兲 or, in terms of C M , CR = 27 C M − 1

CN = − 23 C M .

and

共26兲

Note that this model and the other models derived in this work tend to imply that CN is less than zero. In contrast, the widely used nonlinear model of Sarkar and Speziale4 has a positive value for this constant. The implications of this difference are examined in detail in Sec. VIII. Various choices of C M are possible. The simple choice C M = 1 leads to CR = 25 , and CN = − 23 . These values produce a model which is very similar to the two models examined in detail below. The equally simple choice Kˆ / K = 1 implies CR =

7 2K2 2K2 − 1 and CN = − . 3 RnmRnm RnmRnm

共27兲

ˆ / K implies the “target” distribution has the This choice of K same energy as the PDF but a spherical shape. The performance of this model is shown in Sec. V and is referred to as Model-1. The realizability condition, CR − CN





2K2 RnkRkn 4 = , 2 − 3 2K RmnRnm

indicates that Model-1 is weakly realizable. In general the realizability condition for these relaxation models is CR − CN





ˆ RnkRkn 4 K − = 3 2K2 K

冒冉

1+



ˆ RnmRmn K − , 2K2 K

ˆ / K vanish in the two-component limit so choices where K will satisfy the strong realizability condition. The quantity 3 F = det共Rij兲 / 共 32 K兲 is 1 in isotropic turbulence and 0 in the two component limit. The choice Kˆ / K = F means that

冋 冋

册 册

4 RmnRnm − +F 3 2K2 CR = RmnRnm 1+ −F 2K2

and CN =



−1 . RmnRnm 1+ − F 2K2



共28兲 The other strong realizability condition 共CN 艌 −2K2 / RnmRmn when F = 0兲 is also satisfied by this model. Referred to as Model-F, the performance of this model is also shown in Sec. V. This model has a target distribution that has less energy, and in this sense it is similar to the simple BGK model of Sec. III. However, unlike the BGK model, this model has the spherical Gaussian as a solution, is strongly realizable and does not produce a spike in the PDF in the two-component limit. In addition, unlike the simple BGK model, the decay constant h now depends on the velocity v

and acts preferentially on the tails of the distribution, damping extreme events more strongly. While there is one freeparameter left in this model, Kˆ / K, it is far more restricted in its behavior than the arbitrary constant CBGK = 1 + CR = CˆR found in the simpler BGK type relation model. All possible ˆ / K that have been tried 共three so far兲 give very choices for K similar results for the actual model predictions, so this type of model is far less “tunable.” V. MODEL PERFORMANCE

In this section the performance of these models is compared with experimental data for return to isotropy. For each test case, we present both the Reynolds stresses as a function of time and the Reynolds stress anisotropy as a function of time. The anisotropy is the standard method for looking at return-to-isotropy, since it eliminates much of the dependence on the dissipation. However, due to the nondimensionalization with respect to K the anisotropy can cause errors in one turbulence component 共possibly even experimental errors兲 to appear as a general failure of the entire model. For this reason we retain the direct Reynolds stress decay plots as well. In all models the dissipation is determined from the model transport equation

⳵␧ ␧2 = − C␧2 . ⳵t K

共29兲

The value of C␧2 is taken to be 11/6, which is the high Reynolds number analytical solution for turbulence with a low wavenumber k2 spectrum.22 In most of the experiments the initial value of the dissipation is not known, and is obtained by attempting to match the K profile as well as possible. In each case, we have solved the Reynolds stress ordinary differential equation 共ODE兲 associated with the model, using fourth order Runge–Kutta and very small time steps. We have also solved the corresponding PDF relaxation models and obtained very similar results. However, there are further numerical issues associated with solving the PDF equations which we do not wish to address here, so we simply present the ODE results in this paper. Because C␧2 and the return process are believed to be Reynolds number dependent, we have selected only high Re number experiments for comparison and no direct numerical simulation 共DNS兲 test cases. It must be noted that there is some uncertainty associated with the experimental results. First, while the geometry of these experiments changes abruptly from a straining section to a straight section, the actual cessation of the mean strain may not be quite so abrupt due to the long range effects of pressure. As a result, these decay experiments may have some residual straining in them at early times. The translation of the zero time in the Le Penven experiment, case III⬍0, suggests that the experimenters were aware of this problem. More importantly, the initial turbulence for these experiments has structure, due to the strains imposed to make the turbulence anisotropic. It is likely that at early times the relaxation of these structures also affects the return process.

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-6

B. Perot and C. Chartrand

Phys. Fluids 17, 035101 共2005兲

FIG. 2. Reynolds stresses and anisotropy for the case III⬎ 0 from Le Penven, Gence, and Comte-Bellot. Symbols are the experimental data, lines are the Model-1 predictions, and the dashed lines are the Model-F predictions.

FIG. 3. Reynolds stresses and anisotropy for case III ⬍ 0 from Le Penven, Gence, and Comte-Bellot. Symbols are the experimental data, lines are the Model-1 predictions, and the dashed lines are the Model-F predictions.

FIG. 4. Reynolds stresses and anisotropy for case A of Choi and Lumley. Symbols are the experimental data, lines are the Model-1 predictions, and the dashed lines are the Model-F predictions.

FIG. 5. Reynolds stresses and anisotropy for case B of Choi and Lumley. Symbols are the experimental data, lines are the Model-1 predictions, and the dashed lines are the Model-F predictions.

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-7

Phys. Fluids 17, 035101 共2005兲

Modeling return to isotropy

FIG. 6. Reynolds stresses and anisotropy for case C-2 of Choi and Lumley. Symbols are the experimental data, lines are the Model-1 predictions, and the dashed lines are the Model-F predictions.

Figures 2 and 3 are Le Penven et al. cases III ⬎0 共expansion兲 and III ⬍0 共contraction兲. Figures 4–6 are the data of Choi and Lumley7 for their cases A 共plane distortion兲, B 共axisymmetric expansion兲, and C-2 共axisymmetric contraction兲, respectively. Despite the fact that model-F is strongly realizable and Model-1 is not, the two models behave very similarly for all five experimental test cases. With the exception of Fig. 3 共Le Penven, case III ⬍0兲 and Fig. 6 共Choi and Lumley, case C-2兲 the models show poor agreement with the experimental data, and tend to return to isotropy too quickly. VI. GENERAL RELAXATION MODELS

Rather than assuming a spherical Gaussian, let us assume that the anisotropic ellipsoidal Gaussian 关Eq. 共23兲兴 is a solution to the relaxation equation 关Eq. 共17兲兴. It will ultimately be seen that this gives much better model predictions. With this very broad assumption,





−1 ⳵f 1 ⳵ /⳵ tdet共Rnm兲 ⳵ Rnm =− f + v ⬘v ⬘ . 2 ⳵t det共Rnm兲 ⳵t n m

共30兲

−1 −1 and Since ⳵R−1 ij / ⳵t = −Rim 共⳵Rmn / ⳵t兲Rnj −1 = det共Rnm兲共⳵Rip / ⳵t兲R pi 共Jacobi’s formula兲 this

⳵ / ⳵t det共Rnm兲 reduces to

⳵f ⳵R 1 −1 −1 ⬘ 兲 ij . = − f共R−1 ij − Rim Rnj v⬘ nvm 2 ⳵t ⳵t

共31兲

Let us further assume that ⳵Rij / ⳵t = −␧ / 2K共⌸imRmj + ⌸ jmRmi兲 which is the general Reynolds stress transport model 关Eq. 共5兲兴. Then

⳵f ␧ −1 ⬘ vn⬘兲 . = f 共⌸ii − ⌸inRim vm ⳵ t 2K

共32兲

This implies that for any desired Reynolds stress model ⌸ij a corresponding relaxation model can be constructed, ␧ e−1/2Rnmvˆ ⬘nvˆ m⬘ g共v兲 = C M ⌸ii , 2K 关共2␲兲3det共Rˆ 兲兴1/2 ˆ −1

␧ 2K



共33兲

−1 ˜vm ⬘ ˜vn⬘ ⌸inRim −1 1 + 共un − ˜un兲共um − ˜um兲Rim



−1 共uˆ p − u p兲 1 + 共un − ˜un兲共um − ˜um兲Rim

=

⌸in −1 R ⌸ii im

⌸in ⌸ pp



.

When the PDF is an elliptic Gaussian we expect C M = 1, ˜vn⬘ = vˆ n⬘ = vn⬘, and Rˆnm = Rnm. The constant C M can be a function

再冕

⌸in ⌸ pp



⬘ v⬘n fdv + Rnp共um − ˜um兲 v⬘pvm



+ Rmp共un − ˜un兲 .

共34兲

The simplest choice is ˜u p = u p then uˆ p = u p +

⌸in −1 R ⌸ii im

冕 ⬘⬘⬘

v pvmvn fdv.

共35兲

The choice uˆ p = u p is more complicated and requires a sym−1 ˜ n − u n兲 metric matrix inversion 共⌸ipR−1 in + ⌸inRip 兲共u −1 −1 = ⌸inRim Rtp 兰v⬘t vm ⬘ vn⬘ fdv. For certain models 共like the one shown below兲, this matrix problem is easy to invert analytically, and this choice is also viable. The Reynolds stress transport equation is derived in Appendix E. Assuming the choice ˜u p = u p it requires that ⌸ pn −1 Rˆij = R ⌸ss pm −

nm

h共v兲 = C M

of tilde and hat quantities 共such as ˜u p − u p and uˆ p − u p兲 but it can no longer be a function of the Reynolds stress invariants 共like it was in the simpler spherical relaxation model兲. This is because the elliptic Gaussian PDF 共unlike the spherical Gaussian PDF兲 can represent any state of the Reynolds stress invariants. Note that the relaxation equation places constraints on the underlying Reynolds stress model. It implies that ⌸ii −1 ⬎ 0, and ⌸inRim must be a positive definite tensor. Conservation of probability 共or mass兲 is already satisfied by this model. Conservation of momentum requires a relation between uˆ p and ˜u p 共see Appendix D兲,



冕 ⬘⬘⬘⬘

vmvnvi v j fdv



1 ⌸im ⌸ jm Rmj + Rmi − 共uˆi − ui兲共uˆ j − u j兲. 共36兲 C M ⌸ss ⌸ss

If the PDF is an ellipsoidal Gaussian then uˆ p = u p 关by Eq. 共35兲兴, and C M = 1 共by definition兲. In addition, since

冕 ⬘⬘⬘⬘

vmvnvi v j fdv = RmnRij + RmiRnj + RmjRni

共37兲

Eq. 共36兲 gives the correct limit, Rˆij = Rij for an elliptic Gaussian PDF. The hat and tilde quantities can be seen to be slight

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-8

Phys. Fluids 17, 035101 共2005兲

B. Perot and C. Chartrand

perturbations to the standard quantities that precisely account for any deviation of the PDF from an elliptic Gaussian shape. The model given by Eqs. 共33兲–共36兲 represents the general relaxation model. Using this formulation, any Reynolds stress transport model can also be implemented as a PDF relaxation model, which has the elliptic Gaussian as a solution. Remember that ⌸ij = ␦ij corresponds to the case of no 兲 return-to-isotropy, and ⌸ij = ␦ij + CR共␦ij − 32 KR−1 ij is the classic linear return-to-isotropy model. Substituting these expressions into Eqs. 共33兲–共36兲 will produce the corresponding PDF relaxation model. However, in this paper, we do not wish to specify ⌸ij, but to determine what the general relaxation model 关Eqs. 共33兲–共36兲兴 imply about how it should be specified. The general relaxation model as described above has singular h, uˆ p, and Rˆij in the two-component limit due to the presence of R−1 ij . This singularity is removed by the parameter-free Reynolds stress model ⌸ij = 共2K / RnmRmn兲Rij. Making ⌸ij directly proportional to the Reynolds stress tensor removes the singularities. The constant of proportionality is determined from the decay condition ⌸ijRij = 2K 共see Sec. III兲. In the relaxation context this model is given by ␧ 2K2 e−1/2Rnmvˆ ⬘nvˆ m⬘ , g = CM K RnmRmn 关共2␲兲3det共Rˆ 兲兴1/2 ˆ −1

nm

共38兲

˜vn⬘˜vn⬘ ␧ 2K2 , h = CM K RnmRmn 关2K + 共u − ˜u兲2兴 where 共uˆ p − u p兲关2K + 共u − ˜u兲2兴 = 2Rip共ui − ˜ui兲 +

冕 ⬘⬘⬘

v pvi vi fdv.

共39兲 Note that Eq. 共39兲 is a particular case of the general Eq. 共34兲 共for this ⌸ij model兲. It also happens to be identical to Eq. 共21兲, the general expression for the spherical relaxation models in Sec. IV. As in Sec. IV, the choice of uˆ p = u p or ˜u p = u p is up to the user. For symmetric PDFs it makes no difference what the choice is, since then uˆ p = ˜u p = u p. For inhomogeneous flows, the PDFs will be skewed and this choice may make some difference. For this model we also require the condition on Rˆij that





1 RimRmj Rˆij + 共uˆi − ui兲共uˆ j − u j兲 + 关2K + 共u − u兲2兴 CM K =

冕 ⬘⬘⬘⬘

vnvnvi v j fdv + 共un − un兲

冕 ⬘⬘⬘

vnvi v j fdv + Rij共u − u兲2 .

共40兲 This model 关Eqs. 共38兲–共40兲兴 differs from those in Sec. IV, in that it has the ellipsoidal Gaussian as a solution. The choices for C M are now far more restrictive. The simplest choice is simply to set C M = 1. Equations 共40兲 and 共39兲 are simplified considerably by choosing

˜u p = u p. Then the hat quantities are defined by uˆ p = u p + 共1 / 2K兲兰v⬘pv⬘i v⬘i fdv and Rˆij = 共1 / 2K兲兰v⬘nvn⬘vi⬘v⬘j fdv − 共uˆi − ui兲共uˆ j − u j兲 − 共1 / C M 兲共RimRmj / K兲. The equivalent Reynolds stress transport model can be derived from this relaxation model by assuming the PDF is an elliptic Gaussian. Under this assumption, the various possible choices of the hat and tilde quantities are irrelevant and we find that all these choices are equivalent to

⳵ Rij RisRsj = − 2␧ , RmnRnm ⳵t

共41兲

which implies the model parameters are CR =

4 2K2 2K2 − 1 and CN = − . 3 RmnRnm RmnRnm

共42兲

We note that this model satisfies the strong realizability constraint, CR − CN关RnkRkn / 共2K2兲 − 34 兴 = 0, and sits on the cusp of the strong realizability condition CN 艌 −2K2 / RmnRnm. In the two-component limit, this model returns to isotropy as slowly as physically possible. The performance of this model is shown below and it is referred to as Model-EG 共for elliptic Gaussian兲. The fact that the resulting Reynolds stress model is very simple, entirely nonlinear, contains no model parameters, and satisfies strong realizability at its cusp, makes Model-EG very intriguing. In Figs. 7–11 the performance of Model-EG is compared with experimental data for return-to-isotropy, the classic linear Rotta model 共with CR = 0.8兲, and the nonlinear model of Sarkar and Speziale 共CR = 0.7, CN = 1.05兲. The most interesting result is that these three very different models perform very similarly for all five test cases. The Sarkar and Speziale model is slightly better than the other two, but it has two adjustable model constants that were tuned to exactly these test cases. The linear Rotta model also performs surprisingly well. It can be made even better by adjusting the standard value 共0.8兲 downwards 共to 0.7 or 0.6兲. Model-EG matches the data the least well, but gives quite good agreement considering there are no adjustable parameters in this model. As noted earlier, the greatest uncertainty in both the models and the experiments lies in the initial conditions. To see that the assessment of the models performance is not affected by these initial conditions, Fig. 8 was recalculated using a later time for initialization. Figure 12 shows that the point of initialization does not fundamentally change the results. We conclude this section by noting that other return models have been proposed that are nonlinear, which parameterize CR and CN as functions of the Reynolds stress invariants 共or anisotropy invariants兲, and which satisfy strong realizability.7,17 However, these models assume that CR and CN are polynomial functions of the invariants. In contrast, the model described above uses linear rational polynomial functions of the invariants to represent the return parameters CR and CN. We note that rational polynomials tend to have better fitting properties than polynomials, and that the formulated rational polynomials are the result of physical assumptions but not assumptions about functional behavior.

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-9

Modeling return to isotropy

Phys. Fluids 17, 035101 共2005兲

FIG. 7. Reynolds stresses and anisotropy for the case III⬎ 0 from Le Penven, Gence, and Comte-Bellot. Symbols are the experimental data, lines are the Rotta model predictions 共CR = 0.8兲, the dashed lines are the SS model predictions, and large dashed lines are the Model-EG predictions.

FIG. 8. Reynolds stresses and anisotropy for case III ⬍ 0 from Le Penven, Gence, and Comte-Bellot. Symbols are the experimental data, lines are the Rotta model predictions 共CR = 0.8兲, the dashed lines are the SS model predictions, and large dashed lines are the Model-EG predictions.

FIG. 9. Reynolds stresses and anisotropy for case A of Choi and Lumley. Symbols are the experimental data, lines are the Rotta model predictions 共CR = 0.8兲, the dashed lines are the SS model predictions, and large dashed lines are the Model-EG predictions.

FIG. 10. Reynolds stresses and anisotropy for case B of Choi and Lumley. Symbols are the experimental data, lines are the Rotta model predictions 共CR = 0.8兲, the dashed lines are the SS model predictions, and large dashed lines are the Model-EG predictions.

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-10

Phys. Fluids 17, 035101 共2005兲

B. Perot and C. Chartrand

FIG. 11. Reynolds stresses and anisotropy for case C-2 of Choi and Lumley. Symbols are the experimental data, lines are the Rotta model predictions 共CR = 0.8兲, the dashed lines are the SS model predictions, and large dashed lines are the Model-EG predictions.

VII. FOKKER–PLANCK COLLISION MODELS

An alternative to relaxation models is the Fokker–Planck collision model. This model is frequently used to model Brownian motion, liquid collisions, and some plasmas. Langevin models for turbulence19,23 are directly related to the Fokker–Planck equation and therefore effectively use this type of model. A generalized Fokker–Planck collision operator involves two as yet unspecified matrices, Gij and Hij, −

冋 册

⳵ 共Gijv⬘j f兲 ⳵f ⳵f ⳵ =− + Hij . ⳵ Vj ⳵t ⳵ Vi ⳵ Vi

共43兲

The tensor Hij should be positive definite for stability reasons. In Langevin models it is convenient to make Hij isotropic as well. However, in general the Fokker–Planck collision model has considerable flexibility in the choice of both the model tensors. The model automatically satisfies conservation of probability and momentum. It also has the ellipsoidal Gaussian as a solution.24 Multiplying Eq. 共43兲 by v⬘nvm ⬘ and integrating over all velocity space gives the equivalent Reynolds stress transport equation

⳵ Rnm = GmjR jn + GnjR jm + Hmn + Hnm . ⳵t

共44兲

By comparing this with the generic Reynolds stress transport equation 关Eq. 共5兲兴, it can be seen that −1 =− Gij + HimRmj

␧ ⌸ij . 2K

共45兲

In this way, classic return models 共given in terms of 兿ij兲 can be implemented in the generalized Fokker–Planck con-

text. This transformation is also discussed in Pope.19 The general nonlinear Reynolds stress return model 关Eq. 共8兲兴 is equivalent to

再 冋

⌸ij = ␦ij + CR + CN − CN



4 RnkRkn − 3 2K2



册冎

共␦ij − 32 KR−1 ij 兲

Rij RnkRkn − ␦ij . K 2K2

共46兲

In the Fokker–Planck context this implies that −1 =− Gij + HimRmj

冋 再



4 R R ␧ ␦ij + CR + CN − nk 2kn 3 2K 2K

⫻共␦ij − 32 KR−1 ij 兲 − CN



册冎

Rij RnkRkn − ␦ij K 2K2

冊册

. 共47兲

There are many possible choices of Gij and Hij which satisfy this constraint. The simplest and most numerically attractive choice for Hij is that this tensor is isotropic, Hij = CD␧␦ij, where CD is an arbitrary model constant. This means that Gij = − +



4 Rij ␧ 共1 + CR + CN兲␦ij − CN 3 K 2K







册 冊

4 RnkRkn ␧ − CR + CN − 3CD R−1 ij . 3 3 2K2

共48兲

The singularity in Gij is removed by the particular choice 3CD = CR + CN关 34 − 共RnkRkn / 2K2兲兴, which is the choice used in

FIG. 12. Reynolds stresses and anisotropy for the case III⬍ 0 from Le Penven, Gence, and Comte-Bellot, initialized at 0.037. Symbols are the experimental data, lines are the Rotta model predictions 共CR = 0.8兲, the dotted lines are the SS model predictions, and dashed lines are the Model-EG predictions.

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-11

Phys. Fluids 17, 035101 共2005兲

Modeling return to isotropy

most Langevin turbulence models. This gives the following model constants:



Hij = CR + CN



4 RnkRkn − 3 2K2

冊册





4 ␧ Rij 共1 + CR兲␦ij + CN ␦ij − 3 2K K

冊册

共49兲

.

Note that with this choice the weak realizability constraint CR + CN关 34 − 共RnkRkn / 2K2兲兴 艌 0 is equivalent to the requirement that Hij be positive definite. Under these circumstances, the classic linear return model 共with CN = 0兲 is and Him obtained using Gij = −共␧ / 2K兲共1 + CR兲␦ij = 共␧ / 3兲CR␦im. Model-1 given in Sec. IV 关with CR = 37 共2K2 / RnkRkn兲 − 1 and CN = −2K2 / RnkRkn兴 is obtained using







2K2 ␧ 2K2 ␧ ␦ij and Gij = − Hij = RnkRkn 3 2K RnkRkn

冊冋



册 册

4 RmnRnm − +F 3 2K2 CR = RmnRnm 1+ −F 2K2

冒冋

1+

and Gij = −



␧ R ␦ij + ij 2K K

and CN =



−1 , RmnRnm 1+ − F 2K2



RmnRnm −F 2K2

册冒冋

1+

册 册

RmnRnm −F . 2K2

4 2K2 2K2 − 1 and CN = − , 3 RmnRnm RmnRnm

is obtained using Hij = 0 and Gij = −



CR =

4 2K2 −1 3 RnkRkn

and

CN = −

2K2 , RnkRkn

␧ ␧ Rij − CE␦ij RnkRkn K

and

Hij =

␧ CERij , K

where CE is again an arbitrary parameter. Note that CE can actually be determined by a dispersion analysis and is related to the Kolmorgorov constant.

Note that in the two-component limit Hij now goes to zero. This particular splitting 关Eq. 共7.7兲兴 will be unstable in this limit. Model-EG, with CR =

where CS = CR + CN共 34 − RnkRkn / 2K2兲. Again, to remove the near singularity CD = 31 CS can be chosen, but because of the more general form for Hij the 共realizability兲 restriction CS 艌 0 is no longer required for a well-posed model. The classic linear return model is obtained using Gij = −共␧ / 2K兲共1 + CR + 2CE兲␦ij and Him = 共␧ / 3兲CR␦im + CE共␧ / K兲Rij. Note that this splitting has an extra free parameter CE which does not change the Reynolds stress evolution, but does change the model. A nonsingular splitting for Model-EG, with

Gij = −

is obtained using ␧ Hij = ␦ijF 3

共50兲

is now given by

R ␦ij + ij . K

Model-F, with

冋 冋

4 ␧ 共1 + CR + CN + 2CE兲␦ij 3 2K

␧ ␧ + 共CS − 3CD兲 R−1 + CN 2 Rij , 2K 3 ij

␧ ␦ij 3

and Gij = −

Gij = −



2K2 Rij ␧ . 2K RmnRnm K

This model is therefore incompatible with this splitting 共unstable兲. If Hij is assumed to be isotropic 共and nonzero兲, then Gij must become singular in the two-component limit. A more general splitting is possible if Hij is allowed to be anisotropic. Classic Langevin models require isotropic Hij, but the Fokker–Planck model itself only requires Hij to be positive definite. Assuming a positive definite form, Hij = CD␧␦ij + CE共␧ / K兲Rij implies

VIII. DISCUSSION

The return-to-isotropy problem of anisotropic turbulence has been studied via three very different collision models for the evolution of the velocity PDF. The simplest collision operator is the BGK approximation to the Boltzmann collision integral. This collision model, 共−␧ / k兲CBGK共f − f eq兲, is characterized by an inverse time scale 共which does not depend on the velocity兲. It was shown that if this model is to dissipate energy correctly, the target state must have considerably less energy than the current PDF state. Some models even use a target state with zero energy 共a delta function兲. The BGK model produces the classic linear return-to-isotropy model, with the rate of return CR = 1 / 共K / Kˆ − 1兲 determined by the energy of the target state. The Gaussian PDF is not a solution of the BGK model even though theoretical and experimental evidence might suggest that this is desirable. To overcome the limitations of the BGK model, more general relaxation models were constructed in which the collision operator g共v兲 − h共v兲f has a positive-definite velocity dependent source term and a velocity dependent sink term that is proportional to the PDF. Previous analysis of this collision model in the context of turbulence is unknown to the authors. In Sec. IV prescriptions for the model parameters g and h were derived such that the spherical Gaussian is a solution to the evolution equation. Two models were derived from this analysis, Model-1 assumed that the target distribution has the same energy as the PDF, Kˆ / K = 1. It is only weakly realizable. Model-F assumed that the target disˆ / K = F. tribution has less energy than the PDF in the ratio K This ratio was chosen because it makes the resulting model strongly realizable. While these initial parameter-free relax-

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-12

Phys. Fluids 17, 035101 共2005兲

B. Perot and C. Chartrand

FIG. 13. Invariant triangle. 共a兲 Sarkar and Speziale, 共b兲 Model-1, 共c兲 Model-F, 共d兲 Model-EG.

ation models did not perform as well as might be hoped, they set the stage for the development of the parameter-free Model-EG. Model-EG was shown to be the only nonsingular relaxation model that has the elliptic Gaussian as a solution. The equivalent Reynolds stress transport model is totally nonlinear in the Reynolds stresses and was shown to be strongly realizable. Interestingly, the performance of Model-EG is quite similar to the linear return to isotropy model. Even the Sarkar and Speziale model with an opposite sign for the nonlinear term, CN, performs similarly for the test cases studied. To investigate these models further, their trajectories on the anisotropy invariant map were plotted, and are presented in Fig. 13. It is well known that the linear Rotta model has linear trajectories when plotted on this anisotropy invariant map. The trajectories of the model of Sarkar and Speziale tend to move downwards and from left to right on this map. This means that turbulence with two large Reynolds stresses and one small stress will tend to first approach a state with only one large stress before approaching full isotropy. This implies that the intermediate stress decays faster than the maximum and minimum stresses, which is somewhat counter intuitive. The models developed in this paper tend to have the opposite behavior. Turbulence with one large stress will first decay to a state with two large stresses before approaching total isotropy. There is no experimental data in the middle of the triangle that allows us to determine which behavior is actually correct. The top boundary of the “triangle” is the two-component line. The strongly realizable models have trajectories that stay on this line and move to the left if they start on the two-component line. This means that if one component of the turbulence is zero it stays zero for all time, and the two nonzero stresses approach each other 共mutual isotropy兲. This is the expected behavior for two-dimensional turbulence,

which is sometimes 共but by no means always兲 found when the turbulence is two component. More information about the turbulence 共than the Reynolds stress兲 is clearly necessary to make return models behave correctly in the twocomponent limit. Strong realizability seems appropriate when the two-component turbulence is also two dimensional, and weak realizability seems appropriate otherwise. Finally, the relationship between the relaxation models and the Fokker–Planck collision model





⳵ 共Gijv⬘j f兲 ⳵f ⳵ + Hij ⳵ Vi ⳵ Vi ⳵ V j



was investigated. Like the general relaxation model 共Sec. VI兲 the Fokker–Planck model has the ellipsoidal Gaussian as a solution. Because it involves derivatives in velocity space, the Fokker–Planck model is more difficult to implement numerically than relaxation models. However, the Fokker– Planck model 共with isotropic Hij兲 has a direct correspondence with the Langevin PDF models. Examination of Model-EG in this context showed that this model cannot be implemented with isotropic Hij. Instead, the diffusion coefficient Hij must be proportional to Rij. In this work, fairly reasonable assumptions have been explored for how large collections of interacting dissipative particles 共turbulent eddies兲 might be expected to behave. We have then explored the modeling implications of these assumptions, and then tested against experimental data and theoretical analysis 共realizability considerations兲 to determine which assumptions are the most reasonable. The models are therefore telling us information about the physics. The primary assumption has been that turbulence tends toward a spherical or an elliptic Gaussian PDF distribution. While it is not entirely clear that high Reynolds number decaying anisotropic turbulence should become Gaussian 共spherical or elliptic兲, these are certainly very reasonable and

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-13

Phys. Fluids 17, 035101 共2005兲

Modeling return to isotropy

relatively broad starting points. The evidence from this paper indicates that turbulence does not try to approach a spherical Gaussian 共except in the limit as time goes to infinity兲. However we have found that models that assume that the ellipsoidal Gaussian distribution is a solution appear to have very interesting predictive and theoretical properties. In particular, we have determined the unique nonsingular 共assuming nature abhors a singularity兲 model of this form. The equivalent Reynolds stress transport model 关Eq. 共41兲兴 has not been proposed in the past but does have a number of attractive predictive and theoretical properties. In particular, and probably most importantly for many modelers, this model 共in its transport equation form兲 has absolutely no tunable constants. Less useful, but perhaps just as interesting, this model cannot be implemented as a classical Lagrangian PDF method 共Langevin equation兲.

⳵ ⳵t



vi f ⳵ v

=−



␧ 共K − Kˆ兲

冕⬘

再冕

vi f ⳵ v − ui



共 34 ␲Kˆ兲−3/2e3vn⬘v⬘n/4K ⳵ v



4 ˆ −3/2 3v⬘v⬘/4K 兲 e n n ⳵v . vi 共 3 ␲K

By definition 兰vi f ⳵ v = ui. The second integral on the righthand side is equal to 1 and the last integral is zero since it has an odd integrand, so finally

⳵ ␧ 兵ui − ui − 0其 = 0. ui = − ⳵t 共K − Kˆ兲

共A5兲

The Reynolds transport equation is obtained by multiplying the PDF relation equation by vi⬘v⬘j and then integrating over all velocity space,

ACKNOWLEDGMENTS

The authors gratefully acknowledge the financial support of this work by the Office of Naval Research 共Grant Nos. N00014-01-1-0483 and N00014-04-1-0267兲 under the supervision of Dr. Ronald Joslin.

冕 ⬘⬘ vi v j

=

⳵f ⳵v ⳵t

⳵ Rij ␧ =− ⳵t 共K − Kˆ兲

冕 ⬘ ⬘共



4 ˆ −3/2 3v⬘v⬘/4K 兲 e nn ⳵ v. vi v j f − 共 3 ␲K

APPENDIX A: BHATNAGAR–GROSS–KROOK MOMENTS

共A6兲

Conservation of mass 共or probability兲 requires that the integral of the PDF be equivalent to one for all time. This means that the integral of its time derivative must be zero. Starting from the BGK model for the time derivative of distribution gives the following expression:



⳵f ␧ ⳵v=− ˆ兲 ⳵t 共K − K

冕共



共A2兲

Conservation of momentum requires that no mean flow be created by the relaxation process. The mean velocity of the flow is equivalent to the first moment of the PDF. By taking the integral over all velocity space, we can show that



vi

⳵f ␧ ⳵v=− ⳵t ˆ兲 共K − K



⳵ ⳵t

4 ˆ −3/2 3v⬘v⬘/4K 兲 e n n 兲 ⳵ v. vi共f − 共 3 ␲K

共A3兲 Using the fact that the velocity is an independent variable 共from time兲 and splitting the velocity into its mean and fluctuating parts gives

冕 ⬘⬘

vi v j f ⳵ v

=−



Since both distributions integrate to 1, we can see that

f ⳵ v = 0.

This then becomes

ˆ 兲−3/2e3vn⬘vn⬘/4K ⳵ v . f − 共 34 ␲K 共A1兲

⳵ ⳵t

共A4兲

␧ 共K − Kˆ兲

再冕

− 共 34 ␲Kˆ兲−3/2

v⬘i v⬘j f ⳵ v

冕 ⬘⬘

vi v j e

3vn⬘vn⬘ 4K



⳵v .

共A7兲

Since 兰v⬘i v⬘j f ⳵ v = Rij by definition, and the last integral must be isotropic

⳵ Rij ␧ 兵Rij − 32 Kˆ␦ij其. =− ⳵t 共K − Kˆ兲

共A8兲

APPENDIX B: RELAXATION MODEL MOMENTS

Here we verify conservation of mass for the relaxation models derived in Sec. IV. The method is the same as before starting from the relaxation model for the PDF,



⳵f ⳵v= ⳵t

冕 冉 CM



3␧ 4 ˆ −3/2 −3vˆ ⬘vˆ ⬘/4Kˆ 共 ␲K兲 e n n 2K 3



˜v⬘n˜v⬘n 3␧ f ⳵v 2K 2K + 共u − ˜u兲2

共B1兲

with vˆ i⬘ = vi − uˆi, ˜vi⬘ = vi − ˜ui,

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-14



Phys. Fluids 17, 035101 共2005兲

B. Perot and C. Chartrand

冕 ⬘⬘

3␧ 3␧ ⳵f − CM ⳵ v = CM 2K 2K ⳵t ⫻



vi v j

⳵f ⳵ Rij ⳵v= ⳵t ⳵t

共vn⬘ + un − ˜un兲共vn⬘ + un − ˜un兲 f ⳵ v = 0. 2K + 共u − ˜u兲2

=

共B2兲 Since 兰v⬘n f ⳵ v = 0 we get



冕关

3␧ ⳵ Rij = CM 2K ⳵t

共un − ˜un兲共un − ˜un兲 + v⬘nv⬘n兴 f ⳵ v = 0.

The integrals can be evaluated to give



vi

3␧ 2K = 0. 2K 2K + 共u − ˜u兲2

⳵f ⳵ ui = ⳵v= ⳵t ⳵t

冕 冉 C M vi



共B8兲



关vˆ ⬘i + 共uˆi − ui兲兴 关vˆ ⬘j + 共uˆ j − u j兲兴

冕 ⬘ ⬘„

vi v j 共un − ˜un兲共un − ˜un兲 + 2v⬘n共un − ˜un兲

共B4兲

3␧ ⳵ Rij = CM 2K ⳵t





关vˆ ⬘i vˆ ⬘j + 共uˆi − ui兲共uˆ j − u j兲兴

3␧ 1 ˆ ⫻共 34 ␲Kˆ兲−3/2e−3vˆ n⬘vˆ n⬘/4K ⳵ v − C M 2K 2K + 共u − ˜u兲2

3␧ 4 ˆ −3/2 −3vˆ ⬘vˆ ⬘/4Kˆ 共 ␲K兲 e n n 2K 3

˜vn⬘˜vn⬘ 3␧ − f ⳵v 2K 2K + 共u − ˜u兲2

共B9兲

−3/2 ˆ Since 兰vˆ ⬘i 共 34 ␲Kˆ兲 e−3vˆ n⬘vˆ n⬘/4K ⳵ v = 0 共due to the odd integrand兲, we get

Similarly, to verify conservation of momentum we continue as follows:



˜v⬘n˜v⬘n 3␧ f ⳵ v. 2K 2K + 共u − ˜u兲2

+ vn⬘v⬘n… f ⳵ v .

3␧ 3␧ 共u − ˜u兲2 ⳵f − CM ⳵ v = CM 2K 2K 2K + 共u − ˜u兲2 ⳵t − CM



3␧ 4 ˆ −3/2 −3vˆ ⬘vˆ ⬘/4Kˆ 共 ␲K兲 e n n 2K 3

3␧ 1 ˆ ⫻共 34 ␲Kˆ兲−3/2e−3vˆ n⬘vˆ n⬘/4K ⳵ v − C M 2K 2K + 共u − ˜u兲2

共B3兲



C M vi⬘v⬘j

By substituting in the relations vˆ i⬘ = vi − uˆi and ˜v⬘i = vi − ˜ui the integrals can be reduced,

1 3␧ 3␧ ⳵f − CM ⳵ v = CM 2K 2K 2K + 共u − ˜u兲2 ⳵t ⫻







⫻ 共u − ˜u兲2Rij + 2共un − ˜un兲 共B5兲 +

expanding v⬘n = 共un − ˜un兲 + v⬘n and vi = ui + v⬘i gives

冕 ⬘⬘⬘⬘ 冊

冕 ⬘⬘⬘

vi v j vn f ⳵ v

vi v f vnvn f ⳵ v .

共B10兲

The first integral is reduced in terms of “hats,” 3␧ ⳵ ui = C M uˆi 2K ⳵t − CM



3␧ ⳵ Rij = C M 关 32 Kˆ␦ij + 共uˆi − ui兲共uˆ j − u j兲兴 2K ⳵t

3␧ 1 共u − ˜u兲2ui 2K 2K + 共u − ˜u兲2

+ 2共un − ˜un兲Rin + 2Kui +

冕 ⬘⬘⬘ 冊

vi vnvn f ⳵ v .

− CM 共B6兲

Conservation requires the above equation be equal to zero, this implies that

关2K + 共u − ˜u兲2兴共uˆi − ui兲 = 2Rin共un − ˜un兲 +

冕 ⬘⬘⬘

vi vnvn f ⳵ v .

共B7兲 From Appendix C, we see that if f is Gaussian, the last term on the right-hand side goes to zero, and ui = ˜ui = uˆi confirming conservation of momentum for Gaussian PDFs. For nonGaussian PDFs the above relation must be satisfied. The Reynolds stress transport equation is also derived similarly



3␧ 1 共u − ˜u兲2Rij 2K 2K + 共u − ˜u兲2

+ 2共un − ˜un兲

冕 ⬘⬘⬘

vi v j vn f ⳵ v +

冕 ⬘⬘⬘⬘ 冊

vi v j vnvn f ⳵ v .

共B11兲 To ensure the correct dissipation of energy, we require that the model satisfies the equation ⳵K / ⳵t = 21 ⳵ Rii / ⳵t = −␧, −

3␧ ˆ 1 1 ⳵ Rii + 2 共uˆ − u兲2兴 = ␧ = − C M 关K 2 ⳵t 2K + CM ⫻



1 3⑀ 共u − ˜u兲2K + 共un − ˜un兲 2K 2K + 共u − ˜u兲2

冕 ⬘⬘⬘

vi v j vn f ⳵ v +

1 2

冕 ⬘⬘⬘⬘ 冊

vi v j vnvn f ⳵ v . 共B12兲

This can be simplified to

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-15

Phys. Fluids 17, 035101 共2005兲

Modeling return to isotropy

ˆ + 1 共uˆ − u兲2 + 2K K 2 3C M =

1 2K + 共u − ˜u兲2 ⫻

冕 ⬘⬘⬘



ˆ R R K mn mn − 2K2 ⳵ Rij ␧ ␧ K = − Rij − 共Rij − 32 K␦ij兲 ˆ ⳵t K K K RmnRmn 1− + K 2K2

共u − ˜u兲2K + 共un − ˜un兲 1 2

vi vi vn f ⳵ v +

冕 ⬘⬘⬘⬘ 冊 vi vi vnvn f ⳵ v

共B13兲

+

␧ K2



␧ K2

or finally



ˆ 2 1 K + 共uˆ − u兲2 + K 2K 3CM = 共u − ˜u兲2 + +

1 2K

un − ˜un K

冕 ⬘⬘⬘⬘

冊共

2K + 共u − ˜u兲2兲

冕 ⬘⬘⬘

冕 ⬘⬘⬘⬘

vi vi vnvn f ⳵ v =

1 共RnnRii + 2RniRni兲 4K2 共B15兲

or alternatively 2 Kˆ RniRni =1− + . K 3C M 2K2

共B16兲

This can be rearranged to become



−1

ˆ R R ⳵ Rij ␧ K ni ni = 1− + ⳵t K K 2K2

共B17兲

.





冊冉 冊 冊 −1

ˆ R R ␧ K ni ni 1− + K K 2K2

2ˆ K␦ij 3



−1

1 共RnnRij + 2RniRnj兲. 2K

This simplifies to 1 ˆ K RmnRmn 1− + K 2K2



4 Kˆ RniRni + − 3 K −1 4 2K2 + Kˆ RniRni 3 Kˆ RniRni 1− + 1− + 2 K 2K K 2K2



2 ␧ ⫻ Rij − K␦ij + 2 3 K

␦ij 3



−1 RniRnj ˆ K RniRni 1− + K 2K2 共B21兲

.

f = 关共2␲兲3det共Rnm兲兴−1/2e共1/2兲Rnmv⬘nvm⬘ . −1

2 ˆ 3 K␦ij

− Rij −



RniRnj . 共B19兲 K

Rearranging into the classic return model form gives

共C1兲

Since only Rij is a function of space, this implies that the derivative satisfies

⳵f = v⬘l f . ⳵ vk

共C2兲

This allows us to write the third moment as a second moment and then apply the chain rule,

冕 ⬘⬘⬘

vi vnvm f ⳵ v = − Rkm







APPENDIX C: MOMENTS OF A GAUSSIAN PROBABILITY DENSITY FUNCTION

− Rkl

共B18兲

⳵ Rij ␧ = ⳵t K

共B20兲

RniRnj .

If we have a PDF of elliptic Gaussian shape we can write the PDF as

So if we assume Gaussian form for the PDF, the Reynolds transport equation can be written as follows:



⳵ Rij ␧ ␧ = − Rij − ⳵t K K

− RniRni



Kˆ RniRni 1− + K 2K2

共B14兲

For an elliptic Gaussian PDF the above equation reduces to

ˆ R R 3 K ni ni CM = 1 − + 2 K 2K2

1

And when written as follows, the values of CR and CN become apparent:

vi vi vn f ⳵ v

vi vi vnvn f ⳵ v .

ˆ 2 1 K + = K 3CM 4K2

1 ␦ij RmnRmn 3 ˆ K RmnRmn 1− + K 2K2

= − Rkm

冕 ⬘⬘ 冕冉 ⬘ ⬘ vi vn

⳵f ⳵v ⳵ vk



⳵ v⬘i v⬘n ⳵ vi vn f −f ⳵ v . 共C3兲 ⳵ vk ⳵ vk

By Gauss’s divergence theorem the first integral term goes to zero, since f → 0 at infinity. Differentiating the second term gives

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-16

Phys. Fluids 17, 035101 共2005兲

B. Perot and C. Chartrand

冕 ⬘⬘⬘

vi vnvm f ⳵ v = Rkm

冕共⬘ 冕⬘



f vi ␦nk + vn⬘␦ik兲 ⳵ v

vi f ⳵ v + Rim

= Rnm

冕⬘

vn f ⳵ v = 0,

共C4兲

冕 冕

vmvnvi v j fdv = − Rkj

= Rkj

f

⳵f ⬘ vn⬘vi⬘ dv vm ⳵ vk

⳵ vm⬘ vn⬘v⬘i dv. ⳵ vk





⬘ v⬘i ␦nk兲dv f vmvn␦ik + vn⬘vi⬘␦mk + vm

1 + 共un − ˜un兲共um − ˜um兲

共C6兲

␧ 2K



⌸ii关共2␲兲 det共Rˆnm兲兴−1/2 ␧ 2K



−1 ˜vm ⬘ ˜vn⬘ ⌸inRim

1 + 共un − ˜un兲共um − ˜um兲

⌸in −1 R ⌸ pp im

f ⳵ v,

⳵f ␧ ␧ ⳵ v = CM ⌸ii − CM 2K 2K ⳵t ⌸in −1 1 + 共un − ˜un兲共um − ˜um兲 R ⌸ pp im ˜vm˜vn f ⳵ v .



⌸in −1 1 + 共un − ˜un兲共um − ˜um兲 R ⌸ pp im

vp f ⳵ v ,

−1 ⌸inRim

1 + 共un − ˜un兲共um − ˜um兲

⌸in −1 R ⌸ pp im

⬘ + 共um − ˜um兲兴关v⬘n + 共un − ˜un兲兴v p f ⳵ v , 关vm

− CM



−1 ⌸inRim

冕⬘⬘

−1 ˜vm ⬘ ˜v⬘n ⌸inRim

共D6兲

⳵ up ␧ = C M ⌸iiuˆ p 2K ⳵t 共D1兲



␧ 2K

3

ˆ −1

ˆ −1 v p关共2␲兲3det共Rˆnm兲兴−1/2e−共1/2兲Rnmvˆ n⬘vˆ m⬘ ⳵ v

⳵ up ␧ = C M ⌸iiuˆ p 2K ⳵t − CM

⫻e−共1/2兲Rnmvˆ n⬘vˆ m⬘ ⳵ v − C M



共D4兲

共D5兲

Here we verify conservation of mass for the more general relaxation models derived in Sec. VI. The method is the same as before starting from the general relaxation model for the PDF, Eq. 共33兲. Conservation of mass then requires that the right-hand side of the zeroth moment be equal to zero,

⳵f ␧ ⳵ v = CM 2K ⳵t

冕 冕

⳵ up ␧ = CM 兿 2K ii ⳵t

APPENDIX D: GENERAL RELAXATION MOMENTS



⌸in −1 R ⌸ pp im

⫻关Rmn + 共um − ˜um兲共un − ˜un兲兴 = 0.

− CM



共D3兲

The momentum equation is the first moment

冕共⬘⬘



⬘ v⬘n + 共um − ˜um兲共un − ˜un兲兴f ⳵ v , 关vm

−1 ⌸inRim



共C5兲

= RmnRij + RniRmj + RmiRnj .





⌸in −1 R ⌸ pp im

⳵f ␧ ␧ ⳵ v = CM ⌸ii − CM 2K 2K ⳵t

Taking the derivative gives the fourth moment in terms of the second moments

=Rkj

−1 ⌸inRim



1 + 共un − ˜un兲共um − ˜um兲

which is what we would expect since the PDF is an even function and the integrand is an odd function 共cubic兲. The expression 共A2兲 and the chain rule is also useful for evaluating the fourth moment of an elliptic Gaussian,

冕 ⬘⬘⬘⬘

⳵f ␧ ␧ ⳵ v = CM ⌸ii − CM 2K 2K ⳵t

冉冕

− CM

共D2兲

Substituting with ˜vi⬘ = vi⬘ + ui − ˜ui, and recalling that 兰vm ⬘ f ⳵ v = 0 gives

␧ 2K

−1 ⌸inRim

1 + 共un − ˜un兲共um − ˜um兲

⬘ v⬘nv⬘p f ⳵ v + Rmnu p vm

␧ 2K



⌸in −1 R ⌸ pp im

−1 ⌸inRim

1 + 共un − ˜un兲共um − ˜um兲

⌸in −1 R ⌸ pp im

⫻关Rmp共un − ˜un兲 + Rnp共um − ˜um兲 + 共um − ˜um兲共un − ˜un兲u p兴.

共D7兲

Conservation of momentum therefore requires that

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-17

CM

Phys. Fluids 17, 035101 共2005兲

Modeling return to isotropy



再冕



⳵ Rlj ␧ = CM ⌸iiRˆlj + ⌸ii共uˆl − ul兲共uˆ j − u j兲 ˆ ⳵t 2K

−1 ⌸inRim

␧ ␧ ⌸iiuˆ p = C M 2K 2K

1 + 共un − ˜un兲共um − ˜um兲

⌸in −1 R ⌸ pp im

⬘ vn⬘v⬘p f ⳵ v + Rmnu p + 关Rmp共un − ˜un兲 vm





+ Rnp共um − ˜um兲 + 共um − ˜um兲共un − ˜un兲u p兴 ,



uˆ p 1 + 共un − ˜un兲共um − ˜um兲 =

−1 ⌸inRim ⌸ii

冕 ⬘⬘⬘

⌸in −1 R ⌸ pp im



v mv n v p f ⳵ v + u p +



共uˆ p − u p兲 1 + 共un − ˜un兲共um − ˜um兲 ⌸in −1 = R ⌸ii im

冉冕



= C M ⌸ii Rˆlj + 共uˆl − ul兲共uˆ j − u j兲

⌸in −1 R 关Rmp共un − ˜un兲 ⌸ii im

⌸in −1 R ⌸ pp im

共D9兲



+



共D10兲

Below, the Reynolds transport equation is derived for the general relaxation model,

冕 ⬘⬘

vl v p关共2␲兲3det共Rˆnm兲兴−1/2

⫻e−共1/2兲Rnmvˆ ⬘nvˆ m⬘ ⳵ v ˆ −1



−1 ˜vm ⬘ ˜vn⬘ ⌸inRim

1 + 共un − ˜un兲共um − ˜um兲

⌸in −1 R ⌸ pp im



vl⬘v⬘p f ⳵ v .

共E1兲

Substitution gives

⳵ Rlj ␧ = C M ⌸ii ˆ ⳵t 2K



关vˆ l⬘vˆ ⬘j + 共uˆl − ul兲共uˆ j − u j兲兴

ˆ −1 ⫻关共2␲兲3det共Rˆnm兲兴−1/2e−共1/2兲Rnmvˆ ⬘nvˆ m⬘ ⳵ v

− CM



␧ ˆ 2K

−1 ⌸inRim

1 + 共un − ˜un兲共um − ˜um兲

冕 ⬘⬘⬘⬘

vl v j ˜vm˜vn f ⳵ v ,

−1 ⌸inRim ⌸ii

冕 ⬘⬘⬘⬘ 冊

v l v j v mv n f ⳵ v ,

共E4兲

−1 Rˆlj = 共⌸imRmj + ⌸ jmRmi兲 − 共uˆl − ul兲共uˆ j − u j兲 C M ⌸ii

APPENDIX E: GENERAL REYNOLDS TRANSPORT EQUATION





which gives us the definition of Rˆ

+ Rnp共um − ˜um兲 .





vl v j ˜vm˜vn f ⳵ v .

If we choose un = ˜un and use the general form of the Reynolds stress model, ⳵Rlj / ⳵t = −共␧ / 2K兲共⌸imRmj + ⌸ jmRmi兲, we arrive at the following equation:

⬘ vn⬘v⬘p f ⳵ v + Rmp共un − ˜un兲 vm

⳵ Rlp ␧ = CM ⌸ii 2K ⳵t

冕 ⬘⬘⬘⬘

− 共⌸imRmj + ⌸ jmRmi兲

⌸in −1 R 共um − ˜um兲共un − ˜un兲u p + Rnp共um − ˜um兲兴 + ⌸ii im and

⌸in −1 1 + 共un − ˜un兲共um − ˜um兲 R ⌸ pp im

共E3兲

共D8兲 which simplifies to

−1 ⌸inRim

⌸in −1 R ⌸ pp im 共E2兲

−1 ⌸inRim ⌸ii

冕 ⬘⬘⬘⬘

v l v j v mv n f ⳵ v .

共E5兲

1

J. Rotta, “Stastische theorie nichthomegener turbulenz I,” Z. Phys. 129, 547 共1951兲. 2 J. L. Lumley, “Computation modeling of turbulent flows,” Adv. Appl. Mech. 18, 126 共1978兲. 3 L. Le Penven, J. N. Gence, and G. Comte-Bellot, “On the approach to isotropy of homogeneous turbulence: Effect of the partiton of kinetic energy among the velocity components,” in Frontiers in Fluid Mechanics, edited by S. H. Davis and J. L. Lumley 共Springer, New York, 1985兲, p. 1. 4 S. Sarkar and C. G. Speziale, “A simple nonlinear model for the return to isotropy in turbulence,” Phys. Fluids A 2, 84 共1990兲. 5 D. C. Haworth and S. B. Pope, “A generalized Langevin model for turbulent flows,” Phys. Fluids 29, 2 共1986兲. 6 K. S. Choi and J. L. Lumley, “Turbulence and chaotic phenomena in fluids,” in Proceedings of the IUTAM Symposium (Kyoto, Japan), edited by T. Tatsumi 共North-Holland, Amsterdam, 1984兲, p. 267. 7 K. S. Choi and J. L. Lumley, “The return to isotropy of homogeneous turbulence,” J. Fluid Mech. 436, 57 共2001兲. 8 T. J. Craft and B. E. Launder, “Computation of impinging flows using second-moment closure,” in Proceedings of the Eighth Symposium on Turbulent Shear Flows 共Technical University of Munich, Munich, Germany, 1991兲, pp. 8-5-1–8-5-6. 9 M. Yamamoto and C. Arakawa, “Study on the pressure-strain term in Reynolds stress models,” in Proceedings of the Eighth Symposium on Turbulent Shear Flows 共Technical University of Munich, Munich, Germany, 1991兲, pp. III-17-1–III-17-2. 10 C. G. Speziale, S. Sarkar, and T. B. Gatski, “Modeling the pressure-strain correlation of turbulence: An invariant dynamical systems approach,” J. Fluid Mech. 227, 245 共1991兲. 11 P. Degond and M. Lemou, in Hyperbolic Problems, Theory, Numerics and Applications Eighth International Conference in Magdeburg, February– March 2000, edited by H. Freistueler and G. Warnecke 共Birkhauser, Boston, 2001兲. 12 P. Bhatnagar, E. P. Gross, and M. K. Krook, “A model of collision processes in gases. I. Small amplitude processes in charged and neutral onecomponent systems,” Phys. Rev. 94, 511 共1954兲. 13 M. Hallbäck, J. Groth, and A. V. Johansson, “An algebraic model for nonisotropic turbulent dissipation rate in Reynolds stress closures,” Phys. Fluids A 2, 1859 共1990兲. 14 J. B. Perot and P. Moin, “A near wall model for the dissipation tensor,”

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp

035101-18

Phys. Fluids 17, 035101 共2005兲

B. Perot and C. Chartrand

Eleventh Australasian Fluid Mechanics Conference, Hobart, Tasmania, Australia, 13-18, 1992. 15 P. A. Durbin and B. A. Pettersson Reif, Statistical Theory and Modeling for Turbulent Flows 共Wiley, New York, 2001兲. 16 B. E. Launder, G. J. Reece, and W. Rodi, “Progress in the development of a Reynolds stress turbulence closure,” J. Fluid Mech. 68, 537 共1975兲. 17 T. Sjögren and A. V. Johansson, “Development and calibration of algebraic nonlinear models for terms in the Reynolds stress transport equation,” Phys. Fluids 12, 1554 共2000兲. 18 U. Schumann, “Realizability of Reynolds stress turbulence models,” Phys. Fluids 20, 721 共1977兲. 19 S. B. Pope, Turbulent Flows 共Cambridge University Press, Cambridge,

2000兲. S. Tavoularis and S. Corrsin, “Experiments in nearly homogeneous turbulent shear flow with a uniform mean temperature gradient. Part 1,” J. Fluid Mech. 104, 311 共1981兲. 21 P. K. Yeung and S. B. Pope, “Lagrangian statistics from direct numerical simulations of isotropic turbulence,” J. Fluid Mech. 207, 531 共1989兲. 22 P. G. Saffman, “Note on decay of homogeneous turbulence,” Phys. Fluids 10, 1349 共1967兲. 23 J. P. Laval, B. Dubrulle, and J. C. McWilliams, “Langevin models of turbulence: Renormalization group, distant interaction algorithms, or rapid distortion theory?” Phys. Fluids 15, 5 共2003兲. 24 C. Chartrand, “Eddy collision models for turbulence,” Masters thesis, University of Massachusetts, Amherst, 2003. 20

Downloaded 22 Apr 2005 to 128.119.88.168. Redistribution subject to AIP license or copyright, see http://pof.aip.org/pof/copyright.jsp