You are on page 1of 1

Search Wikipedia

Turbulence modeling
Article Talk

This article needs additional citations for verification. Please help improve this article by adding
citations to reliable sources. Unsourced material may be challenged and removed.
Find sources: "Turbulence modeling" – news · newspapers · books · scholar · JSTOR (November 2016)
(Learn how and when to remove this template message)

In fluid dynamics, turbulence modeling is the construction and use of a mathematical model to predict the
effects of turbulence. Turbulent flows are commonplace in most real-life scenarios, including the flow of
blood through the cardiovascular system,[1] the airflow over an aircraft wing,[2] the re-entry of space
vehicles,[3] besides others. In spite of decades of research, there is no analytical theory to predict the
evolution of these turbulent flows. The equations governing turbulent flows can only be solved directly for
simple cases of flow. For most real-life turbulent flows, CFD simulations use turbulent models to predict the
evolution of turbulence. These turbulence models are simplified constitutive equations that predict the
statistical evolution of turbulent flows.[4]

Contents

Closure problem

Eddy viscosity

Prandtl's mixing-length concept

Smagorinsky model for the sub-grid scale eddy viscosity A simulation of a physical wind tunnel
airplane model
Spalart–Allmaras, k–ε and k–ω models

Common models

References
Notes

Other

Closure problem

The Navier–Stokes equations govern the velocity and pressure of a fluid flow. In a turbulent flow, each of
these quantities may be decomposed into a mean part and a fluctuating part. Averaging the equations gives
the Reynolds-averaged Navier–Stokes (RANS) equations, which govern the mean flow. However, the
nonlinearity of the Navier–Stokes equations means that the velocity fluctuations still appear in the RANS
equations, in the nonlinear term from the convective acceleration. This term is known as the
Reynolds stress, .[5] Its effect on the mean flow is like that of a stress term, such as from pressure or
viscosity.

To obtain equations containing only the mean velocity and pressure, we need to close the RANS equations
by modelling the Reynolds stress term as a function of the mean flow, removing any reference to the
fluctuating part of the velocity. This is the closure problem.

Eddy viscosity

Joseph Valentin Boussinesq was the first to attack the closure problem,[6] by introducing the concept of
eddy viscosity. In 1877 Boussinesq proposed relating the turbulence stresses to the mean flow to close the
system of equations. Here the Boussinesq hypothesis is applied to model the Reynolds stress term. Note
that a new proportionality constant , the turbulence eddy viscosity, has been introduced. Models of
this type are known as eddy viscosity models or EVM's.

which can be written in shorthand as

where
is the mean rate of strain tensor

is the turbulence eddy viscosity

is the turbulence kinetic energy

and is the Kronecker delta.

In this model, the additional turbulence stresses are given by augmenting the molecular viscosity with an
eddy viscosity.[7] This can be a simple constant eddy viscosity (which works well for some free shear flows
such as axisymmetric jets, 2-D jets, and mixing layers).

The Boussinesq hypothesis – although not explicitly stated by Boussinesq at the time – effectively consists
of the assumption that the Reynolds stress tensor is aligned with the strain tensor of the mean flow (i.e.:
that the shear stresses due to turbulence act in the same direction as the shear stresses produced by the
averaged flow). It has since been found to be significantly less accurate than most practitioners would
assume.[8] Still, turbulence models which employ the Boussinesq hypothesis have demonstrated significant
practical value. In cases with well-defined shear layers, this is likely due the dominance of streamwise shear
components, so that considerable relative errors in flow-normal components are still negligible in absolute
terms. Beyond this, most eddy viscosity turbulence models contain coefficients which are calibrated
against measurements, and thus produce reasonably accurate overall outcomes for flow fields of similar
type as used for calibration.

Prandtl's mixing-length concept

Later, Ludwig Prandtl introduced the additional concept of the mixing length,[9] along with the idea of a
boundary layer. For wall-bounded turbulent flows, the eddy viscosity must vary with distance from the wall,
hence the addition of the concept of a 'mixing length'. In the simplest wall-bounded flow model, the eddy
viscosity is given by the equation:

where:

is the partial derivative of the streamwise velocity (u) with respect to the wall normal direction (y);

is the mixing length.

This simple model is the basis for the "law of the wall", which is a surprisingly accurate model for wall-
bounded, attached (not separated) flow fields with small pressure gradients.

More general turbulence models have evolved over time, with most modern turbulence models given by
field equations similar to the Navier–Stokes equations.

Smagorinsky model for the sub-grid scale eddy viscosity

Joseph Smagorinsky was the first who proposed a formula for the eddy viscosity in Large Eddy Simulation
models,[10] based on the local derivatives of the velocity field and the local grid size:

In the context of Large Eddy Simulation, turbulence modeling refers to the need to parameterize the subgrid
scale stress in terms of features of the filtered velocity field. This field is called subgrid-scale modeling .

Spalart–Allmaras, k–ε and k–ω models

The Boussinesq hypothesis is employed in the Spalart–Allmaras (S–A), k–ε (k–epsilon), and k–ω (k–omega)
models and offers a relatively low cost computation for the turbulence viscosity . The S–A model uses
only one additional equation to model turbulence viscosity transport, while the k–ε and k–ω models use
two.

Common models

The following is a brief overview of commonly employed models in modern engineering applications.

Spalart–Allmaras (S–A)

The Spalart–Allmaras model[11] is a one-equation model that solves a modelled transport equation for the
kinematic eddy turbulent viscosity. The Spalart–Allmaras model was designed specifically for aerospace
applications involving wall-bounded flows and has been shown to give good results for boundary layers
subjected to adverse pressure gradients. It is also gaining popularity in turbomachinery
applications.[citation needed]

k–ε (k–epsilon)

K-epsilon (k-ε) turbulence model[12] is the most common model used in computational fluid dynamics
(CFD) to simulate mean flow characteristics for turbulent flow conditions. It is a two-equation model
which gives a general description of turbulence by means of two transport equations (PDEs). The original
impetus for the K-epsilon model was to improve the mixing-length model, as well as to find an alternative
to algebraically prescribing turbulent length scales in moderate to high complexity flows.

k–ω (k–omega)

In computational fluid dynamics, the k–omega (k–ω) turbulence model[13] is a common two-equation
turbulence model that is used as a closure for the Reynolds-averaged Navier–Stokes equations (RANS
equations). The model attempts to predict turbulence by two partial differential equations for two
variables, k and ω, with the first variable being the turbulence kinetic energy (k) while the second (ω) is
the specific rate of dissipation (of the turbulence kinetic energy k into internal thermal energy).

SST (Menter’s Shear Stress Transport)

SST (Menter's shear stress transport) turbulence model[14] is a widely used and robust two-equation
eddy-viscosity turbulence model used in computational fluid dynamics. The model combines the k-
omega turbulence model and K-epsilon turbulence model such that the k-omega is used in the inner
region of the boundary layer and switches to the k-epsilon in the free shear flow.

Reynolds stress equation model

The Reynolds stress equation model (RSM), also referred to as second moment closure model,[15] is the
most complete classical turbulence modelling approach. Popular eddy-viscosity based models like the k–
ε (k–epsilon) model and the k–ω (k–omega) models have significant shortcomings in complex engineering
flows. This arises due to the use of the eddy-viscosity hypothesis in their formulation. For instance, in
flows with high degrees of anisotropy, significant streamline curvature, flow separation, zones of
recirculating flow or flows influenced by rotational effects, the performance of such models is
unsatisfactory.[16] In such flows, Reynolds stress equation models offer much better accuracy.[17]

Eddy viscosity based closures cannot account for the return to isotropy of turbulence,[18] observed in
decaying turbulent flows. Eddy-viscosity based models cannot replicate the behaviour of turbulent flows
in the Rapid Distortion limit,[19] where the turbulent flow essentially behaves like an elastic medium.[20]

References

Notes
1. ^ Sallam, Ahmed; Hwang, Ned (1984). "Human red 11. ^ Spalart, P.; Allmaras, S. (1992). "A one-equation
blood cell hemolysis in a turbulent shear flow: turbulence model for aerodynamic flows". 30th
contribution of Reynolds shear stresses" . Aerospace Sciences Meeting and Exhibit, AIAA.
Biorheology. 21 (6): 783–97. doi:10.3233/BIR-1984- doi:10.2514/6.1992-439 .
21605 . PMID 6240286 .
12. ^ Hanjalic, K.; Launder, B. (1972). "A Reynolds stress
2. ^ Rhie, C; Chow, Li (1983). "Numerical study of the model of turbulence and its application to thin shear
turbulent flow past an airfoil with trailing edge flows" . Journal of Fluid Mechanics. 52 (4): 609–
separation" (PDF). AIAA Journal. 21 (11): 1525– 638. Bibcode:1972JFM....52..609H .
1532. Bibcode:1983AIAAJ..21.1525R . doi:10.1017/S002211207200268X .
[dead link]
doi:10.2514/3.8284 . S2CID 122631170 .

3. ^ Reddy, K; Silva, D; Krishnendu, Sinha (1983). 13. ^ Wilcox, D. C. (2008). "Formulation of the k-omega
"Hypersonic turbulent flow simulation of Fire II Turbulence Model Revisited". AIAA Journal. 46 (11):
reentry vehicle afterbody" (PDF). AIAA Journal. 2823–2838. Bibcode:2008AIAAJ..46.2823W .
doi:10.2514/1.36541 .
4. ^ Pope, Stephen (2000). Turbulent Flows.
14. ^ Menter, F. R. (1994). "Two-Equation Eddy-
5. ^ Andersson, Bengt; et al. (2012). Computational
Viscosity Turbulence Models for Engineering
fluid dynamics for engineers . Cambridge:
Applications" (PDF). AIAA Journal. 32 (8): 1598–
Cambridge University Press. p. 83 . ISBN 978-1-
1605. Bibcode:1994AIAAJ..32.1598M .
107-01895-2.
doi:10.2514/3.12149 . S2CID 120712103 .[dead link]
6. ^ Boussinesq, Joseph (1903). Boussinesq, J. (1903).
15. ^ Hanjalić, Hanjalić; Launder, Brian (2011). Modelling
Thōrie analytique de la chaleur mise en harmonie
Turbulence in Engineering and the Environment:
avec la thermodynamique et avec la thōrie mc̄ anique
Second-Moment Routes to Closure.
de la lumi_re: Refroidissement et c̄ hauffement par
rayonnement, conductibilit ̄ des tiges, lames et 16. ^ Mishra, Aashwin; Girimaji, Sharath (2013).
masses cristallines, courants de convection, thōrie "Intercomponent energy transfer in incompressible
mc̄ anique de la lumi_re. Gauthier-Villars. homogeneous turbulence: multi-point physics and
amenability to one-point closures". Journal of Fluid
7. ^ John J. Bertin; Jacques Periaux; Josef Ballmann
Mechanics. 731: 639–681.
(1992), Advances in Hypersonics: Modeling
Bibcode:2013JFM...731..639M .
hypersonic flows , Springer, ISBN 9780817636630
doi:10.1017/jfm.2013.343 . S2CID 122537381 .
8. ^ François G. Schmitt (2007), "About Boussinesq's
17. ^ Pope, Stephen. "Turbulent Flows". Cambridge
turbulent viscosity hypothesis: historical remarks and
University Press, 2000.
a direct evaluation of its validity" , Comptes Rendus
Mécanique, 335 (9–10): 617–627, 18. ^ Lumley, John; Newman, Gary (1977). "The return to
doi:10.1016/j.crme.2007.08.004 , isotropy of homogeneous turbulence". Journal of
S2CID 32637068 Fluid Mechanics. 82: 161–178.
Bibcode:1977JFM....82..161L .
9. ^ Prandtl, Ludwig (1925). "Bericht uber
doi:10.1017/s0022112077000585 .
Untersuchungen zur ausgebildeten Turbulenz". Z.
S2CID 39228898 .
Angew. Math. Mech. 5 (2): 136.
Bibcode:1925ZaMM....5..136P . 19. ^ Mishra, Aashwin; Girimaji, Sharath (2013).
doi:10.1002/zamm.19250050212 . "Intercomponent energy transfer in incompressible
homogeneous turbulence: multi-point physics and
10. ^ Smagorinsky, Joseph (1963). "Smagorinsky,
amenability to one-point closures". Journal of Fluid
Joseph. "General circulation experiments with the
Mechanics. 731: 639–681.
primitive equations: I. The basic experiment" .
Bibcode:2013JFM...731..639M .
Monthly Weather Review. 91 (3): 99–164.
doi:10.1017/jfm.2013.343 . S2CID 122537381 .
Bibcode:1963MWRv...91...99S . doi:10.1175/1520-
0493(1963)091<0099:GCEWTP>2.3.CO;2 . 20. ^ Sagaut, Pierre; Cambon, Claude (2008).
Homogeneous Turbulence Dynamics.

Other
Absi, R. (2019) "Eddy Viscosity and Velocity Profiles in Fully-Developed Turbulent Channel Flows" Fluid
Dyn (2019) 54: 137. https://doi.org/10.1134/S0015462819010014

Absi, R. (2021) "Reinvestigating the Parabolic-Shaped Eddy Viscosity Profile for Free Surface Flows"
Hydrology 2021, 8(3), 126. https://doi.org/10.3390/hydrology8030126

Townsend, A.A. (1980) "The Structure of Turbulent Shear Flow" 2nd Edition (Cambridge Monographs on
Mechanics), ISBN 0521298199

Bradshaw, P. (1971) "An introduction to turbulence and its measurement" (Pergamon Press),
ISBN 0080166210

Wilcox C. D., (1998), "Turbulence Modeling for CFD" 2nd Ed., (DCW Industries, La Cañada),
ISBN 0963605100

Last edited on 9 December 2023, at 04:14

Content is available under CC BY-SA 4.0 unless otherwise noted.

Terms of Use • Privacy policy • Desktop


:

You might also like