Fluent Theory Guide-101-136
Fluent Theory Guide-101-136
Fluent Theory Guide-101-136
Turbulence
This chapter provides theoretical background about the turbulence models available in
ANSYS FLUENT. Information is presented in the following sections:
For more information about using these turbulence models in ANSYS FLUENT, see Chap-
ter 12: Modeling Turbulence in the separate User’s Guide.
4.1 Introduction
Turbulent flows are characterized by fluctuating velocity fields. These fluctuations mix
transported quantities such as momentum, energy, and species concentration, and cause
the transported quantities to fluctuate as well. Since these fluctuations can be of small
scale and high frequency, they are too computationally expensive to simulate directly in
practical engineering calculations. Instead, the instantaneous (exact) governing equations
can be time-averaged, ensemble-averaged, or otherwise manipulated to remove the reso-
lution of small scales, resulting in a modified set of equations that are computationally
less expensive to solve. However, the modified equations contain additional unknown
Release 12.0
c ANSYS, Inc. January 29, 2009 4-1
Turbulence
variables, and turbulence models are needed to determine these variables in terms of
known quantities.
ANSYS FLUENT provides the following choices of turbulence models:
• Spalart-Allmaras model
• k- models
– Standard k- model
– Renormalization-group (RNG) k- model
– Realizable k- model
• k-ω models
– Standard k-ω model
– Shear-stress transport (SST) k-ω model
• v 2 -f model (add-on)
• Detached eddy simulation (DES) model, which includes one of the following RANS
models.
– Spalart-Allmaras RANS model
– Realizable k- RANS model
– SST k-ω RANS model
• Large eddy simulation (LES) model, which includes one of the following sub-scale
models.
– Smagorinsky-Lilly subgrid-scale model
– WALE subgrid-scale model
– Dynamic Smagorinsky model
– Kinetic-energy transport subgrid-scale model
Release 12.0
c ANSYS, Inc. January 29, 2009 4-3
Turbulence
LES provides an alternative approach in which large eddies are explicitly computed (re-
solved) in a time-dependent simulation using the “filtered” Navier-Stokes equations. The
rationale behind LES is that by modeling less of turbulence (and resolving more), the
error introduced by turbulence modeling can be reduced. It is also believed to be easier
to find a “universal” model for the small scales, since they tend to be more isotropic and
less affected by the macroscopic features like boundary conditions, than the large eddies.
Filtering is essentially a mathematical manipulation of the exact Navier-Stokes equations
to remove the eddies that are smaller than the size of the filter, which is usually taken as
the mesh size when spatial filtering is employed as in ANSYS FLUENT. Like Reynolds-
averaging, the filtering process creates additional unknown terms that must be modeled
to achieve closure. Statistics of the time-varying flow-fields such as time-averages and
r.m.s. values of the solution variables, which are generally of most engineering interest,
can be collected during the time-dependent simulation.
LES for high Reynolds number industrial flows requires a significant amount of compu-
tational resources. This is mainly because of the need to accurately resolve the energy-
containing turbulent eddies in both space and time domains, which becomes most acute
in near-wall regions where the scales to be resolved become much smaller. Wall functions
in combination with a coarse near wall mesh can be employed, often with some success, to
reduce the cost of LES for wall-bounded flows. However, one needs to carefully consider
the ramification of using wall functions for the flow in question. For the same reason (to
accurately resolve the eddies), LES also requires highly accurate spatial and temporal
discretizations.
where ūi and u0i are the mean and fluctuating velocity components (i = 1, 2, 3).
Likewise, for pressure and other scalar quantities:
φ = φ̄ + φ0 (4.2-2)
∂ρ ∂
+ (ρui ) = 0 (4.2-3)
∂t ∂xi
" !#
∂ ∂ ∂p ∂ ∂ui ∂uj 2 ∂ul ∂
(ρui )+ (ρui uj ) = − + µ + − δij + (−ρu0i u0j ) (4.2-4)
∂t ∂xj ∂xi ∂xj ∂xj ∂xi 3 ∂xl ∂xj
Equations 4.2-3 and 4.2-4 are called Reynolds-averaged Navier-Stokes (RANS) equations.
They have the same general form as the instantaneous Navier-Stokes equations, with
the velocities and other solution variables now representing ensemble-averaged (or time-
averaged) values. Additional terms now appear that represent the effects of turbulence.
These Reynolds stresses, −ρu0i u0j , must be modeled in order to close Equation 4.2-4.
For variable-density flows, Equations 4.2-3 and 4.2-4 can be interpreted as Favre-averaged
Navier-Stokes equations [130], with the velocities representing mass-averaged values. As
such, Equations 4.2-3 and 4.2-4 can be applied to density-varying flows.
The Boussinesq hypothesis is used in the Spalart-Allmaras model, the k- models, and
the k-ω models. The advantage of this approach is the relatively low computational
cost associated with the computation of the turbulent viscosity, µt . In the case of the
Spalart-Allmaras model, only one additional transport equation (representing turbulent
viscosity) is solved. In the case of the k- and k-ω models, two additional transport
equations (for the turbulence kinetic energy, k, and either the turbulence dissipation
rate, , or the specific dissipation rate, ω) are solved, and µt is computed as a function of
k and or k and ω. The disadvantage of the Boussinesq hypothesis as presented is that
it assumes µt is an isotropic scalar quantity, which is not strictly true.
The alternative approach, embodied in the RSM, is to solve transport equations for each
of the terms in the Reynolds stress tensor. An additional scale-determining equation
(normally for ) is also required. This means that five additional transport equations are
required in 2D flows and seven additional transport equations must be solved in 3D.
Release 12.0
c ANSYS, Inc. January 29, 2009 4-5
Turbulence
In many cases, models based on the Boussinesq hypothesis perform very well, and the
additional computational expense of the Reynolds stress model is not justified. However,
the RSM is clearly superior in situations where the anisotropy of turbulence has a dom-
inant effect on the mean flow. Such cases include highly swirling flows and stress-driven
secondary flows.
For details about using the model in ANSYS FLUENT, see Chapter 12: Modeling Turbulence
and Section 12.5: Setting Up the Spalart-Allmaras Model in the separate User’s Guide.
4.3.1 Overview
The Spalart-Allmaras model is a relatively simple one-equation model that solves a mod-
eled transport equation for the kinematic eddy (turbulent) viscosity. This embodies a
relatively new class of one-equation models in which it is not necessary to calculate a
length scale related to the local shear layer thickness. 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 the turbomachinery applications.
( ) !2
∂ ∂ 1 ∂ ∂ νe ∂ νe − Yν + S (4.3-1)
(ρνe) + (ρνeui ) = Gν + (µ + ρνe) + Cb2 ρ ν
∂t ∂xi σeν ∂xj ∂xj ∂xj e
Release 12.0
c ANSYS, Inc. January 29, 2009 4-7
Turbulence
µt = ρνefv1 (4.3-2)
χ3
fv1 = 3
(4.3-3)
χ3 + Cv1
and
νe
χ≡ (4.3-4)
ν
where
νe
Se ≡ S + fv2 (4.3-6)
κ2 d2
and
χ
fv2 = 1 − (4.3-7)
1 + χfv1
Cb1 and κ are constants, d is the distance from the wall, and S is a scalar measure of the
deformation tensor. By default in ANSYS FLUENT, as in the original model proposed by
Spalart and Allmaras, S is based on the magnitude of the vorticity:
q
S≡ 2Ωij Ωij (4.3-8)
The justification for the default expression for S is that, in the wall-bounded flows that
were of most interest when the model was formulated, the turbulence production found
only where vorticity is generated near walls. However, it has since been acknowledged that
one should also take into account the effect of mean strain on the turbulence production,
and a modification to the model has been proposed [65] and incorporated into ANSYS
FLUENT.
This modification combines the measures of both vorticity and the strain tensors in the
definition of S:
where
q q
Cprod = 2.0, |Ωij | ≡ 2Ωij Ωij , |Sij | ≡ 2Sij Sij
Including both the rotation and strain tensors reduces the production of eddy viscosity
and consequently reduces the eddy viscosity itself in regions where the measure of vortic-
ity exceeds that of strain rate. One such example can be found in vortical flows, i.e., flow
near the core of a vortex subjected to a pure rotation where turbulence is known to be
suppressed. Including both the rotation and strain tensors more correctly accounts for
the effects of rotation on turbulence. The default option (including the rotation tensor
only) tends to overpredict the production of eddy viscosity and hence overpredicts the
eddy viscosity itself in certain circumstances.
You can select the modified form for calculating production in the Viscous Model dialog
box.
Release 12.0
c ANSYS, Inc. January 29, 2009 4-9
Turbulence
2
νe
Yν = Cw1 ρfw (4.3-12)
d
where
#1/6
6
"
1 + Cw3
fw = g 6 6
(4.3-13)
g + Cw3
g = r + Cw2 r6 − r (4.3-14)
νe
r≡ e (4.3-15)
Sκ2 d2
Cw1 , Cw2 , and Cw3 are constants, and Se is given by Equation 4.3-6. Note that the
modification described above to include the effects of mean strain on S will also affect
the value of Se used to compute r.
2
Cb1 = 0.1355, Cb2 = 0.622, σeν = , Cv1 = 7.1
3
Cb1 (1 + Cb2 )
Cw1 = + , Cw2 = 0.3, Cw3 = 2.0, κ = 0.4187
κ2 σeν
u ρuτ y
= (4.3-16)
uτ µ
If the mesh is too coarse to resolve the viscous sublayer, then it is assumed that the
centroid of the wall-adjacent cell falls within the logarithmic region of the boundary
layer, and the law-of-the-wall is employed:
!
u 1 ρuτ y
= ln E (4.3-17)
uτ κ µ
where u is the velocity parallel to the wall, uτ is the shear velocity, y is the distance from
the wall, κ is the von Kármán constant (0.4187), and E = 9.793.
where k, in this case, is the thermal conductivity, E is the total energy, and (τij )eff is the
deviatoric stress tensor, defined as
!
∂uj ∂ui 2 ∂uk
(τij )eff = µeff + − µeff δij
∂xi ∂xj 3 ∂xk
• Section 4.4.7: Convective Heat and Mass Transfer Modeling in the k- Models
Release 12.0
c ANSYS, Inc. January 29, 2009 4-11
Turbulence
For details about using the models in ANSYS FLUENT, see Chapter 12: Modeling Turbulence
and Section 12.6: Setting Up the k- Model in the separate User’s Guide.
This section presents the standard, RNG, and realizable k- models. All three models
have similar forms, with transport equations for k and . The major differences in the
models are as follows:
The transport equations, the methods of calculating turbulent viscosity, and model con-
stants are presented separately for each model. The features that are essentially common
to all models follow, including turbulent generation due to shear buoyancy, accounting
for the effects of compressibility, and modeling heat and mass transfer.
and
2
" #
∂ ∂ ∂ µt ∂
(ρ) + (ρui ) = µ+ + C1 (Gk + C3 Gb ) − C2 ρ + S (4.4-2)
∂t ∂xi ∂xj σ ∂xj k k
In these equations, Gk represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 4.4.4: Modeling Turbulent
Production in the k- Models. Gb is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 4.4.5: Effects of Buoyancy on Turbu-
lence in the k- Models. YM represents the contribution of the fluctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Sec-
tion 4.4.6: Effects of Compressibility on Turbulence in the k- Models. C1 , C2 , and C3
are constants. σk and σ are the turbulent Prandtl numbers for k and , respectively. Sk
and S are user-defined source terms.
k2
µt = ρCµ (4.4-3)
where Cµ is a constant.
Model Constants
The model constants C1 , C2 , Cµ , σk , and σ have the following default values [180]:
These default values have been determined from experiments with air and water for funda-
mental turbulent shear flows including homogeneous shear flows and decaying isotropic
grid turbulence. They have been found to work fairly well for a wide range of wall-
bounded and free shear flows.
Release 12.0
c ANSYS, Inc. January 29, 2009 4-13
Turbulence
Although the default values of the model constants are the standard ones most widely
accepted, you can change them (if needed) in the Viscous Model dialog box.
• The RNG model has an additional term in its equation that significantly improves
the accuracy for rapidly strained flows.
• The effect of swirl on turbulence is included in the RNG model, enhancing accuracy
for swirling flows.
• The RNG theory provides an analytical formula for turbulent Prandtl numbers,
while the standard k- model uses user-specified, constant values.
• While the standard k- model is a high-Reynolds-number model, the RNG theory
provides an analytically-derived differential formula for effective viscosity that ac-
counts for low-Reynolds-number effects. Effective use of this feature does, however,
depend on an appropriate treatment of the near-wall region.
These features make the RNG k- model more accurate and reliable for a wider class of
flows than the standard k- model.
The RNG-based k- turbulence model is derived from the instantaneous Navier-Stokes
equations, using a mathematical technique called “renormalization group” (RNG) meth-
ods. The analytical derivation results in a model with constants different from those in
the standard k- model, and additional terms and functions in the transport equations
for k and . A more comprehensive description of RNG theory and its application to
turbulence can be found in [259].
and
2
!
∂ ∂ ∂ ∂
(ρ) + (ρui ) = α µeff + C1 (Gk + C3 Gb ) − C2 ρ − R + S (4.4-5)
∂t ∂xi ∂xj ∂xj k k
In these equations, Gk represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 4.4.4: Modeling Turbulent
Production in the k- Models. Gb is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 4.4.5: Effects of Buoyancy on Turbu-
lence in the k- Models. YM represents the contribution of the fluctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Sec-
tion 4.4.6: Effects of Compressibility on Turbulence in the k- Models. The quantities αk
and α are the inverse effective Prandtl numbers for k and , respectively. Sk and S are
user-defined source terms.
ρ2 k
!
ν̂
d √ = 1.72 √ dν̂ (4.4-6)
µ ν̂ 3 − 1 + Cν
where
ν̂ = µeff /µ
Cν ≈ 100
Equation 4.4-6 is integrated to obtain an accurate description of how the effective tur-
bulent transport varies with the effective Reynolds number (or eddy scale), allowing the
model to better handle low-Reynolds-number and near-wall flows.
Release 12.0
c ANSYS, Inc. January 29, 2009 4-15
Turbulence
k2
µt = ρCµ (4.4-7)
with Cµ = 0.0845, derived using RNG theory. It is interesting to note that this value
of Cµ is very close to the empirically-determined value of 0.09 used in the standard k-
model.
In ANSYS FLUENT, by default, the effective viscosity is computed using the high-
Reynolds-number form in Equation 4.4-7. However, there is an option available that
allows you to use the differential relation given in Equation 4.4-6 when you need to
include low-Reynolds-number effects.
where µt0 is the value of turbulent viscosity calculated without the swirl modification
using either Equation 4.4-6 or Equation 4.4-7. Ω is a characteristic swirl number eval-
uated within ANSYS FLUENT, and αs is a swirl constant that assumes different values
depending on whether the flow is swirl-dominated or only mildly swirling. This swirl
modification always takes effect for axisymmetric, swirling flows and three-dimensional
flows when the RNG model is selected. For mildly swirling flows (the default in ANSYS
FLUENT), αs is set to 0.07. For strongly swirling flows, however, a higher value of αs
can be used.
0.6321 0.3679
α − 1.3929 α + 2.3929 µmol
= (4.4-9)
α0 − 1.3929
α0 + 2.3929
µeff
Cµ ρη 3 (1 − η/η0 ) 2
R = (4.4-10)
1 + βη 3 k
2
!
∂ ∂ ∂ ∂ ∗
(ρ) + (ρui ) = α µeff + C1 (Gk + C3 Gb ) − C2 ρ (4.4-11)
∂t ∂xi ∂xj ∂xj k k
∗
where C2 is given by
∗ Cµ η 3 (1 − η/η0 )
C2 ≡ C2 + (4.4-12)
1 + βη 3
∗
In regions where η < η0 , the R term makes a positive contribution, and C2 becomes
larger than C2 . In the logarithmic layer, for instance, it can be shown that η ≈ 3.0,
∗
giving C2 ≈ 2.0, which is close in magnitude to the value of C2 in the standard k-
model (1.92). As a result, for weakly to moderately strained flows, the RNG model tends
to give results largely comparable to the standard k- model.
In regions of large strain rate (η > η0 ), however, the R term makes a negative contribu-
∗
tion, making the value of C2 less than C2 . In comparison with the standard k- model,
the smaller destruction of augments , reducing k and, eventually, the effective viscosity.
As a result, in rapidly strained flows, the RNG model yields a lower turbulent viscosity
than the standard k- model.
Thus, the RNG model is more responsive to the effects of rapid strain and streamline
curvature than the standard k- model, which explains the superior performance of the
RNG model for certain classes of flows.
Model Constants
The model constants C1 and C2 in Equation 4.4-5 have values derived analytically by
the RNG theory. These values, used by default in ANSYS FLUENT, are
Release 12.0
c ANSYS, Inc. January 29, 2009 4-17
Turbulence
• The realizable k- model contains a new formulation for the turbulent viscosity.
• A new transport equation for the dissipation rate, , has been derived from an exact
equation for the transport of the mean-square vorticity fluctuation.
The term “realizable” means that the model satisfies certain mathematical constraints
on the Reynolds stresses, consistent with the physics of turbulent flows. Neither the
standard k- model nor the RNG k- model is realizable.
An immediate benefit of the realizable k- model is that it more accurately predicts
the spreading rate of both planar and round jets. It is also likely to provide superior
performance for flows involving rotation, boundary layers under strong adverse pressure
gradients, separation, and recirculation.
To understand the mathematics behind the realizable k- model, consider combining
the Boussinesq relationship (Equation 4.2-5) and the eddy viscosity definition (Equa-
tion 4.4-3) to obtain the following expression for the normal Reynolds stress in an in-
compressible strained mean flow:
2 ∂U
u2 = k − 2 νt (4.4-13)
3 ∂x
Using Equation 4.4-3 for νt ≡ µt /ρ, one obtains the result that the normal stress, u2 ,
which by definition is a positive quantity, becomes negative, i.e., “non-realizable”, when
the strain is large enough to satisfy
k ∂U 1
> ≈ 3.7 (4.4-14)
∂x 3Cµ
Similarly, it can also be shown that the Schwarz inequality for shear stresses (uα uβ 2 ≤
u2α u2β ; no summation over α and β) can be violated when the mean strain rate is large.
The most straightforward way to ensure the realizability (positivity of normal stresses
and Schwarz inequality for shear stresses) is to make Cµ variable by sensitizing it to
the mean flow (mean deformation) and the turbulence (k, ). The notion of variable
Cµ is suggested by many modelers including Reynolds [291], and is well substantiated
by experimental evidence. For example, Cµ is found to be around 0.09 in the inertial
sublayer of equilibrium boundary layers, and 0.05 in a strong homogeneous shear flow.
Both the realizable and RNG k- models have shown substantial improvements over the
standard k- model where the flow features include strong streamline curvature, vortices,
and rotation. Since the model is still relatively new, it is not clear in exactly which
instances the realizable k- model consistently outperforms the RNG model. However,
initial studies have shown that the realizable model provides the best performance of all
the k- model versions for several validations of separated flows and flows with complex
secondary flow features.
One of the weaknesses of the standard k- model or other traditional k- models lies with
the modeled equation for the dissipation rate (). The well-known round-jet anomaly
(named based on the finding that the spreading rate in planar jets is predicted reasonably
well, but prediction of the spreading rate for axisymmetric jets is unexpectedly poor) is
considered to be mainly due to the modeled dissipation equation.
The realizable k- model proposed by Shih et al. [313] was intended to address these
deficiencies of traditional k- models by adopting the following:
• A new model equation for dissipation () based on the dynamic equation of the
mean-square vorticity fluctuation.
One limitation of the realizable k- model is that it produces non-physical turbulent
viscosities in situations when the computational domain contains both rotating and sta-
tionary fluid zones (e.g., multiple reference frames, rotating sliding meshes). This is due
to the fact that the realizable k- model includes the effects of mean rotation in the
definition of the turbulent viscosity (see Equations 4.4-17–4.4-19). This extra rotation
effect has been tested on single rotating reference frame systems and showed superior
behavior over the standard k- model. However, due to the nature of this modification,
its application to multiple reference frame systems should be taken with some caution.
See Section 4.4.3: Modeling the Turbulent Viscosity for information about how to include
or exclude this term from the model.
Release 12.0
c ANSYS, Inc. January 29, 2009 4-19
Turbulence
" #
∂ ∂ ∂ µt ∂k
(ρk) + (ρkuj ) = µ+ + Gk + Gb − ρ − YM + Sk (4.4-15)
∂t ∂xj ∂xj σk ∂xj
and
2
" #
∂ ∂ ∂ µt ∂
(ρ) + (ρuj ) = µ+ + ρ C1 S − ρ C2 √ + C1 C3 Gb + S
∂t ∂xj ∂xj σ ∂xj k + ν k
(4.4-16)
where
" #
η k q
C1 = max 0.43, , η=S , S= 2Sij Sij
η+5
In these equations, Gk represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 4.4.4: Modeling Turbulent
Production in the k- Models. Gb is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 4.4.5: Effects of Buoyancy on Turbu-
lence in the k- Models. YM represents the contribution of the fluctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Sec-
tion 4.4.6: Effects of Compressibility on Turbulence in the k- Models. C2 and C1 are
constants. σk and σ are the turbulent Prandtl numbers for k and , respectively. Sk and
S are user-defined source terms.
Note that the k equation (Equation 4.4-15) is the same as that in the standard k-
model (Equation 4.4-1) and the RNG k- model (Equation 4.4-4), except for the model
constants. However, the form of the equation is quite different from those in the
standard and RNG-based k- models (Equations 4.4-2 and 4.4-5). One of the noteworthy
features is that the production term in the equation (the second term on the right-hand
side of Equation 4.4-16) does not involve the production of k; i.e., it does not contain
the same Gk term as the other k- models. It is believed that the present form better
represents the spectral energy transfer. Another desirable feature is that the destruction
term (the next to last term on the right-hand side of Equation 4.4-16) does not have any
singularity; i.e., its denominator never vanishes, even if k vanishes or becomes smaller
than zero. This feature is contrasted with traditional k- models, which have a singularity
due to k in the denominator.
This model has been extensively validated for a wide range of flows [167, 313], including
rotating homogeneous shear flows, free flows including jets and mixing layers, channel
and boundary layer flows, and separated flows. For all these cases, the performance of
the model has been found to be substantially better than that of the standard k- model.
Especially noteworthy is the fact that the realizable k- model resolves the round-jet
anomaly; i.e., it predicts the spreading rate for axisymmetric jets as well as that for
planar jets.
k2
µt = ρCµ (4.4-17)
The difference between the realizable k- model and the standard and RNG k- models
is that Cµ is no longer constant. It is computed from
1
Cµ = ∗ (4.4-18)
A0 + As kU
where
q
∗
U ≡ Sij Sij + Ω
e Ω e
ij ij (4.4-19)
and
Ω ij = Ωij − 2ijk ωk
e
where Ωij is the mean rate-of-rotation tensor viewed in a rotating reference frame with
the angular velocity ωk . The model constants A0 and As are given by
√
A0 = 4.04, As = 6 cos φ
where
√
!
1 Sij Sjk Ski e q 1 ∂uj ∂ui
φ = cos−1 ( 6W ), W = , S = S ij S ij , S ij = +
3 Se3 2 ∂xi ∂xj
Release 12.0
c ANSYS, Inc. January 29, 2009 4-21
Turbulence
It can be seen that Cµ is a function of the mean strain and rotation rates, the angular
velocity of the system rotation, and the turbulence fields (k and ). Cµ in Equation 4.4-17
can be shown to recover the standard value of 0.09 for an inertial sublayer in an equilib-
rium boundary layer.
Model Constants
The model constants C2 , σk , and σ have been established to ensure that the model
performs well for certain canonical flows. The model constants are
∂uj
Gk = −ρu0i u0j (4.4-20)
∂xi
G k = µt S 2 (4.4-21)
q
S≡ 2Sij Sij (4.4-22)
i When using the high-Reynolds number k- versions, µeff is used in lieu of
µt in Equation 4.4-21.
µt ∂T
Gb = βgi (4.4-23)
Prt ∂xi
where Prt is the turbulent Prandtl number for energy and gi is the component of the
gravitational vector in the ith direction. For the standard and realizable k- models,
the default value of Prt is 0.85. In the case of the RNG k- model, Prt = 1/α, where
α is given by Equation 4.4-9, but with α0 = 1/Pr = k/µcp . The coefficient of thermal
expansion, β, is defined as
!
1 ∂ρ
β=− (4.4-24)
ρ ∂T p
µt ∂ρ
Gb = −gi (4.4-25)
ρPrt ∂xi
It can be seen from the transport equations for k (Equations 4.4-1, 4.4-4, and 4.4-15)
that turbulence kinetic energy tends to be augmented (Gb > 0) in unstable stratification.
For stable stratification, buoyancy tends to suppress the turbulence (Gb < 0). In ANSYS
FLUENT, the effects of buoyancy on the generation of k are always included when you
have both a non-zero gravity field and a non-zero temperature (or density) gradient.
While the buoyancy effects on the generation of k are relatively well understood, the
effect on is less clear. In ANSYS FLUENT, by default, the buoyancy effects on are
neglected simply by setting Gb to zero in the transport equation for (Equation 4.4-2,
4.4-5, or 4.4-16).
However, you can include the buoyancy effects on in the Viscous Model dialog box.
In this case, the value of Gb given by Equation 4.4-25 is used in the transport equation
for (Equation 4.4-2, 4.4-5, or 4.4-16).
The degree to which is affected by the buoyancy is determined by the constant C3 . In
ANSYS FLUENT, C3 is not specified, but is instead calculated according to the following
relation [127]:
Release 12.0
c ANSYS, Inc. January 29, 2009 4-23
Turbulence
v
C3 = tanh (4.4-26)
u
where v is the component of the flow velocity parallel to the gravitational vector and
u is the component of the flow velocity perpendicular to the gravitational vector. In
this way, C3 will become 1 for buoyant shear layers for which the main flow direction is
aligned with the direction of gravity. For buoyant shear layers that are perpendicular to
the gravitational vector, C3 will become zero.
YM = 2ρM2t (4.4-27)
4.4.7 Convective Heat and Mass Transfer Modeling in the k- Models
In ANSYS FLUENT, turbulent heat transport is modeled using the concept of Reynolds’
analogy to turbulent momentum transfer. The “modeled” energy equation is thus given
by the following:
!
∂ ∂ ∂ ∂T
(ρE) + [ui (ρE + p)] = keff + ui (τij )eff + Sh (4.4-29)
∂t ∂xi ∂xj ∂xj
where E is the total energy, keff is the effective thermal conductivity, and
(τij )eff is the deviatoric stress tensor, defined as
!
∂uj ∂ui 2 ∂uk
(τij )eff = µeff + − µeff δij
∂xi ∂xj 3 ∂xk
The term involving (τij )eff represents the viscous heating, and is always computed in the
density-based solvers. It is not computed by default in the pressure-based solver, but it
can be enabled in the Viscous Model dialog box.
Additional terms may appear in the energy equation, depending on the physical models
you are using. See Section 5.2.1: Heat Transfer Theory for more details.
For the standard and realizable k- models, the effective thermal conductivity is given
by
c p µt
keff = k +
Prt
where k, in this case, is the thermal conductivity. The default value of the turbulent
Prandtl number is 0.85. You can change the value of the turbulent Prandtl number in
the Viscous Model dialog box.
For the RNG k- model, the effective thermal conductivity is
Release 12.0
c ANSYS, Inc. January 29, 2009 4-25
Turbulence
For details about using the models in ANSYS FLUENT, see Chapter 12: Modeling Turbulence
and Section 12.7: Setting Up the k-ω Model in the separate User’s Guide.
This section presents the standard [379] and shear-stress transport (SST) [224] k-ω mod-
els. Both models have similar forms, with transport equations for k and ω. The major
ways in which the SST model [225] differs from the standard model are as follows:
• gradual change from the standard k-ω model in the inner region of the boundary
layer to a high-Reynolds-number version of the k- model in the outer part of the
boundary layer
• modified turbulent viscosity formulation to account for the transport effects of the
principal turbulent shear stress
and
!
∂ ∂ ∂ ∂ω
(ρω) + (ρωui ) = Γω + G ω − Yω + S ω (4.5-2)
∂t ∂xi ∂xj ∂xj
In these equations, Gk represents the generation of turbulence kinetic energy due to mean
velocity gradients. Gω represents the generation of ω. Γk and Γω represent the effective
diffusivity of k and ω, respectively. Yk and Yω represent the dissipation of k and ω due
to turbulence. All of the above terms are calculated as described below. Sk and Sω are
user-defined source terms.
µt
Γk = µ + (4.5-3)
σk
µt
Γω = µ+ (4.5-4)
σω
where σk and σω are the turbulent Prandtl numbers for k and ω, respectively. The
turbulent viscosity, µt , is computed by combining k and ω as follows:
ρk
µt = α ∗ (4.5-5)
ω
Release 12.0
c ANSYS, Inc. January 29, 2009 4-27
Turbulence
Low-Reynolds-Number Correction
where
ρk
Ret = (4.5-7)
µω
Rk = 6 (4.5-8)
βi
α0∗ = (4.5-9)
3
βi = 0.072 (4.5-10)
The term Gk represents the production of turbulence kinetic energy. From the exact
equation for the transport of k, this term may be defined as
∂uj
Gk = −ρu0i u0j (4.5-11)
∂xi
G k = µt S 2 (4.5-12)
where S is the modulus of the mean rate-of-strain tensor, defined in the same way as for
the k- model (see Equation 4.4-22).
Production of ω
ω
Gω = α Gk (4.5-13)
k
where Gk is given by Equation 4.5-11.
The coefficient α is given by
!
α∞ α0 + Ret /Rω
α= ∗ (4.5-14)
α 1 + Ret /Rω
where Rω = 2.95. α∗ and Ret are given by Equations 4.5-6 and 4.5-7, respectively.
Note that, in the high-Reynolds-number form of the k-ω model, α = α∞ = 1.
Yk = ρ β ∗ fβ ∗ k ω (4.5-15)
where
1 χk ≤ 0
fβ ∗ = 1+680χ2k (4.5-16)
1+400χ2k
χk > 0
where
1 ∂k ∂ω
χk ≡ (4.5-17)
ω 3 ∂xj ∂xj
and
Release 12.0
c ANSYS, Inc. January 29, 2009 4-29
Turbulence
Rβ = 8 (4.5-21)
∗
β∞ = 0.09 (4.5-22)
Dissipation of ω
Yω = ρ β fβ ω 2 (4.5-23)
where
1 + 70χω
fβ = (4.5-24)
1 + 80χω
Ω Ω S
ij jk ki
χω = (4.5-25)
(β ∗ ω)3
∞
!
1 ∂ui ∂uj
Ωij = − (4.5-26)
2 ∂xj ∂xi
β∗
" #
β = βi 1 − i ζ ∗ F (Mt ) (4.5-27)
βi
βi∗ and F (Mt ) are defined by Equations 4.5-19 and 4.5-28, respectively.
Compressibility Correction
where
2k
M2t ≡ (4.5-29)
a2
Mt0 = 0.25 (4.5-30)
q
a = γRT (4.5-31)
Model Constants
∗ 1 ∗
α∞ = 1, α∞ = 0.52, α0 = , β∞ = 0.09, βi = 0.072, Rβ = 8
9
Rk = 6, Rω = 2.95, ζ ∗ = 1.5, Mt0 = 0.25, σk = 2.0, σω = 2.0
• The standard k-ω model and the transformed k- model are both multiplied by a
blending function and both models are added together. The blending function is
designed to be one in the near-wall region, which activates the standard k-ω model,
and zero away from the surface, which activates the transformed k- model.
• The definition of the turbulent viscosity is modified to account for the transport of
the turbulent shear stress.
These features make the SST k-ω model more accurate and reliable for a wider class
of flows (e.g., adverse pressure gradient flows, airfoils, transonic shock waves) than the
standard k-ω model. Other modifications include the addition of a cross-diffusion term
in the ω equation and a blending function to ensure that the model equations behave
appropriately in both the near-wall and far-field zones.
Release 12.0
c ANSYS, Inc. January 29, 2009 4-31
Turbulence
and
!
∂ ∂ ∂ ∂ω
(ρω) + (ρωui ) = Γω + G ω − Y ω + Dω + S ω (4.5-33)
∂t ∂xi ∂xj ∂xj
µt
Γk = µ + (4.5-34)
σk
µt
Γω = µ+ (4.5-35)
σω
where σk and σω are the turbulent Prandtl numbers for k and ω, respectively. The
turbulent viscosity, µt , is computed as follows:
ρk 1
µt = h i (4.5-36)
ω max 1∗ , SF2
α a1 ω
1
σk = (4.5-37)
F1 /σk,1 + (1 − F1 )/σk,2
1
σω = (4.5-38)
F1 /σω,1 + (1 − F1 )/σω,2
F1 = tanh Φ41 (4.5-39)
" √ ! #
k 500µ 4ρk
Φ1 = min max , 2 , (4.5-40)
0.09ωy ρy ω σω,2 Dω+ y 2
" #
1 1 ∂k ∂ω −10
Dω+ = max 2ρ , 10 (4.5-41)
σω,2 ω ∂xj ∂xj
F2 = tanh Φ22 (4.5-42)
" √ #
k 500µ
Φ2 = max 2 , (4.5-43)
0.09ωy ρy 2 ω
where y is the distance to the next surface and Dω+ is the positive portion of the cross-
diffusion term (see Equation 4.5-52).
The term G
e represents the production of turbulence kinetic energy, and is defined as:
k
where Gk is defined in the same manner as in the standard k-ω model. See Sec-
tion 4.5.1: Modeling the Turbulence Production for details.
Production of ω
α e
Gω = Gk (4.5-45)
νt
Note that this formulation differs from the standard k-ω model. The difference between
the two models also exists in the way the term α∞ is evaluated. In the standard k-ω
model, α∞ is defined as a constant (0.52). For the SST k-ω model, α∞ is given by
Release 12.0
c ANSYS, Inc. January 29, 2009 4-33
Turbulence
where
βi,1 κ2
α∞,1 = ∗
− q (4.5-47)
β∞ ∗
σw,1 β∞
βi,2 κ2
α∞,2 = ∗
− q (4.5-48)
β∞ ∗
σw,2 β∞
where κ is 0.41.
The term Yk represents the dissipation of turbulence kinetic energy, and is defined in a
similar manner as in the standard k-ω model (see Section 4.5.1: Modeling the Turbulence
Dissipation). The difference is in the way the term fβ ∗ is evaluated. In the standard k-ω
model, fβ ∗ is defined as a piecewise function. For the SST k-ω model, fβ ∗ is a constant
equal to 1. Thus,
Yk = ρβ ∗ kω (4.5-49)
Dissipation of ω
Yk = ρβω 2 (4.5-50)
Cross-Diffusion Modification
The SST k-ω model is based on both the standard k-ω model and the standard k- model.
To blend these two models together, the standard k- model has been transformed into
equations based on k and ω, which leads to the introduction of a cross-diffusion term
(Dω in Equation 4.5-33). Dω is defined as
1 ∂k ∂ω
Dω = 2 (1 − F1 ) ρσω,2 (4.5-52)
ω ∂xj ∂xj
For details about the various k- models, see Section 4.4: Standard, RNG, and Realizable
k- Models.
Model Constants
σk,1 = 1.176, σω,1 = 2.0, σk,2 = 1.0, σω,2 = 1.168
ρ (u∗ )2 +
ωw = ω (4.5-53)
µ
Release 12.0
c ANSYS, Inc. January 29, 2009 4-35
Turbulence
where
2
50
ks+
ks+ < 25
ωw+ = (4.5-55)
100
ks+ ≥ 25
ks+
where
ρks u∗
!
ks+ = max 1.0, (4.5-56)
µ
1 du+turb
ω+ = q +
(4.5-57)
∗
β∞ dy
u∗
ω=q (4.5-58)
∗ κy
β∞
Note that in the case of a wall cell being placed in the buffer region, ANSYS FLUENT
will blend ω + between the logarithmic and laminar sublayer values.