You are on page 1of 88

Chapter 11.

Modeling Turbulence
This chapter provides details about the turbulence models available in FLUENT.
Information is presented in the following sections:
Section 11.1: Introduction
Section 11.2: Choosing a Turbulence Model
Section 11.3: The Spalart-Allmaras Model
Section 11.4: The Standard, RNG, and Realizable k- Models
Section 11.5: The Standard and SST k- Models
Section 11.6: The Reynolds Stress Model (RSM)
Section 11.7: The Large Eddy Simulation (LES) Model
Section 11.8: The Detached Eddy Simulation (DES) Model
Section 11.9: Near-Wall Treatments for Wall-Bounded Turbulent Flows
Section 11.10: Grid Considerations for Turbulent Flow Simulations
Section 11.11: Problem Setup for Turbulent Flows
Section 11.12: Solution Strategies for Turbulent Flow Simulations
Section 11.13: Postprocessing for Turbulent Flows
c Fluent Inc. January 11, 2005 11-1
Modeling Turbulence
11.1 Introduction
Turbulent ows are characterized by uctuating velocity elds. These uctuations mix
transported quantities such as momentum, energy, and species concentration, and cause
the transported quantities to uctuate as well. Since these uctuations 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 small
scales, resulting in a modied set of equations that are computationally less expensive
to solve. However, the modied equations contain additional unknown variables, and
turbulence models are needed to determine these variables in terms of known quantities.
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- models
Standard k- model
Shear-stress transport (SST) k- model
v
2
-f model
Reynolds stress model (RSM)
Detached eddy simulation (DES) model
Large eddy simulation (LES) model
11.2 Choosing a Turbulence Model
It is an unfortunate fact that no single turbulence model is universally accepted as be-
ing superior for all classes of problems. The choice of turbulence model will depend on
considerations such as the physics encompassed in the ow, the established practice for
a specic class of problem, the level of accuracy required, the available computational
resources, and the amount of time available for the simulation. To make the most ap-
propriate choice of model for your application, you need to understand the capabilities
and limitations of the various options.
The purpose of this section is to give an overview of issues related to the turbulence
models provided in FLUENT. The computational eort and cost in terms of CPU time and
11-2 c Fluent Inc. January 11, 2005
11.2 Choosing a Turbulence Model
memory of the individual models is discussed. While it is impossible to state categorically
which model is best for a specic application, general guidelines are presented to help
you choose the appropriate turbulence model for the ow you want to model.
11.2.1 Reynolds-Averaged Approach vs. LES
Time-dependent solutions of the Navier-Stokes equations for high Reynolds-number tur-
bulent ows in complex geometries which set out to resolve all the way down to the
smallest scales of the motions are unlikely to be attainable for some time to come. Two
alternative methods can be employed to render the Navier-Stokes equations tractable
so that the small-scale turbulent uctuations do not have to be directly simulated:
Reynolds-averaging (or ensemble-averaging) and ltering. Both methods introduce ad-
ditional terms in the governing equations that need to be modeled in order to achieve a
closure for the unknowns.
The Reynolds-averaged Navier-Stokes (RANS) equations govern the transport of the aver-
aged ow quantities, with the whole range of the scales of turbulence being modeled. The
RANS-based modeling approach therefore greatly reduces the required computational ef-
fort and resources, and is widely adopted for practical engineering applications. An entire
hierarchy of closure models are available in FLUENTincluding Spalart-Allmaras, k- and
its variants, k- and its variants, and the RSM. The RANS equations are often used
to compute time-dependent ows, whose unsteadiness may be externally imposed (e.g.,
time-dependent boundary conditions or sources) or self-sustained (e.g., vortex-shedding,
ow instabilities).
LES provides an alternative approach in which large eddies are explicitly computed (re-
solved) in a time-dependent simulation using the ltered 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 nd a universal model for the small scales, since they tend to be more isotropic and
less aected 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 lter, which is usually taken as
the mesh size when spatial ltering is employed as in FLUENT. Like Reynolds-averaging,
the ltering process creates additional unknown terms that must be modeled to achieve
closure. Statistics of the time-varying ow-elds such as time-averages and r.m.s. val-
ues 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 ows requires a signicant amount of compute
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 increasingly 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 ows. However, one needs to carefully consider
c Fluent Inc. January 11, 2005 11-3
Modeling Turbulence
the ramication of using wall functions for the ow in question. For the same reason (to
accurately resolve the eddies), LES also requires highly accurate spatial and temporal
discretizations.
LES/RANS Coupling: Detached Eddy Simulation (DES) Model
The detached eddy simulation (DES) model in FLUENT is based on a modied version of
the Spalart-Allmaras model and can be considered a more practical alternative to LES for
predicting the ow around high-Reynolds-number, high-lift airfoils. The DES approach
combines an unsteady RANS version of the Spalart-Allmaras model with a ltered version
of the same model to create two separate regions inside the ow domain: one that is
LES-based and another that is close to the wall where the modeling is dominated by the
RANS-based approach. The LES region is normally associated with the high-Re core
turbulent region where large turbulence scales play a dominant role. In this region, the
DES model recovers the pure LES model based on a one-equation sub-grid model. Close
to the wall, where viscous eects prevail, the standard RANS model is recovered.
The application of DES, however, may still require signicant CPU resources and there-
fore, as a general guideline, it is recommended that the conventional turbulence models
employing the Reynolds-averaged approach be used for practical calculations. The LES
approach, described further in Section 11.7: The Large Eddy Simulation (LES) Model,
has been made available for you to try if you have the computational resources and are
willing to invest the eort. The rest of this section will deal with the choice of models
using the Reynolds-averaged approach.
11.2.2 Reynolds (Ensemble) Averaging
In Reynolds averaging, the solution variables in the instantaneous (exact) Navier-Stokes
equations are decomposed into the mean (ensemble-averaged or time-averaged) and uc-
tuating components. For the velocity components:
u
i
= u
i
+ u

i
(11.2-1)
where u
i
and u

i
are the mean and uctuating velocity components (i = 1, 2, 3).
Likewise, for pressure and other scalar quantities:
=

+

(11.2-2)
where denotes a scalar such as pressure, energy, or species concentration.
Substituting expressions of this form for the ow variables into the instantaneous conti-
nuity and momentum equations and taking a time (or ensemble) average (and dropping
11-4 c Fluent Inc. January 11, 2005
11.2 Choosing a Turbulence Model
the overbar on the mean velocity, u) yields the ensemble-averaged momentum equations.
They can be written in Cartesian tensor form as:

t
+

x
i
(u
i
) = 0 (11.2-3)

t
(u
i
) +

x
j
(u
i
u
j
) =
p
x
i
+

x
j
_

_
u
i
x
j
+
u
j
x
i

2
3

ij
u
l
x
l
__
+

x
j
(u

i
u

j
)
(11.2-4)
Equations 11.2-3 and 11.2-4 are called Reynolds-averaged Navier-Stokes (RANS) equa-
tions. 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 eects of tur-
bulence. These Reynolds stresses, u

i
u

j
, must be modeled in order to close Equa-
tion 11.2-4.
For variable-density ows, Equations 11.2-3 and 11.2-4 can be interpreted as Favre-
averaged Navier-Stokes equations [131], with the velocities representing mass-averaged
values. As such, Equations 11.2-3 and 11.2-4 can be applied to density-varying ows.
11.2.3 Boussinesq Approach vs. Reynolds Stress Transport Models
The Reynolds-averaged approach to turbulence modeling requires that the Reynolds
stresses in Equation 11.2-4 be appropriately modeled. A common method employs the
Boussinesq hypothesis [131] to relate the Reynolds stresses to the mean velocity gradients:
u

i
u

j
=
t
_
u
i
x
j
+
u
j
x
i
_

2
3
_
k +
t
u
i
x
i
_

ij
(11.2-5)
The Boussinesq hypothesis is used in the Spalart-Allmaras model, the k- models, and
the k- models. The advantage of this approach is the relatively low computational
cost associated with the computation of the turbulent viscosity,
t
. 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- models, two additional transport
equations (for the turbulence kinetic energy, k, and either the turbulence dissipation
rate, , or the specic dissipation rate, ) are solved, and
t
is computed as a function of
k and . The disadvantage of the Boussinesq hypothesis as presented is that it assumes

t
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 ve additional transport equations are
required in 2D ows and seven additional transport equations must be solved in 3D.
c Fluent Inc. January 11, 2005 11-5
Modeling 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 justied. However,
the RSM is clearly superior for situations in which the anisotropy of turbulence has a
dominant eect on the mean ow. Such cases include highly swirling ows and stress-
driven secondary ows.
11.2.4 The Spalart-Allmaras Model
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 specically for aerospace applications involving wall-bounded ows and has been
shown to give good results for boundary layers subjected to adverse pressure gradients.
It is also gaining popularity for turbomachinery applications.
In its original form, the Spalart-Allmaras model is eectively a low-Reynolds-number
model, requiring the viscous-aected region of the boundary layer to be properly resolved.
In FLUENT, however, the Spalart-Allmaras model has been implemented to use wall
functions when the mesh resolution is not suciently ne. This might make it the best
choice for relatively crude simulations on coarse meshes where accurate turbulent ow
computations are not critical. Furthermore, the near-wall gradients of the transported
variable in the model are much smaller than the gradients of the transported variables
in the k- or k- models. This might make the model less sensitive to numerical error
when non-layered meshes are used near walls. See Section 6.1.3: Numerical Diusion for
further discussion of numerical error.
On a cautionary note, however, the Spalart-Allmaras model is still relatively new, and
no claim is made regarding its suitability to all types of complex engineering ows. For
instance, it cannot be relied on to predict the decay of homogeneous, isotropic turbu-
lence. Furthermore, one-equation models are often criticized for their inability to rapidly
accommodate changes in length scale, such as might be necessary when the ow changes
abruptly from a wall-bounded to a free shear ow.
11.2.5 The Standard k- Model
The simplest complete models of turbulence are two-equation models in which the so-
lution of two separate transport equations allows the turbulent velocity and length scales
to be independently determined. The standard k- model in FLUENT falls within this
class of turbulence model and has become the workhorse of practical engineering ow
calculations in the time since it was proposed by Launder and Spalding [183]. Robust-
ness, economy, and reasonable accuracy for a wide range of turbulent ows explain its
popularity in industrial ow and heat transfer simulations. It is a semi-empirical model,
and the derivation of the model equations relies on phenomenological considerations and
empiricism.
11-6 c Fluent Inc. January 11, 2005
11.2 Choosing a Turbulence Model
As the strengths and weaknesses of the standard k- model have become known, improve-
ments have been made to the model to improve its performance. Two of these variants
are available in FLUENT: the RNG k- model [383] and the realizable k- model [306].
11.2.6 The RNG k- Model
The RNG k- model was derived using a rigorous statistical technique (called renormal-
ization group theory). It is similar in form to the standard k- model, but includes the
following renements:
The RNG model has an additional term in its equation that signicantly improves
the accuracy for rapidly strained ows.
The eect of swirl on turbulence is included in the RNG model, enhancing accuracy
for swirling ows.
The RNG theory provides an analytical formula for turbulent Prandtl numbers,
while the standard k- model uses user-specied, constant values.
While the standard k- model is a high-Reynolds-number model, the RNG theory
provides an analytically-derived dierential formula for eective viscosity that ac-
counts for low-Reynolds-number eects. Eective 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
ows than the standard k- model.
11.2.7 The Realizable k- Model
The realizable k- model is a relatively recent development and diers from the standard
k- model in two important ways:
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 uctuation.
The term realizable means that the model satises certain mathematical constraints
on the Reynolds stresses, consistent with the physics of turbulent ows. Neither the
standard k- model nor the RNG k- model is realizable.
An immediate benet 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
c Fluent Inc. January 11, 2005 11-7
Modeling Turbulence
performance for ows involving rotation, boundary layers under strong adverse pressure
gradients, separation, and recirculation.
Both the realizable and RNG k- models have shown substantial improvements over the
standard k- model where the ow 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 ows and ows with complex
secondary ow features.
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 uid zones (e.g., multiple reference frames, rotating sliding meshes). This is due
to the fact that the realizable k- model includes the eects of mean rotation in the
denition of the turbulent viscosity (see Equations 11.4-1711.4-19). This extra rotation
eect has been tested on single rotating reference frame systems and showed superior be-
havior over the standard k- model. However, due to the nature of this modication, its
application to multiple reference frame systems should be taken with some caution. See
Section 11.4.3: Modeling the Turbulent Viscosity for information about how to include
or exclude this term from the model.
11.2.8 The Standard k- Model
The standard k- model in FLUENT is based on the Wilcox k- model [378], which
incorporates modications for low-Reynolds-number eects, compressibility, and shear
ow spreading. The Wilcox model predicts free shear ow spreading rates that are in
close agreement with measurements for far wakes, mixing layers, and plane, round, and
radial jets, and is thus applicable to wall-bounded ows and free shear ows. A variation
of the standard k- model called the SST k- model is also available in FLUENT, and is
described in Section 11.2.9: The Shear-Stress Transport (SST) k- Model.
11.2.9 The Shear-Stress Transport (SST) k- Model
The shear-stress transport (SST) k- model was developed by Menter [221] to eectively
blend the robust and accurate formulation of the k- model in the near-wall region with
the free-stream independence of the k- model in the far eld. To achieve this, the k-
model is converted into a k- formulation. The SST k- model is similar to the standard
k- model, but includes the following renements:
The standard k- 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- model,
and zero away from the surface, which activates the transformed k- model.
11-8 c Fluent Inc. January 11, 2005
11.2 Choosing a Turbulence Model
The SST model incorporates a damped cross-diusion derivative term in the
equation.
The denition of the turbulent viscosity is modied to account for the transport of
the turbulent shear stress.
The modeling constants are dierent.
These features make the SST k- model more accurate and reliable for a wider class
of ows (e.g., adverse pressure gradient ows, airfoils, transonic shock waves) than the
standard k- model.
11.2.10 The v
2
-f Model
The v
2
-f model is similar to the standard k- model, but incorporates near-wall turbu-
lence anisotropy and non-local pressure-strain eects. A limitation of the v
2
-f model is
that it cannot be used to solve Eulerian multiphase problems, whereas the k- model is
typically used in such applications. The v
2
-f model is a general low-Reynolds-number
turbulence model that is valid all the way up to solid walls, and therefore does not need
to make use of wall functions. Although the model was originally developed for attached
or mildly separated boundary layers [84], it also accurately simulates ows dominated by
separation [27].
The distinguishing feature of the v
2
-f model is its use of the velocity scale, v
2
, instead
of the turbulent kinetic energy, k, for evaluating the eddy viscosity. v
2
, which can be
thought of as the velocity uctuation normal to the streamlines, has shown to provide
the right scaling in representing the damping of turbulent transport close to the wall, a
feature that k does not provide.
For more information about the theoretical background and usage of the v
2
-f model,
please visit the Fluent User Services Center (www.fluentusers.com).
11.2.11 The Reynolds Stress Model (RSM)
The Reynolds stress model (RSM) is the most elaborate turbulence model that FLU-
ENT provides. Abandoning the isotropic eddy-viscosity hypothesis, the RSM closes
the Reynolds-averaged Navier-Stokes equations by solving transport equations for the
Reynolds stresses, together with an equation for the dissipation rate. This means that
ve additional transport equations are required in 2D ows and seven additional trans-
port equations must be solved in 3D.
Since the RSM accounts for the eects of streamline curvature, swirl, rotation, and rapid
changes in strain rate in a more rigorous manner than one-equation and two-equation
models, it has greater potential to give accurate predictions for complex ows. However,
the delity of RSM predictions is still limited by the closure assumptions employed to
c Fluent Inc. January 11, 2005 11-9
Modeling Turbulence
model various terms in the exact transport equations for the Reynolds stresses. The
modeling of the pressure-strain and dissipation-rate terms is particularly challenging, and
often considered to be responsible for compromising the accuracy of RSM predictions.
The RSM might not always yield results that are clearly superior to the simpler models
in all classes of ows to warrant the additional computational expense. However, use
of the RSM is a must when the ow features of interest are the result of anisotropy in
the Reynolds stresses. Among the examples are cyclone ows, highly swirling ows in
combustors, rotating ow passages, and the stress-induced secondary ows in ducts.
11.2.12 Computational Effort: CPU Time and Solution Behavior
In terms of computation, the Spalart-Allmaras model is the least expensive turbulence
model of the options provided in FLUENT, since only one turbulence transport equation
is solved.
The standard k- model clearly requires more computational eort than the Spalart-
Allmaras model since an additional transport equation is solved. The realizable k-
model requires only slightly more computational eort than the standard k- model.
However, due to the extra terms and functions in the governing equations and a greater
degree of non-linearity, computations with the RNG k- model tend to take 1015% more
CPU time than with the standard k- model. Like the k- models, the k- models are
also two-equation models, and thus require about the same computational eort.
Compared with the k- and k- models, the RSM requires additional memory and CPU
time due to the increased number of the transport equations for Reynolds stresses. How-
ever, ecient programming in FLUENT has reduced the CPU time per iteration signi-
cantly. On average, the RSM in FLUENT requires 5060% more CPU time per iteration
compared to the k- and k- models. Furthermore, 1520% more memory is needed.
Aside from the time per iteration, the choice of turbulence model can aect the ability of
FLUENT to obtain a converged solution. For example, the standard k- model is known
to be slightly over-diusive in certain situations, while the RNG k- model is designed
such that the turbulent viscosity is reduced in response to high rates of strain. Since
diusion has a stabilizing eect on the numerics, the RNG model is more likely to be
susceptible to instability in steady-state solutions. However, this should not necessarily
be seen as a disadvantage of the RNG model, since these characteristics make it more
responsive to important physical instabilities such as time-dependent turbulent vortex
shedding.
Similarly, the RSM may take more iterations to converge than the k- and k- models
due to the strong coupling between the Reynolds stresses and the mean ow.
11-10 c Fluent Inc. January 11, 2005
11.3 The Spalart-Allmaras Model
11.3 The Spalart-Allmaras Model
In turbulence models that employ the Boussinesq approach, the central issue is how the
eddy viscosity is computed. The model proposed by Spalart and Allmaras [325] solves
a transport equation for a quantity that is a modied form of the turbulent kinematic
viscosity.
11.3.1 Transport Equation for the Spalart-Allmaras Model
The transported variable in the Spalart-Allmaras model, , is identical to the turbu-
lent kinematic viscosity except in the near-wall (viscous-aected) region. The transport
equation for is

t
( )+

x
i
( u
i
) = G

+
1


_
_

x
j
_
( + )

x
j
_
+ C
b2

_

x
j
_
2
_
_
Y

+S

(11.3-1)
where G

is the production of turbulent viscosity and Y

is the destruction of turbulent


viscosity that occurs in the near-wall region due to wall blocking and viscous damping.


and C
b2
are constants and is the molecular kinematic viscosity. S

is a user-dened
source term. Note that since the turbulence kinetic energy k is not calculated in the
Spalart-Allmaras model, the last term in Equation 11.2-5 is ignored when estimating the
Reynolds stresses.
11.3.2 Modeling the Turbulent Viscosity
The turbulent viscosity,
t
, is computed from

t
= f
v1
(11.3-2)
where the viscous damping function, f
v1
, is given by
f
v1
=

3

3
+ C
3
v1
(11.3-3)
and

(11.3-4)
c Fluent Inc. January 11, 2005 11-11
Modeling Turbulence
11.3.3 Modeling the Turbulent Production
The production term, G

, is modeled as
G

= C
b1

S (11.3-5)
where

S S +

2
d
2
f
v2
(11.3-6)
and
f
v2
= 1

1 + f
v1
(11.3-7)
C
b1
and are constants, d is the distance from the wall, and S is a scalar measure of the
deformation tensor. By default in FLUENT, as in the original model proposed by Spalart
and Allmaras, S is based on the magnitude of the vorticity:
S
_
2
ij

ij
(11.3-8)
where
ij
is the mean rate-of-rotation tensor and is dened by

ij
=
1
2
_
u
i
x
j

u
j
x
i
_
(11.3-9)
The justication for the default expression for S is that, for the wall-bounded ows that
were of most interest when the model was formulated, turbulence is found only where
vorticity is generated near walls. However, it has since been acknowledged that one
should also take into account the eect of mean strain on the turbulence production, and
a modication to the model has been proposed [67] and incorporated into FLUENT.
This modication combines measures of both rotation and strain tensors in the denition
of S:
S |
ij
| + C
prod
min (0, |S
ij
| |
ij
|) (11.3-10)
where
C
prod
= 2.0, |
ij
|
_
2
ij

ij
, |S
ij
|
_
2S
ij
S
ij
11-12 c Fluent Inc. January 11, 2005
11.3 The Spalart-Allmaras Model
with the mean strain rate, S
ij
, dened as
S
ij
=
1
2
_
u
j
x
i
+
u
i
x
j
_
(11.3-11)
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 ows, i.e., ow
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 eects 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 modied form for calculating production in the Viscous Model panel.
11.3.4 Modeling the Turbulent Destruction
The destruction term is modeled as
Y

= C
w1
f
w
_

d
_
2
(11.3-12)
where
f
w
= g
_
1 + C
6
w3
g
6
+ C
6
w3
_
1/6
(11.3-13)
g = r + C
w2
_
r
6
r
_
(11.3-14)
r

S
2
d
2
(11.3-15)
C
w1
, C
w2
, and C
w3
are constants, and

S is given by Equation 11.3-6. Note that the
modication described above to include the eects of mean strain on S will also aect
the value of

S used to compute r.
c Fluent Inc. January 11, 2005 11-13
Modeling Turbulence
11.3.5 Model Constants
The model constants C
b1
, C
b2
,

, C
v1
, C
w1
, C
w2
, C
w3
, and have the following default
values [325]:
C
b1
= 0.1355, C
b2
= 0.622,

=
2
3
, C
v1
= 7.1
C
w1
=
C
b1

2
+
(1 + C
b2
)


, C
w2
= 0.3, C
w3
= 2.0, = 0.4187
11.3.6 Wall Boundary Conditions
At walls, the modied turbulent kinematic viscosity, , is set to zero.
When the mesh is ne enough to resolve the laminar sublayer, the wall shear stress is
obtained from the laminar stress-strain relationship:
u
u

=
u

(11.3-16)
If the mesh is too coarse to resolve the laminar sublayer, 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
u

=
1

ln E
_
u

_
(11.3-17)
where u is the velocity parallel to the wall, u

is the shear velocity, y is the distance from


the wall, is the von Karm an constant (0.4187), and E = 9.793.
11.3.7 Convective Heat and Mass Transfer Modeling
In 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
(E) +

x
i
[u
i
(E + p)] =

x
j
_
_
k +
c
p

t
Pr
t
_
T
x
j
+ u
i
(
ij
)
e
_
+ S
h
(11.3-18)
where k, in this case, is the thermal conductivity, E is the total energy, and (
ij
)
e
is the
deviatoric stress tensor, dened as
11-14 c Fluent Inc. January 11, 2005
11.4 The Standard, RNG, and Realizable k- Models
(
ij
)
e
=
e
_
u
j
x
i
+
u
i
x
j
_

2
3

e
u
i
x
i

ij
The term involving (
ij
)
e
represents the viscous heating, and is always computed in the
coupled solvers. It is not computed by default in the segregated solver, but it can be
enabled in the Viscous Model panel. The default value of the turbulent Prandtl number
is 0.85. You can change the value of Pr
t
in the Viscous Model panel.
Turbulent mass transfer is treated similarly, with a default turbulent Schmidt number of
0.7. This default value can be changed in the Viscous Model panel.
Wall boundary conditions for scalar transport are handled analogously to momentum,
using the appropriate law-of-the-wall.
11.4 The Standard, RNG, and Realizable k- Models
This section presents the standard, RNG, and realizable k- models. All three models
have similar forms, with transport equations for k and . The major dierences in the
models are as follows:
the method of calculating turbulent viscosity
the turbulent Prandtl numbers governing the turbulent diusion of k and
the generation and destruction terms in the equation
The transport equations, methods of calculating turbulent viscosity, and model constants
are presented separately for each model. The features that are essentially common to all
models follow, including turbulent production, generation due to buoyancy, accounting
for the eects of compressibility, and modeling heat and mass transfer.
11.4.1 The Standard k- Model
The standard k- model [183] is a semi-empirical model based on model transport equa-
tions for the turbulence kinetic energy (k) and its dissipation rate (). The model trans-
port equation for k is derived from the exact equation, while the model transport equation
for was obtained using physical reasoning and bears little resemblance to its mathe-
matically exact counterpart.
In the derivation of the k- model, it was assumed that the ow is fully turbulent, and
the eects of molecular viscosity are negligible. The standard k- model is therefore valid
only for fully turbulent ows.
c Fluent Inc. January 11, 2005 11-15
Modeling Turbulence
Transport Equations for the Standard k- Model
The turbulence kinetic energy, k, and its rate of dissipation, , are obtained from the
following transport equations:

t
(k) +

x
i
(ku
i
) =

x
j
_
_
+

t

k
_
k
x
j
_
+ G
k
+ G
b
Y
M
+ S
k
(11.4-1)
and

t
() +

x
i
(u
i
) =

x
j
_
_
+

t

_

x
j
_
+C
1

k
(G
k
+ C
3
G
b
) C
2

2
k
+S

(11.4-2)
In these equations, G
k
represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 11.4.4: Modeling Turbulent
Production in the k- Models. G
b
is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 11.4.5: Eects of Buoyancy on Turbu-
lence in the k- Models. Y
M
represents the contribution of the uctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Sec-
tion 11.4.6: Eects of Compressibility on Turbulence in the k- Models. C
1
, C
2
, and C
3
are constants.
k
and

are the turbulent Prandtl numbers for k and , respectively. S


k
and S

are user-dened source terms.


Modeling the Turbulent Viscosity
The turbulent (or eddy) viscosity,
t
, is computed by combining k and as follows:

t
= C

k
2

(11.4-3)
where C

is a constant.
Model Constants
The model constants C
1
, C
2
, C

,
k
, and

have the following default values [183]:


C
1
= 1.44, C
2
= 1.92, C

= 0.09,
k
= 1.0,

= 1.3
These default values have been determined from experiments with air and water for funda-
mental turbulent shear ows including homogeneous shear ows and decaying isotropic
grid turbulence. They have been found to work fairly well for a wide range of wall-
bounded and free shear ows.
11-16 c Fluent Inc. January 11, 2005
11.4 The Standard, RNG, and Realizable k- Models
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 panel.
11.4.2 The RNG 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 dierent 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 [54].
Transport Equations for the RNG k- Model
The RNG k- model has a similar form to the standard k- model:

t
(k) +

x
i
(ku
i
) =

x
j
_

e
k
x
j
_
+ G
k
+ G
b
Y
M
+ S
k
(11.4-4)
and

t
() +

x
i
(u
i
) =

x
j
_

x
j
_
+C
1

k
(G
k
+ C
3
G
b
) C
2

2
k
R

+S

(11.4-5)
In these equations, G
k
represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 11.4.4: Modeling Turbulent
Production in the k- Models. G
b
is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 11.4.5: Eects of Buoyancy on Turbu-
lence in the k- Models. Y
M
represents the contribution of the uctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Sec-
tion 11.4.6: Eects of Compressibility on Turbulence in the k- Models. The quantities

k
and

are the inverse eective Prandtl numbers for k and , respectively. S


k
and S

are user-dened source terms.


Modeling the Effective Viscosity
The scale elimination procedure in RNG theory results in a dierential equation for
turbulent viscosity:
d
_

2
k

_
= 1.72


3
1 + C

d (11.4-6)
c Fluent Inc. January 11, 2005 11-17
Modeling Turbulence
where
=
e
/
C

100
Equation 11.4-6 is integrated to obtain an accurate description of how the eective tur-
bulent transport varies with the eective Reynolds number (or eddy scale), allowing the
model to better handle low-Reynolds-number and near-wall ows.
In the high-Reynolds-number limit, Equation 11.4-6 gives

t
= C

k
2

(11.4-7)
with C

= 0.0845, derived using RNG theory. It is interesting to note that this value
of C

is very close to the empirically-determined value of 0.09 used in the standard k-


model.
In FLUENT, by default, the eective viscosity is computed using the high-Reynolds-
number form in Equation 11.4-7. However, there is an option available that allows you
to use the dierential relation given in Equation 11.4-6 when you need to include low-
Reynolds-number eects.
RNG Swirl Modication
Turbulence, in general, is aected by rotation or swirl in the mean ow. The RNG model
in FLUENT provides an option to account for the eects of swirl or rotation by modifying
the turbulent viscosity appropriately. The modication takes the following functional
form:

t
=
t0
f
_

s
, ,
k

_
(11.4-8)
where
t0
is the value of turbulent viscosity calculated without the swirl modication
using either Equation 11.4-6 or Equation 11.4-7. is a characteristic swirl number eval-
uated within FLUENT, and
s
is a swirl constant that assumes dierent values depending
on whether the ow is swirl-dominated or only mildly swirling. This swirl modication
always takes eect for axisymmetric, swirling ows and three-dimensional ows when the
RNG model is selected. For mildly swirling ows (the default in FLUENT),
s
is set to
0.05 and cannot be modied. For strongly swirling ows, however, a higher value of
s
can be used.
11-18 c Fluent Inc. January 11, 2005
11.4 The Standard, RNG, and Realizable k- Models
Calculating the Inverse Effective Prandtl Numbers
The inverse eective Prandtl numbers,
k
and

, are computed using the following


formula derived analytically by the RNG theory:

1.3929

0
1.3929

0.6321

+ 2.3929

0
+ 2.3929

0.3679
=

mol

e
(11.4-9)
where
0
= 1.0. In the high-Reynolds-number limit (
mol
/
e
1),
k
=

1.393.
The R

Term in the Equation


The main dierence between the RNG and standard k- models lies in the additional
term in the equation given by
R

=
C

3
(1 /
0
)
1 +
3

2
k
(11.4-10)
where Sk/,
0
= 4.38, = 0.012.
The eects of this term in the RNG equation can be seen more clearly by rearranging
Equation 11.4-5. Using Equation 11.4-10, the third and fourth terms on the right-hand
side of Equation 11.4-5 can be merged, and the resulting equation can be rewritten as

t
() +

x
i
(u
i
) =

x
j
_

x
j
_
+ C
1

k
(G
k
+ C
3
G
b
) C

2
k
(11.4-11)
where C

2
is given by
C

2
C
2
+
C

3
(1 /
0
)
1 +
3
(11.4-12)
In regions where <
0
, the R term makes a positive contribution, and C

2
becomes
larger than C
2
. In the logarithmic layer, for instance, it can be shown that 3.0,
giving C

2
2.0, which is close in magnitude to the value of C
2
in the standard k-
model (1.92). As a result, for weakly to moderately strained ows, the RNG model tends
to give results largely comparable to the standard k- model.
In regions of large strain rate ( >
0
), however, the R term makes a negative contribu-
tion, making the value of C

2
less than C
2
. In comparison with the standard k- model,
the smaller destruction of augments , reducing k and, eventually, the eective viscosity.
As a result, in rapidly strained ows, the RNG model yields a lower turbulent viscosity
than the standard k- model.
c Fluent Inc. January 11, 2005 11-19
Modeling Turbulence
Thus, the RNG model is more responsive to the eects of rapid strain and streamline
curvature than the standard k- model, which explains the superior performance of the
RNG model for certain classes of ows.
Model Constants
The model constants C
1
and C
2
in Equation 11.4-5 have values derived analytically by
the RNG theory. These values, used by default in FLUENT, are
C
1
= 1.42, C
2
= 1.68
11.4.3 The Realizable k- Model
In addition to the standard and RNG-based k- models described in Sections 11.4.1 and
11.4.2, FLUENT also provides the so-called realizable k- model [306]. The term real-
izable means that the model satises certain mathematical constraints on the normal
stresses, consistent with the physics of turbulent ows. To understand this, consider
combining the Boussinesq relationship (Equation 11.2-5) and the eddy viscosity deni-
tion (Equation 11.4-3) to obtain the following expression for the normal Reynolds stress
in an incompressible strained mean ow:
u
2
=
2
3
k 2
t
U
x
(11.4-13)
Using Equation 11.4-3 for
t

t
/, one obtains the result that the normal stress, u
2
,
which by denition is a positive quantity, becomes negative, i.e., non-realizable, when
the strain is large enough to satisfy
k

U
x
>
1
3C

3.7 (11.4-14)
Similarly, it can also be shown that the Schwarz inequality for shear stresses (u

u
2

u
2

; no summation over and ) 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 C

variable by sensitizing it to
the mean ow (mean deformation) and the turbulence (k, ). The notion of variable
C

is suggested by many modelers including Reynolds [279], and is well substantiated


by experimental evidence. For example, C

is found to be around 0.09 in the inertial


sublayer of equilibrium boundary layers, and 0.05 in a strong homogeneous shear ow.
Another weakness 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 nding that the spreading rate in planar jets is predicted reasonably well, but
11-20 c Fluent Inc. January 11, 2005
11.4 The Standard, RNG, and Realizable k- Models
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. [306] was intended to address these
deciencies of traditional k- models by adopting the following:
a new eddy-viscosity formula involving a variable C

originally proposed by
Reynolds [279].
a new model equation for dissipation () based on the dynamic equation of the
mean-square vorticity uctuation.
Transport Equations for the Realizable k- Model
The modeled transport equations for k and in the realizable k- model are

t
(k) +

x
j
(ku
j
) =

x
j
_
_
+

t

k
_
k
x
j
_
+ G
k
+ G
b
Y
M
+ S
k
(11.4-15)
and

t
() +

x
j
(u
j
) =

x
j
_
_
+

t

_

x
j
_
+ C
1
S C
2

2
k +

+ C
1

k
C
3
G
b
+ S

(11.4-16)
where
C
1
= max
_
0.43,

+ 5
_
, = S
k

, S =
_
2S
ij
S
ij
In these equations, G
k
represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 11.4.4: Modeling Turbulent
Production in the k- Models. G
b
is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 11.4.5: Eects of Buoyancy on Turbu-
lence in the k- Models. Y
M
represents the contribution of the uctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Sec-
tion 11.4.6: Eects of Compressibility on Turbulence in the k- Models. C
2
and C
1
are
constants.
k
and

are the turbulent Prandtl numbers for k and , respectively. S


k
and
S

are user-dened source terms.


Note that the k equation (Equation 11.4-15) is the same as that in the standard k-
model (Equation 11.4-1) and the RNG k- model (Equation 11.4-4), except for the
c Fluent Inc. January 11, 2005 11-21
Modeling Turbulence
model constants. However, the form of the equation is quite dierent from those in
the standard and RNG-based k- models (Equations 11.4-2 and 11.4-5). One of the
noteworthy features is that the production term in the equation (the second term on
the right-hand side of Equation 11.4-16) does not involve the production of k; i.e., it does
not contain the same G
k
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 11.4-16)
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 ows [171, 306], including
rotating homogeneous shear ows, free ows including jets and mixing layers, channel
and boundary layer ows, and separated ows. 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.
Modeling the Turbulent Viscosity
As in other k- models, the eddy viscosity is computed from

t
= C

k
2

(11.4-17)
The dierence between the realizable k- model and the standard and RNG k- models
is that C

is no longer constant. It is computed from


C

=
1
A
0
+ A
s
kU

(11.4-18)
where
U


_
S
ij
S
ij
+

ij

ij
(11.4-19)
and

ij
=
ij
2
ijk

ij
=
ij

ijk

k
11-22 c Fluent Inc. January 11, 2005
11.4 The Standard, RNG, and Realizable k- Models
where
ij
is the mean rate-of-rotation tensor viewed in a rotating reference frame with
the angular velocity
k
. The model constants A
0
and A
s
are given by
A
0
= 4.04, A
s
=

6 cos
where
=
1
3
cos
1
(

6W), W =
S
ij
S
jk
S
ki

S
3
,

S =
_
S
ij
S
ij
, S
ij
=
1
2
_
u
j
x
i
+
u
i
x
j
_
It can be seen that C

is a function of the mean strain and rotation rates, the angular ve-
locity of the system rotation, and the turbulence elds (k and ). C

in Equation 11.4-17
can be shown to recover the standard value of 0.09 for an inertial sublayer in an equilib-
rium boundary layer.
i
In FLUENT, the term 2
ijk

k
is, by default, not included in the
calculation of

ij
. This is an extra rotation term that is not com-
patible with cases involving sliding meshes or multiple reference frames.
If you want to include this term in the model, you can enable it by using the
define/models/viscous/turbulence-expert/rke-cmu-rotation-term?
text command and entering yes at the prompt.
Model Constants
The model constants C
2
,
k
, and

have been established to ensure that the model


performs well for certain canonical ows. The model constants are
C
1
= 1.44, C
2
= 1.9,
k
= 1.0,

= 1.2
11.4.4 Modeling Turbulent Production in the k- Models
The term G
k
, representing the production of turbulence kinetic energy, is modeled iden-
tically for the standard, RNG, and realizable k- models. From the exact equation for
the transport of k, this term may be dened as
G
k
= u

i
u

j
u
j
x
i
(11.4-20)
To evaluate G
k
in a manner consistent with the Boussinesq hypothesis,
G
k
=
t
S
2
(11.4-21)
c Fluent Inc. January 11, 2005 11-23
Modeling Turbulence
where S is the modulus of the mean rate-of-strain tensor, dened as
S
_
2S
ij
S
ij
(11.4-22)
i
When using the high-Reynolds number k- versions,
e
is used in lieu of

t
in Equation 11.4-21.
11.4.5 Effects of Buoyancy on Turbulence in the k- Models
When a non-zero gravity eld and temperature gradient are present simultaneously, the
k- models in FLUENT account for the generation of k due to buoyancy (G
b
in Equa-
tions 11.4-1, 11.4-4, and 11.4-15), and the corresponding contribution to the production
of in Equations 11.4-2, 11.4-5, and 11.4-16.
The generation of turbulence due to buoyancy is given by
G
b
= g
i

t
Pr
t
T
x
i
(11.4-23)
where Pr
t
is the turbulent Prandtl number for energy and g
i
is the component of the
gravitational vector in the ith direction. For the standard and realizable k- models, the
default value of Pr
t
is 0.85. In the case of the RNG k- model, Pr
t
= 1/, where
is given by Equation 11.4-9, but with
0
= 1/Pr = k/c
p
. The coecient of thermal
expansion, , is dened as
=
1

T
_
p
(11.4-24)
For ideal gases, Equation 11.4-23 reduces to
G
b
= g
i

t
Pr
t

x
i
(11.4-25)
It can be seen from the transport equations for k (Equations 11.4-1, 11.4-4, and 11.4-15)
that turbulence kinetic energy tends to be augmented (G
b
> 0) in unstable stratication.
For stable stratication, buoyancy tends to suppress the turbulence (G
b
< 0). In FLU-
ENT, the eects of buoyancy on the generation of k are always included when you have
both a non-zero gravity eld and a non-zero temperature (or density) gradient.
While the buoyancy eects on the generation of k are relatively well understood, the
eect on is less clear. In FLUENT, by default, the buoyancy eects on are neglected
simply by setting G
b
to zero in the transport equation for (Equation 11.4-2, 11.4-5, or
11.4-16).
11-24 c Fluent Inc. January 11, 2005
11.4 The Standard, RNG, and Realizable k- Models
However, you can include the buoyancy eects on in the Viscous Model panel. In this
case, the value of G
b
given by Equation 11.4-25 is used in the transport equation for
(Equation 11.4-2, 11.4-5, or 11.4-16).
The degree to which is aected by the buoyancy is determined by the constant C
3
.
In FLUENT, C
3
is not specied, but is instead calculated according to the following
relation [129]:
C
3
= tanh

v
u

(11.4-26)
where v is the component of the ow velocity parallel to the gravitational vector and
u is the component of the ow velocity perpendicular to the gravitational vector. In
this way, C
3
will become 1 for buoyant shear layers for which the main ow direction is
aligned with the direction of gravity. For buoyant shear layers that are perpendicular to
the gravitational vector, C
3
will become zero.
11.4.6 Effects of Compressibility on Turbulence in the k- Models
For high-Mach-number ows, compressibility aects turbulence through so-called di-
latation dissipation, which is normally neglected in the modeling of incompressible
ows [378]. Neglecting the dilatation dissipation fails to predict the observed decrease
in spreading rate with increasing Mach number for compressible mixing and other free
shear layers. To account for these eects in the k- models in FLUENT, the dilatation
dissipation term, Y
M
, is included in the k equation. This term is modeled according to
a proposal by Sarkar [291]:
Y
M
= 2M
2
t
(11.4-27)
where M
t
is the turbulent Mach number, dened as
M
t
=

k
a
2
(11.4-28)
where a (

RT) is the speed of sound.


This compressibility modication always takes eect when the compressible form of the
ideal gas law is used.
c Fluent Inc. January 11, 2005 11-25
Modeling Turbulence
11.4.7 Convective Heat and Mass Transfer Modeling in the k- Models
In 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
(E) +

x
i
[u
i
(E + p)] =

x
j
_
k
e
T
x
j
+ u
i
(
ij
)
e
_
+ S
h
(11.4-29)
where E is the total energy, k
e
is the eective thermal conductivity, and
(
ij
)
e
is the deviatoric stress tensor, dened as
(
ij
)
e
=
e
_
u
j
x
i
+
u
i
x
j
_

2
3

e
u
i
x
i

ij
The term involving (
ij
)
e
represents the viscous heating, and is always computed in the
coupled solvers. It is not computed by default in the segregated solver, but it can be
enabled in the Viscous Model panel.
Additional terms may appear in the energy equation, depending on the physical models
you are using. See Section 12.2.1: Theory for more details.
For the standard and realizable k- models, the eective thermal conductivity is given
by
k
e
= k +
c
p

t
Pr
t
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 panel.
For the RNG k- model, the eective thermal conductivity is
k
e
= c
p

e
where is calculated from Equation 11.4-9, but with
0
= 1/Pr = k/c
p
.
The fact that varies with
mol
/
e
, as in Equation 11.4-9, is an advantage of the RNG k-
model. It is consistent with experimental evidence indicating that the turbulent Prandtl
number varies with the molecular Prandtl number and turbulence [163]. Equation 11.4-9
works well across a very broad range of molecular Prandtl numbers, from liquid metals
(Pr 10
2
) to paran oils (Pr 10
3
), which allows heat transfer to be calculated in low-
Reynolds-number regions. Equation 11.4-9 smoothly predicts the variation of eective
11-26 c Fluent Inc. January 11, 2005
11.5 The Standard and SST k- Models
Prandtl number from the molecular value ( = 1/Pr) in the viscosity-dominated region
to the fully turbulent value ( = 1.393) in the fully turbulent regions of the ow.
Turbulent mass transfer is treated similarly. For the standard and realizable k- models,
the default turbulent Schmidt number is 0.7. This default value can be changed in the
Viscous Model panel. For the RNG model, the eective turbulent diusivity for mass
transfer is calculated in a manner that is analogous to the method used for the heat
transport. The value of
0
in Equation 11.4-9 is
0
= 1/Sc, where Sc is the molecular
Schmidt number.
11.5 The Standard and SST k- Models
This section presents the standard and shear-stress transport (SST) k- models. Both
models have similar forms, with transport equations for k and . The major ways in
which the SST model [222] diers from the standard model are as follows:
gradual change from the standard k- 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
modied turbulent viscosity formulation to account for the transport eects of the
principal turbulent shear stress
The transport equations, methods of calculating turbulent viscosity, and methods of
calculating model constants and other terms are presented separately for each model.
11.5.1 The Standard k- Model
The standard k- model is an empirical model based on model transport equations for
the turbulence kinetic energy (k) and the specic dissipation rate (), which can also be
thought of as the ratio of to k [378].
As the k- model has been modied over the years, production terms have been added
to both the k and equations, which have improved the accuracy of the model for
predicting free shear ows.
Transport Equations for the Standard k- Model
The turbulence kinetic energy, k, and the specic dissipation rate, , are obtained from
the following transport equations:

t
(k) +

x
i
(ku
i
) =

x
j
_

k
k
x
j
_
+ G
k
Y
k
+ S
k
(11.5-1)
c Fluent Inc. January 11, 2005 11-27
Modeling Turbulence
and

t
() +

x
i
(u
i
) =

x
j
_

x
j
_
+ G

+ S

(11.5-2)
In these equations, G
k
represents the generation of turbulence kinetic energy due to mean
velocity gradients. G

represents the generation of .


k
and

represent the eective


diusivity of k and , respectively. Y
k
and Y

represent the dissipation of k and due


to turbulence. All of the above terms are calculated as described below. S
k
and S

are
user-dened source terms.
Modeling the Effective Diffusivity
The eective diusivities for the k- model are given by

k
= +

t

k
(11.5-3)

= +

t

(11.5-4)
where
k
and

are the turbulent Prandtl numbers for k and , respectively. The


turbulent viscosity,
t
, is computed by combining k and as follows:

t
=

(11.5-5)
Low-Reynolds-Number Correction
The coecient

damps the turbulent viscosity causing a low-Reynolds-number correc-


tion. It is given by

0
+ Re
t
/R
k
1 + Re
t
/R
k
_
(11.5-6)
where
Re
t
=
k

(11.5-7)
R
k
= 6 (11.5-8)

0
=

i
3
(11.5-9)

i
= 0.072 (11.5-10)
11-28 c Fluent Inc. January 11, 2005
11.5 The Standard and SST k- Models
Note that, in the high-Reynolds-number form of the k- model,

= 1.
Modeling the Turbulence Production
Production of k
The term G
k
represents the production of turbulence kinetic energy. From the exact
equation for the transport of k, this term may be dened as
G
k
= u

i
u

j
u
j
x
i
(11.5-11)
To evaluate G
k
in a manner consistent with the Boussinesq hypothesis,
G
k
=
t
S
2
(11.5-12)
where S is the modulus of the mean rate-of-strain tensor, dened in the same way as for
the k- model (see Equation 11.4-22).
Production of
The production of is given by
G

k
G
k
(11.5-13)
where G
k
is given by Equation 11.5-11.
The coecient is given by
=

0
+ Re
t
/R

1 + Re
t
/R

_
(11.5-14)
where R

= 2.95.

and Re
t
are given by Equations 11.5-6 and 11.5-7, respectively.
Note that, in the high-Reynolds-number form of the k- model, =

= 1.
Modeling the Turbulence Dissipation
Dissipation of k
The dissipation of k is given by
Y
k
=

k (11.5-15)
c Fluent Inc. January 11, 2005 11-29
Modeling Turbulence
where
f

=
_
_
_
1
k
0
1+680
2
k
1+400
2
k

k
> 0
(11.5-16)
where

k

1

3
k
x
j

x
j
(11.5-17)
and

i
[1 +

F(M
t
)] (11.5-18)

i
=

_
4/15 + (Re
t
/R

)
4
1 + (Re
t
/R

)
4
_
(11.5-19)

= 1.5 (11.5-20)
R

= 8 (11.5-21)

= 0.09 (11.5-22)
where Re
t
is given by Equation 11.5-7.
Dissipation of
The dissipation of is given by
Y

= f


2
(11.5-23)
where
f

=
1 + 70

1 + 80

(11.5-24)

ij

jk
S
ki
(

)
3

(11.5-25)

ij
=
1
2
_
u
i
x
j

u
j
x
i
_
(11.5-26)
11-30 c Fluent Inc. January 11, 2005
11.5 The Standard and SST k- Models
The strain rate tensor, S
ij
is dened in Equation 11.3-11. Also,
=
i
_
1

F(M
t
)
_
(11.5-27)

i
and F(M
t
) are dened by Equations 11.5-19 and 11.5-28, respectively.
Compressibility Correction
The compressibility function, F(M
t
), is given by
F(M
t
) =
_
0 M
t
M
t0
M
2
t
M
2
t0
M
t
> M
t0
(11.5-28)
where
M
2
t

2k
a
2
(11.5-29)
M
t0
= 0.25 (11.5-30)
a =
_
RT (11.5-31)
Note that, in the high-Reynolds-number form of the k- model,

i
=

. In the incom-
pressible form,

i
.
Model Constants

= 1,

= 0.52,
0
=
1
9
,

= 0.09,
i
= 0.072, R

= 8
R
k
= 6, R

= 2.95,

= 1.5, M
t0
= 0.25,
k
= 2.0,

= 2.0
Wall Boundary Conditions
The wall boundary conditions for the k equation in the k- models are treated in the
same way as the k equation is treated when enhanced wall treatments are used with
the k- models. This means that all boundary conditions for wall-function meshes will
correspond to the wall function approach, while for the ne meshes, the appropriate
low-Reynolds-number boundary conditions will be applied.
In FLUENT the value of at the wall is specied as

w
=
(u

)
2


+
(11.5-32)
c Fluent Inc. January 11, 2005 11-31
Modeling Turbulence
The asymptotic value of
+
in the laminar sublayer is given by

+
= min
_

+
w
,
6

i
(y
+
)
2
_
(11.5-33)
where

+
w
=
_

_
_
50
k
+
s
_
2
k
+
s
< 25
100
k
+
s
k
+
s
25
(11.5-34)
where
k
+
s
= max
_
1.0,
k
s
u

_
(11.5-35)
and k
s
is the roughness height.
In the logarithmic (or turbulent) region, the value of
+
is

+
=
1
_

du
+
turb
dy
+
(11.5-36)
which leads to the value of in the wall cell as
=
u

y
(11.5-37)
Note that in the case of a wall cell being placed in the buer region, FLUENT will blend

+
between the logarithmic and laminar sublayer values.
11.5.2 The Shear-Stress Transport (SST) k- Model
In addition to the standard k- model described in Section 11.5.1: The Standard k-
Model, FLUENT also provides a variation called the shear-stress transport (SST) k-
model, so named because the denition of the turbulent viscosity is modied to account
for the transport of the principal turbulent shear stress. It is this feature that gives the
SST k- model an advantage in terms of performance over both the standard k- model
and the standard k- model. Other modications include the addition of a cross-diusion
term in the equation and a blending function to ensure that the model equations behave
appropriately in both the near-wall and far-eld zones.
11-32 c Fluent Inc. January 11, 2005
11.5 The Standard and SST k- Models
Transport Equations for the SST k- Model
The SST k- model has a similar form to the standard k- model:

t
(k) +

x
i
(ku
i
) =

x
j
_

k
k
x
j
_
+

G
k
Y
k
+ S
k
(11.5-38)
and

t
() +

x
i
(u
i
) =

x
j
_

x
j
_
+ G

+ D

+ S

(11.5-39)
In these equations,

G
k
represents the generation of turbulence kinetic energy due to
mean velocity gradients, calculated as described in Section 11.5.1: Modeling the Tur-
bulence Production. G

represents the generation of , calculated as described in Sec-


tion 11.5.1: Modeling the Turbulence Production.
k
and

represent the eective


diusivity of k and , respectively, which are calculated as described below. Y
k
and
Y

represent the dissipation of k and due to turbulence, calculated as described in


Section 11.5.1: Modeling the Turbulence Dissipation. D

represents the cross-diusion


term, calculated as described below. S
k
and S

are user-dened source terms.


Modeling the Effective Diffusivity
The eective diusivities for the SST k- model are given by

k
= +

t

k
(11.5-40)

= +

t

(11.5-41)
where
k
and

are the turbulent Prandtl numbers for k and , respectively. The


turbulent viscosity,
t
, is computed as follows:

t
=
k

1
max
_
1

,
SF
2
a
1

_
(11.5-42)
where S is the strain rate magnitude and

k
=
1
F
1
/
k,1
+ (1 F
1
)/
k,2
(11.5-43)

=
1
F
1
/
,1
+ (1 F
1
)/
,2
(11.5-44)
c Fluent Inc. January 11, 2005 11-33
Modeling Turbulence

is dened in Equation 11.5-6. The blending functions, F


1
and F
2
, are given by
F
1
= tanh
_

4
1
_
(11.5-45)

1
= min
_
max
_
k
0.09y
,
500
y
2

_
,
4k

,2
D
+

y
2
_
(11.5-46)
D
+

= max
_
2
1

,2
1

k
x
j

x
j
, 10
10
_
(11.5-47)
F
2
= tanh
_

2
2
_
(11.5-48)

2
= max
_
2

k
0.09y
,
500
y
2

_
(11.5-49)
where y is the distance to the next surface and D
+

is the positive portion of the cross-


diusion term (see Equation 11.5-58).
Modeling the Turbulence Production
Production of k
The term

G
k
represents the production of turbulence kinetic energy, and is dened as:

G
k
= min(G
k
, 10

k) (11.5-50)
where G
k
is dened in the same manner as in the standard k- model. See Sec-
tion 11.5.1: Modeling the Turbulence Production for details.
Production of
The term G

represents the production of and is given by


G

t
G
k
(11.5-51)
Note that this formulation diers from the standard k- model. The dierence between
the two models also exists in the way the term

is evaluated. In the standard k-


model,

is dened as a constant (0.52). For the SST k- model,

is given by

= F
1

,1
+ (1 F
1
)
,2
(11.5-52)
11-34 c Fluent Inc. January 11, 2005
11.5 The Standard and SST k- Models
where

,1
=

i,1

w,1
_

(11.5-53)

,2
=

i,2

w,2
_

(11.5-54)
where is 0.41.
Modeling the Turbulence Dissipation
Dissipation of k
The term Y
k
represents the dissipation of turbulence kinetic energy, and is dened in a
similar manner as in the standard k- model (see Section 11.5.1: Modeling the Turbulence
Dissipation). The dierence is in the way the term f

is evaluated. In the standard k-


model, f

is dened as a piecewise function. For the SST k- model, f

is a constant
equal to 1. Thus,
Y
k
=

k (11.5-55)
Dissipation of
The term Y

represents the dissipation of , and is dened in a similar manner as in


the standard k- model (see Section 11.5.1: Modeling the Turbulence Dissipation). The
dierence is in the way the terms
i
and f

are evaluated. In the standard k- model,


i
is dened as a constant (0.072) and f

is dened in Equation 11.5-24. For the SST k-


model, f

is a constant equal to 1. Thus,


Y
k
=
2
(11.5-56)
Instead of a having a constant value,
i
is given by

i
= F
1

i,1
+ (1 F
1
)
i,2
(11.5-57)
and F
1
is obtained from Equation 11.5-45.
c Fluent Inc. January 11, 2005 11-35
Modeling Turbulence
Cross-Diffusion Modication
The SST k- model is based on both the standard k- 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 , which leads to the introduction of a cross-diusion term
(D

in Equation 11.5-39). D

is dened as
D

= 2 (1 F
1
)
,2
1

k
x
j

x
j
(11.5-58)
For details about the various k- models, see Section 11.4: The Standard, RNG, and
Realizable k- Models.
Model Constants

k,1
= 1.176,
,1
= 2.0,
k,2
= 1.0,
,2
= 1.168
a
1
= 0.31,
i,1
= 0.075
i,2
= 0.0828
All additional model constants (

,
0
,

, R

, R
k
, R

, and M
t0
) have the same
values as for the standard k- model (see Section 11.5.1: Model Constants).
11.6 The Reynolds Stress Model (RSM)
The Reynolds stress model [110, 180, 181] involves calculation of the individual Reynolds
stresses, u

i
u

j
, using dierential transport equations. The individual Reynolds stresses
are then used to obtain closure of the Reynolds-averaged momentum equation (Equa-
tion 11.2-4).
The exact form of the Reynolds stress transport equations may be derived by taking mo-
ments of the exact momentum equation. This is a process wherein the exact momentum
equations are multiplied by a uctuating property, the product then being Reynolds-
averaged. Unfortunately, several of the terms in the exact equation are unknown and
modeling assumptions are required in order to close the equations.
In this section, the Reynolds stress transport equations are presented together with the
modeling assumptions required to attain closure.
11-36 c Fluent Inc. January 11, 2005
11.6 The Reynolds Stress Model (RSM)
11.6.1 The Reynolds Stress Transport Equations
The exact transport equations for the transport of the Reynolds stresses, u

i
u

j
, may be
written as follows:

t
( u

i
u

j
)
. .
Local Time Derivative
+

x
k
(u
k
u

i
u

j
)
. .
C
ij
Convection
=

x
k
_
u

i
u

j
u

k
+ p
_

kj
u

i
+
ik
u

j
_
_
. .
D
T,ij
Turbulent Diusion
+

x
k
_


x
k
(u

i
u

j
)
_
. .
D
L,ij
Molecular Diusion

_
u

i
u

k
u
j
x
k
+ u

j
u

k
u
i
x
k
_
. .
P
ij
Stress Production
(g
i
u

j
+ g
j
u

i
)
. .
G
ij
Buoyancy Production
+ p
_
u

i
x
j
+
u

j
x
i
_
. .

ij
Pressure Strain
2
u

i
x
k
u

j
x
k
. .

ij
Dissipation
2
k
_
u

j
u

ikm
+ u

i
u

jkm
_
. .
F
ij
Production by System Rotation
+ S
user
. .
User-Dened Source Term
(11.6-1)
Of the various terms in these exact equations, C
ij
, D
L,ij
, P
ij
, and F
ij
do not require any
modeling. However, D
T,ij
, G
ij
,
ij
, and
ij
need to be modeled to close the equations.
The following sections describe the modeling assumptions required to close the equation
set.
11.6.2 Modeling Turbulent Diffusive Transport
D
T,ij
can be modeled by the generalized gradient-diusion model of Daly and Harlow [69]:
D
T,ij
= C
s

x
k
_

ku

k
u

i
u

j
x

_
(11.6-2)
However, this equation can result in numerical instabilities, so it has been simplied in
FLUENT to use a scalar turbulent diusivity as follows [195]:
D
T,ij
=

x
k
_

k
u

i
u

j
x
k
_
(11.6-3)
The turbulent viscosity,
t
, is computed using Equation 11.6-27.
c Fluent Inc. January 11, 2005 11-37
Modeling Turbulence
Lien and Leschziner [195] derived a value of
k
= 0.82 by applying the generalized
gradient-diusion model, Equation 11.6-2, to the case of a planar homogeneous shear
ow. Note that this value of
k
is dierent from that in the standard and realizable k-
models, in which
k
= 1.0.
11.6.3 Modeling the Pressure-Strain Term
Linear Pressure-Strain Model
By default in FLUENT, the pressure-strain term,
ij
, in Equation 11.6-1 is modeled
according to the proposals by Gibson and Launder [110], Fu et al. [105], and Launder [179,
180].
The classical approach to modeling
ij
uses the following decomposition:

ij
=
ij,1
+
ij,2
+
ij,w
(11.6-4)
where
ij,1
is the slow pressure-strain term, also known as the return-to-isotropy term,

ij,2
is called the rapid pressure-strain term, and
ij,w
is the wall-reection term.
The slow pressure-strain term,
ij,1
, is modeled as

ij,1
C
1


k
_
u

i
u

2
3

ij
k
_
(11.6-5)
with C
1
= 1.8.
The rapid pressure-strain term,
ij,2
, is modeled as

ij,2
C
2
_
(P
ij
+ F
ij
+ G
ij
C
ij
)
2
3

ij
(P + GC)
_
(11.6-6)
where C
2
= 0.60, P
ij
, F
ij
, G
ij
, and C
ij
are dened as in Equation 11.6-1, P =
1
2
P
kk
,
G =
1
2
G
kk
, and C =
1
2
C
kk
.
The wall-reection term,
ij,w
, is responsible for the redistribution of normal stresses near
the wall. It tends to damp the normal stress perpendicular to the wall, while enhancing
the stresses parallel to the wall. This term is modeled as

ij,w
C

k
_
u

k
u

m
n
k
n
m

ij

3
2
u

i
u

k
n
j
n
k

3
2
u

j
u

k
n
i
n
k
_
k
3/2
C

d
+ C

2
_

km,2
n
k
n
m

ij

3
2

ik,2
n
j
n
k

3
2

jk,2
n
i
n
k
_
k
3/2
C

d
(11.6-7)
11-38 c Fluent Inc. January 11, 2005
11.6 The Reynolds Stress Model (RSM)
where C

1
= 0.5, C

2
= 0.3, n
k
is the x
k
component of the unit normal to the wall, d is
the normal distance to the wall, and C

= C
3/4

/, where C

= 0.09 and is the von


Karman constant (= 0.4187).

ij,w
is included by default in the Reynolds stress model.
Low-Re Modications to the Linear Pressure-Strain Model
When the RSM is applied to near-wall ows using the enhanced wall treatment described
in Section 11.9.3: Two-Layer Model for Enhanced Wall Treatment, the pressure-strain
model needs to be modied. The modication used in FLUENT species the values of C
1
,
C
2
, C

1
, and C

2
as functions of the Reynolds stress invariants and the turbulent Reynolds
number, according to the suggestion of Launder and Shima [182]:
C
1
= 1 + 2.58A
_
A
2
_
1 exp
_
(0.0067Re
t
)
2
__
(11.6-8)
C
2
= 0.75

A (11.6-9)
C

1
=
2
3
C
1
+ 1.67 (11.6-10)
C

2
= max
_
2
3
C
2

1
6
C
2
, 0
_
(11.6-11)
with the turbulent Reynolds number dened as Re
t
= (k
2
/). The parameter A and
tensor invariants, A
2
and A
3
, are dened as
A
_
1
9
8
(A
2
A
3
)
_
(11.6-12)
A
2
a
ik
a
ki
(11.6-13)
A
3
a
ik
a
kj
a
ji
(11.6-14)
a
ij
is the Reynolds-stress anisotropy tensor, dened as
a
ij
=
_
u

i
u

j
+
2
3
k
ij
k
_
(11.6-15)
The modications detailed above are employed only when the enhanced wall treatment
is selected in the Viscous Model panel.
c Fluent Inc. January 11, 2005 11-39
Modeling Turbulence
Quadratic Pressure-Strain Model
An optional pressure-strain model proposed by Speziale, Sarkar, and Gatski [328] is
provided in FLUENT. This model has been demonstrated to give superior performance in a
range of basic shear ows, including plane strain, rotating plane shear, and axisymmetric
expansion/contraction. This improved accuracy should be benecial for a wider class of
complex engineering ows, particularly those with streamline curvature. The quadratic
pressure-strain model can be selected as an option in the Viscous Model panel.
This model is written as follows:

ij
= (C
1
+ C

1
P) b
ij
+ C
2

_
b
ik
b
kj

1
3
b
mn
b
mn

ij
_
+
_
C
3
C

3
_
b
ij
b
ij
_
kS
ij
+ C
4
k
_
b
ik
S
jk
+ b
jk
S
ik

2
3
b
mn
S
mn

ij
_
+ C
5
k (b
ik

jk
+ b
jk

ik
) (11.6-16)
where b
ij
is the Reynolds-stress anisotropy tensor dened as
b
ij
=
_
u

i
u

j
+
2
3
k
ij
2k
_
(11.6-17)
The mean strain rate, S
ij
, is dened as
S
ij
=
1
2
_
u
j
x
i
+
u
i
x
j
_
(11.6-18)
The mean rate-of-rotation tensor,
ij
, is dened by

ij
=
1
2
_
u
i
x
j

u
j
x
i
_
(11.6-19)
The constants are
C
1
= 3.4, C

1
= 1.8, C
2
= 4.2, C
3
= 0.8, C

3
= 1.3, C
4
= 1.25, C
5
= 0.4
The quadratic pressure-strain model does not require a correction to account for the
wall-reection eect in order to obtain a satisfactory solution in the logarithmic region
of a turbulent boundary layer. It should be noted, however, that the quadratic pressure-
strain model is not available when the enhanced wall treatment is selected in the Viscous
Model panel.
11-40 c Fluent Inc. January 11, 2005
11.6 The Reynolds Stress Model (RSM)
11.6.4 Effects of Buoyancy on Turbulence
The production terms due to buoyancy are modeled as
G
ij
=

t
Pr
t
_
g
i
T
x
j
+ g
j
T
x
i
_
(11.6-20)
where Pr
t
is the turbulent Prandtl number for energy, with a default value of 0.85.
Using the denition of the coecient of thermal expansion, , given by Equation 11.4-24,
the following expression is obtained for G
ij
for ideal gases:
G
ij
=

t
Pr
t
_
g
i

x
j
+ g
j

x
i
_
(11.6-21)
11.6.5 Modeling the Turbulence Kinetic Energy
In general, when the turbulence kinetic energy is needed for modeling a specic term, it
is obtained by taking the trace of the Reynolds stress tensor:
k =
1
2
u

i
u

i
(11.6-22)
As described in Section 11.6.8: Boundary Conditions for the Reynolds Stresses, an option
is available in FLUENT to solve a transport equation for the turbulence kinetic energy in
order to obtain boundary conditions for the Reynolds stresses. In this case, the following
model equation is used:

t
(k) +

x
i
(ku
i
) =

x
j
_
_
+

t

k
_
k
x
j
_
+
1
2
(P
ii
+ G
ii
) (1+2M
2
t
) +S
k
(11.6-23)
where
k
= 0.82 and S
k
is a user-dened source term. Equation 11.6-23 is obtainable
by contracting the modeled equation for the Reynolds stresses (Equation 11.6-1). As
one might expect, it is essentially identical to Equation 11.4-1 used in the standard k-
model.
Although Equation 11.6-23 is solved globally throughout the ow domain, the values of
k obtained are used only for boundary conditions. In every other case, k is obtained from
Equation 11.6-22. This is a minor point, however, since the values of k obtained with
either method should be very similar.
c Fluent Inc. January 11, 2005 11-41
Modeling Turbulence
11.6.6 Modeling the Dissipation Rate
The dissipation tensor,
ij
, is modeled as

ij
=
2
3

ij
( + Y
M
) (11.6-24)
where Y
M
= 2M
2
t
is an additional dilatation dissipation term according to the model
by Sarkar [291]. The turbulent Mach number in this term is dened as
M
t
=

k
a
2
(11.6-25)
where a (

RT) is the speed of sound. This compressibility modication always takes


eect when the compressible form of the ideal gas law is used.
The scalar dissipation rate, , is computed with a model transport equation similar to
that used in the standard k- model:

t
() +

x
i
(u
i
) =

x
j
_
_
+

t

_

x
j
_
C
1
1
2
[P
ii
+ C
3
G
ii
]

k
C
2

2
k
+S

(11.6-26)
where

= 1.0, C
1
= 1.44, C
2
= 1.92, C
3
is evaluated as a function of the local ow
direction relative to the gravitational vector, as described in Section 11.4.5: Eects of
Buoyancy on Turbulence in the k- Models, and S

is a user-dened source term.


11.6.7 Modeling the Turbulent Viscosity
The turbulent viscosity,
t
, is computed similarly to the k- models:

t
= C

k
2

(11.6-27)
where C

= 0.09.
11.6.8 Boundary Conditions for the Reynolds Stresses
Whenever ow enters the domain, FLUENT requires values for individual Reynolds
stresses, u

i
u

j
, and for the turbulence dissipation rate, . These quantities can be input
directly or derived from the turbulence intensity and characteristic length, as described
in Section 11.11.2: Reynolds Stress Model.
At walls, FLUENT computes the near-wall values of the Reynolds stresses and from wall
functions (see Section 11.9.2: Wall Functions). FLUENT applies explicit wall boundary
11-42 c Fluent Inc. January 11, 2005
11.6 The Reynolds Stress Model (RSM)
conditions for the Reynolds stresses by using the log-law and the assumption of equilib-
rium, disregarding convection and diusion in the transport equations for the stresses
(Equation 11.6-1). Using a local coordinate system, where is the tangential coordinate,
is the normal coordinate, and is the binormal coordinate, the Reynolds stresses at
the wall-adjacent cells are computed from
u

k
= 1.098,
u

k
= 0.247,
u

k
= 0.655,
u

k
= 0.255 (11.6-28)
To obtain k, FLUENT solves the transport equation of Equation 11.6-23. For reasons of
computational convenience, the equation is solved globally, even though the values of k
thus computed are needed only near the wall; in the far eld k is obtained directly from the
normal Reynolds stresses using Equation 11.6-22. By default, the values of the Reynolds
stresses near the wall are xed using the values computed from Equation 11.6-28, and
the transport equations in Equation 11.6-1 are solved only in the bulk ow region.
Alternatively, the Reynolds stresses can be explicitly specied in terms of wall-shear
stress, instead of k:
u

u
2

= 5.1,
u

u
2

= 1.0,
u

u
2

= 2.3,
u

u
2

= 1.0 (11.6-29)
where u

is the friction velocity dened by u

w
/, where
w
is the wall-shear stress.
When this option is chosen, the k transport equation is not solved.
11.6.9 Convective Heat and Mass Transfer Modeling
With the Reynolds stress model in 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
(E) +

x
i
[u
i
(E + p)] =

x
j
_
_
k +
c
p

t
Pr
t
_
T
x
j
+ u
i
(
ij
)
e
_
+ S
h
(11.6-30)
where E is the total energy and (
ij
)
e
is the deviatoric stress tensor, dened as
(
ij
)
e
=
e
_
u
j
x
i
+
u
i
x
j
_

2
3

e
u
i
x
i

ij
The term involving (
ij
)
e
represents the viscous heating, and is always computed in the
coupled solvers. It is not computed by default in the segregated solver, but it can be
enabled in the Viscous Model panel. The default value of the turbulent Prandtl number
is 0.85. You can change the value of Pr
t
in the Viscous Model panel.
c Fluent Inc. January 11, 2005 11-43
Modeling Turbulence
Turbulent mass transfer is treated similarly, with a default turbulent Schmidt number of
0.7. This default value can be changed in the Viscous Model panel.
11.7 The Large Eddy Simulation (LES) Model
Turbulent ows are characterized by eddies with a wide range of length and time scales.
The largest eddies are typically comparable in size to the characteristic length of the
mean ow. The smallest scales are responsible for the dissipation of turbulence kinetic
energy.
It is possible, in theory, to directly resolve the whole spectrum of turbulent scales using
an approach known as direct numerical simulation (DNS). No modeling is required in
DNS. However, DNS is not feasible for practical engineering problems involving high
Reynolds number ows. The cost required for DNS to resolve the entire range of scales
is proportional to Re
3
t
, where Re
t
is the turbulent Reynolds number. Clearly, for high
Reynolds numbers, the cost becomes prohibitive.
In LES, large eddies are resolved directly, while small eddies are modeled. Large eddy
simulation (LES) thus falls between DNS and RANS in terms of the fraction of the
resolved scales. The rationale behind LES can be summarized as follows:
Momentum, mass, energy, and other passive scalars are transported mostly by large
eddies.
Large eddies are more problem-dependent. They are dictated by the geometries
and boundary conditions of the ow involved.
Small eddies are less dependent on the geometry, tend to be more isotropic, and
are consequently more universal.
The chance of nding a universal turbulence model is much higher for small eddies.
Resolving only the large eddies allows one to use much coarser mesh and larger times-
step sizes in LES than in DNS. However, LES still requires substantially ner meshes
than those typically used for RANS calculations. In addition, LES has to be run for
a suciently long ow-time to obtain stable statistics of the ow being modeled. As
a result, the computational cost involved with LES is normally orders of magnitudes
higher than that for steady RANS calculations in terms of memory (RAM) and CPU
time. Therefore, high-performance computing (e.g., parallel computing) is a necessity for
LES, especially for industrial applications.
The following sections give details of the governing equations for LES, the subgrid-scale
turbulence models, and the boundary conditions.
11-44 c Fluent Inc. January 11, 2005
11.7 The Large Eddy Simulation (LES) Model
11.7.1 Filtered Navier-Stokes Equations
The governing equations employed for LES are obtained by ltering the time-dependent
Navier-Stokes equations in either Fourier (wave-number) space or conguration (physical)
space. The ltering process eectively lters out the eddies whose scales are smaller than
the lter width or grid spacing used in the computations. The resulting equations thus
govern the dynamics of large eddies.
A ltered variable (denoted by an overbar) is dened by
(x) =
_
D
(x

)G(x, x

)dx

(11.7-1)
where D is the uid domain, and G is the lter function that determines the scale of the
resolved eddies.
In FLUENT, the nite-volume discretization itself implicitly provides the ltering opera-
tion:
(x) =
1
V
_
V
(x

) dx

, x

V (11.7-2)
where V is the volume of a computational cell. The lter function, G(x, x

), implied here
is then
G(x, x

)
_
1/V, x

V
0, x

otherwise
(11.7-3)
The LES capability in FLUENTis applicable to compressible ows. For the sake of concise
notation, however, the theory is presented here for incompressible ows.
Filtering the Navier-Stokes equations, one obtains

t
+

x
i
(u
i
) = 0 (11.7-4)
and

t
(u
i
) +

x
j
(u
i
u
j
) =

x
j
_

ij
x
j
_

p
x
i


ij
x
j
(11.7-5)
where
ij
is the stress tensor due to molecular viscosity dened by

ij

_

_
u
i
x
j
+
u
j
x
i
__

2
3

u
l
x
l

ij
(11.7-6)
c Fluent Inc. January 11, 2005 11-45
Modeling Turbulence
and
ij
is the subgrid-scale stress dened by

ij
u
i
u
j
u
i
u
j
(11.7-7)
11.7.2 Subgrid-Scale Models
The subgrid-scale stresses resulting from the ltering operation are unknown, and re-
quire modeling. The subgrid-scale turbulence models in FLUENTemploy the Boussinesq
hypothesis [131] as in the RANS models, computing subgrid-scale turbulent stresses from

ij

1
3

kk

ij
= 2
t
S
ij
(11.7-8)
where
t
is the subgrid-scale turbulent viscosity, and S
ij
is the rate-of-strain tensor for
the resolved scale dened by
S
ij

1
2
_
u
i
x
j
+
u
j
x
i
_
(11.7-9)
FLUENT oers four models for
t
: the Smagorinsky-Lilly model, the dynamic Smagorinsky-
Lilly model, the WALE model, and the dynamic kinetic energy subgrid-scale model.
The Smagorinsky-Lilly Model
This simple model was rst proposed by Smagorinsky [313]. In the Smagorinsky-Lilly
model, the eddy-viscosity is modeled by

t
= L
2
s

(11.7-10)
where L
s
is the mixing length for subgrid scales and


_
2S
ij
S
ij
. In FLUENT, L
s
is
computed using
L
s
= min
_
d, C
s
V
1/3
_
(11.7-11)
where is the von Karm an constant, d is the distance to the closest wall, C
s
is the
Smagorinsky constant, and V is the volume of the computational cell.
Lilly derived a value of 0.17 for C
s
for homogeneous isotropic turbulence in the inertial
subrange. However, this value was found to cause excessive damping of large-scale uc-
tuations in the presence of mean shear and in transitional ows as near solid boundary,
and has to be reduced in such regions. In short, C
s
is not an universal constant, which
is the most serious shortcoming of this simple model. Nonetheless, C
s
value of around
11-46 c Fluent Inc. January 11, 2005
11.7 The Large Eddy Simulation (LES) Model
0.1 has been found to yield the best results for a wide range of ows, and is the default
value in FLUENT.
The Dynamic Smagorinsky-Lilly Model
Germano et al. [107] and subsequently Lilly [198] conceived a procedure in which the
Smagorinsky model constant, C
s
, is dynamically computed based on the information
provided by the resolved scales of motion. The dynamic procedure thus obviates the
need for users to specify the model constant C
s
in advance. The details of the model
implementation in FLUENTand its validation can be found in [169].
The C
s
obtained using the dynamic Smagorinsky-Lilly model varies in time and space
over a fairly wide range. To avoid numerical instability, in FLUENT, C
s
is clipped at zero
and 0.23 by default.
The Wall-Adapting Local Eddy-Viscosity (WALE) Model
In the WALE model [242], the eddy viscosity is modeled by

t
= L
2
s
(S
d
ij
S
d
ij
)
3/2
(S
ij
S
ij
)
5/2
+ (S
d
ij
S
d
ij
)
5/4
(11.7-12)
where L
s
and S
d
ij
in the WALE model are dened, respectively, as
L
s
= min
_
d, C
w
V
1/3
_
(11.7-13)
S
d
ij
=
1
2
_
g
2
ij
+ g
2
ji
_

1
3

ij
g
2
kk
, g
ij
=
u
i
x
j
(11.7-14)
In FLUENT, the default value of the WALE constant, C
w
, is 0.325 and has been found
to yield satisfactory results for a wide range of ow. The rest of the notation is the
same as for the Smagorinsky-Lilly model. With this spatial operator, the WALE model
is designed to return the correct wall asymptotic (y
3
) behavior for wall bounded ows.
The Dynamic Kinetic Energy Subgrid-Scale Model
The original and dynamic Smagorinsky-Lilly models discussed previously are essentially
algebraic models in which subgrid-scale stresses are parameterized using the resolved ve-
locity scales. The underlying assumption is the local equilibrium between the transferred
energy through the grid-lter scale and the dissipation of kinetic energy at small sub-
grid scales. The subgrid-scale turbulence can be better modeled by accounting for the
transport of the subgrid-scale turbulence kinetic energy.
c Fluent Inc. January 11, 2005 11-47
Modeling Turbulence
The dynamic SGS kinetic energy model in FLUENTreplicates the model proposed by Kim
and Menon [172].
The subgrid-scale kinetic energy is dened as
k
sgs
=
1
2
_
u
2
k
u
2
k
_
(11.7-15)
which is obtained by contracting the subgrid-scale stress in Equation 11.7-7.
The subgrid-scale eddy viscosity,
t
, is computed using k
sgs
as

t
= C
k
k
1/2
sgs

f
(11.7-16)
where
f
is the lter-size computed from
f
V
1/3
The subgrid-scale stress can then be written as

ij

2
3
k
sgs

ij
= 2C
k
k
1/2
sgs

f
S
ij
(11.7-17)
k
sgs
is obtained by solving its transport equation
k
sgs
t
+
u
j
k
sgs
x
j
=
ij
u
i
x
j
C

k
3/2
sgs

f
+

x
j
_

k
k
sgs
x
j
_
(11.7-18)
In the above equations, the model constants, C
k
and C

, are determined dynamically [172].

k
is hardwired to 1.0. The details of the implementation of this model in FLUENTand
its validation is given by Kim [169].
11.7.3 Inlet Boundary Conditions for the LES Model
This section describes the three algorithms available in FLUENT to model the uctuating
velocity at inlet boundaries.
No Perturbations
The stochastic components of the ow at the velocity-specied inlet boundaries are ne-
glected if the No Perturbations option is used. In such cases, individual instantaneous
velocity components are simply set equal to their mean velocity counterparts. This op-
tion is suitable only when the level of turbulence at the inow boundaries is negligible or
does not play a major role in the accuracy of the overall solution.
11-48 c Fluent Inc. January 11, 2005
11.7 The Large Eddy Simulation (LES) Model
Vortex Method
To generate a time-dependent inlet condition, a random 2D vortex method is considered.
With this approach, a perturbation is added on a specied mean velocity prole via a
uctuating vorticity eld (i.e. two-dimensional in the plane normal to the streamwise
direction). The vortex method is based on the Lagrangian form of the 2D evolution
equation of the vorticity and the Biot-Savart law. A particle discretization is used to
solve this equation. These particles, or vortex points are convected randomly and
carry information about the vorticity eld. If N is the number of vortex points and A
is the area of the inlet section, the amount of vorticity carried by a given particle i is
represented by the circulation
i
and an assumed spatial distribution :

i
(x, y) = 4

_
Ak(x, y)
3N[2 ln(3) 3 ln(2)]
(11.7-19)
(x) =
1
2
2
_
2e
|x|
2
/2
2
1
_
2e
|x|
2
/2
2
(11.7-20)
where k is the turbulence kinetic energy. The parameter provides control over the size
of a vortex particle. The resulting discretization for the velocity eld is given by
u(x) =
1
2
N

i=1

i
((x
i
x) z)(1 e
|xx

|
2
/2
2
)
|x x

i
|
2
(11.7-21)
Originally [303], the size of the vortex was xed by an ad hoc value of . To make the
vortex method generally applicable, a local vortex size is specied through a turbulent
mixing length hypothesis. is calculated from a known prole of mean turbulence kinetic
energy and mean dissipation rate at the inlet according to the following:
=
ck
3/2
2
(11.7-22)
where c = 0.16. To ensure that the vortex will always belong to resolved scales, the
minimum value of in Equation 11.7-22 is bounded by the local grid size. The sign
of the circulation of each vortex is changed randomly each characteristic time scale .
In the general implementation of the vortex method, this time scale represents the time
necessary for a 2D vortex convected by the bulk velocity in the boundary normal direction
to travel along n times its mean characteristic 2D size (
m
), where n is xed equal to
100 from numerical testing. The vortex method considers only uctuations in the plane
normal to the streamwise direction. By their interaction with the mean velocity eld,
these uctuations impact streamwise uctuations downstream from the inlet plane as
well.
c Fluent Inc. January 11, 2005 11-49
Modeling Turbulence
i
Since the vortex method theory is based on the modication of the velocity
eld normal to the streamwise direction, it is imperative that user creates
inlet plane normal (or as close as possible) to the streamwise velocity di-
rection.
Spectral Synthesizer
The spectral synthesizer provides an alternative method of generating uctuating velocity
components. It is based on the random ow generation technique originally proposed by
Kraichnan [174] and modied by Smirnov et al. [314]. In this method, uctuating velocity
components are computed by synthesizing a divergence-free velocity-vector eld from
the summation of Fourier harmonics. In the implementation in FLUENT, the number of
Fourier harmonics is xed to 100.
11.8 The Detached Eddy Simulation (DES) Model
The DES model belongs to the class of models usually referred to as an LES/RANS cou-
pling modeling approach. The main idea of this approach is to combine RANS modeling
with LES for applications in which classical LES is not aordable (e.g., high-Re external
aerodynamics simulations). In FLUENT, the DES model is based on the one-equation
Spalart-Allmaras model.
The standard Spalart-Allmaras model uses the distance to the closest wall as the deni-
tion for the length scale d, which plays a major role in determining the level of production
and destruction of turbulent viscosity (Equations 11.3-6, 11.3-12, and 11.3-15). The DES
model, as proposed by Shur et al. [307] replaces d everywhere with a new length scale

d,
dened as

d = min(d, C
des
) (11.8-1)
where the grid spacing, , is based on the largest grid space in the x, y, or z directions
forming the computational cell. The empirical constant C
des
has a value of 0.65.
11-50 c Fluent Inc. January 11, 2005
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
11.9.1 Overview
Turbulent ows are signicantly aected by the presence of walls. Obviously, the mean
velocity eld is aected through the no-slip condition that has to be satised at the wall.
However, the turbulence is also changed by the presence of the wall in non-trivial ways.
Very close to the wall, viscous damping reduces the tangential velocity uctuations, while
kinematic blocking reduces the normal uctuations. Toward the outer part of the near-
wall region, however, the turbulence is rapidly augmented by the production of turbulence
kinetic energy due to the large gradients in mean velocity.
The near-wall modeling signicantly impacts the delity of numerical solutions, inasmuch
as walls are the main source of mean vorticity and turbulence. After all, it is in the near-
wall region that the solution variables have large gradients, and the momentum and other
scalar transports occur most vigorously. Therefore, accurate representation of the ow in
the near-wall region determines successful predictions of wall-bounded turbulent ows.
The k- models, the RSM, and the LES model are primarily valid for turbulent core
ows (i.e., the ow in the regions somewhat far from walls). Consideration therefore
needs to be given as to how to make these models suitable for wall-bounded ows. The
Spalart-Allmaras and k- models were designed to be applied throughout the boundary
layer, provided that the near-wall mesh resolution is sucient.
Numerous experiments have shown that the near-wall region can be largely subdivided
into three layers. In the innermost layer, called the viscous sublayer, the ow is almost
laminar, and the (molecular) viscosity plays a dominant role in momentum and heat
or mass transfer. In the outer layer, called the fully-turbulent layer, turbulence plays
a major role. Finally, there is an interim region between the viscous sublayer and the
fully turbulent layer where the eects of molecular viscosity and turbulence are equally
important. Figure 11.9.1 illustrates these subdivisions of the near-wall region, plotted in
semi-log coordinates.
Wall Functions vs. Near-Wall Model
Traditionally, there are two approaches to modeling the near-wall region. In one ap-
proach, the viscosity-aected inner region (viscous sublayer and buer layer) is not re-
solved. Instead, semi-empirical formulas called wall functions are used to bridge the
viscosity-aected region between the wall and the fully-turbulent region. The use of wall
functions obviates the need to modify the turbulence models to account for the presence
of the wall.
In another approach, the turbulence models are modied to enable the viscosity-aected
region to be resolved with a mesh all the way to the wall, including the viscous sublayer.
For purposes of discussion, this will be termed the near-wall modeling approach. These
two approaches are depicted schematically in Figure 11.9.2.
c Fluent Inc. January 11, 2005 11-51
Modeling Turbulence
Figure 11.9.1: Subdivisions of the Near-Wall Region
Figure 11.9.2: Near-Wall Treatments in FLUENT
11-52 c Fluent Inc. January 11, 2005
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
In most high-Reynolds-number ows, the wall function approach substantially saves com-
putational resources, because the viscosity-aected near-wall region, in which the solution
variables change most rapidly, does not need to be resolved. The wall function approach
is popular because it is economical, robust, and reasonably accurate. It is a practical
option for the near-wall treatments for industrial ow simulations.
The wall function approach, however, is inadequate in situations where the low-Reynolds-
number eects are pervasive in the ow domain in question, and the hypotheses under-
lying the wall functions cease to be valid. Such situations require near-wall models that
are valid in the viscosity-aected region and accordingly integrable all the way to the
wall.
FLUENT provides both the wall function approach and the near-wall modeling approach.
Near-Wall Treatments for the Spalart-Allmaras, k-, and LES Models
See Sections 11.3.6, 11.5.1, and 11.7.3, respectively, for a description of the near-wall
treatments applied by the Spalart-Allmaras, k-, and LES models.
11.9.2 Wall Functions
Wall functions are a collection of semi-empirical formulas and functions that in eect
bridge or link the solution variables at the near-wall cells and the corresponding
quantities on the wall. The wall functions comprise
laws-of-the-wall for mean velocity and temperature (or other scalars)
formulas for near-wall turbulent quantities
FLUENT oers two choices of wall function approaches:
standard wall functions
non-equilibrium wall functions
Standard Wall Functions
The standard wall functions in FLUENT are based on the proposal of Launder and Spald-
ing [184], and have been most widely used for industrial ows. They are provided as a
default option in FLUENT.
c Fluent Inc. January 11, 2005 11-53
Modeling Turbulence
Momentum
The law-of-the-wall for mean velocity yields
U

=
1

ln(Ey

) (11.9-1)
where
U


U
P
C
1/4

k
1/2
P

w
/
(11.9-2)
y


C
1/4

k
1/2
P
y
P

(11.9-3)
and = von Karman constant (= 0.4187)
E = empirical constant (= 9.793)
U
P
= mean velocity of the uid at point P
k
P
= turbulence kinetic energy at point P
y
P
= distance from point P to the wall
= dynamic viscosity of the uid
The logarithmic law for mean velocity is known to be valid for 30 < y

< 300. In
FLUENT, the log-law is employed when y

> 11.225.
When the mesh is such that y

< 11.225 at the wall-adjacent cells, FLUENT applies the


laminar stress-strain relationship that can be written as
U

= y

(11.9-4)
It should be noted that, in FLUENT, the laws-of-the-wall for mean velocity and temper-
ature are based on the wall unit, y

, rather than y
+
( u

y/). These quantities are


approximately equal in equilibrium turbulent boundary layers.
Energy
Reynolds analogy between momentum and energy transport gives a similar logarithmic
law for mean temperature. As in the law-of-the-wall for mean velocity, the law-of-the-wall
for temperature employed in FLUENT comprises the following two dierent laws:
linear law for the thermal conduction sublayer where conduction is important
logarithmic law for the turbulent region where eects of turbulence dominate con-
duction
11-54 c Fluent Inc. January 11, 2005
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
The thickness of the thermal conduction layer is, in general, dierent from the thickness
of the (momentum) viscous sublayer, and changes from uid to uid. For example, the
thickness of the thermal sublayer for a high-Prandtl-number uid (e.g., oil) is much less
than its momentum sublayer thickness. For uids of low Prandtl numbers (e.g., liquid
metal), on the contrary, it is much larger than the momentum sublayer thickness.
In highly compressible ows, the temperature distribution in the near-wall region can
be signicantly dierent from that of low subsonic ows, due to the heating by viscous
dissipation. In FLUENT, the temperature wall functions include the contribution from
the viscous heating [356].
The law-of-the-wall implemented in FLUENT has the following composite form:
T


(T
w
T
P
) c
p
C
1/4

k
1/2
P
q
=
_

_
Pr y

+
1
2
Pr
C
1/4

k
1/2
P
q
U
2
P
(y

< y

T
)
Pr
t
_
1

ln(Ey

) + P
_
+
1
2

C
1/4

k
1/2
P
q
{Pr
t
U
2
P
+ (Pr Pr
t
)U
2
c
} (y

> y

T
)
(11.9-5)
where P is computed by using the formula given by Jayatilleke [149]:
P = 9.24
_
_
Pr
Pr
t
_
3/4
1
_
_
1 + 0.28e
0.007Pr/Prt
_
(11.9-6)
and
k
P
= turbulent kinetic energy at point P
= density of uid
c
p
= specic heat of uid
q = wall heat ux
T
P
= temperature at the cell adjacent to wall
T
w
= temperature at the wall
Pr = molecular Prandtl number (c
p
/k
f
)
Pr
t
= turbulent Prandtl number (0.85 at the wall)
A = Van Driest constant (= 26)
U
c
= mean velocity magnitude at y

= y

T
Note that, for the segregated solver, the terms
1
2
Pr
C
1/4

k
1/2
P
q
U
2
P
and
1
2

C
1/4

k
1/2
P
q
_
Pr
t
U
2
P
+ (Pr Pr
t
)U
2
c
_
c Fluent Inc. January 11, 2005 11-55
Modeling Turbulence
will be included in Equation 11.9-5 only for compressible ow calculations.
The non-dimensional thermal sublayer thickness, y

T
, in Equation 11.9-5 is computed as
the y

value at which the linear law and the logarithmic law intersect, given the molecular
Prandtl number of the uid being modeled.
The procedure of applying the law-of-the-wall for temperature is as follows. Once the
physical properties of the uid being modeled are specied, its molecular Prandtl number
is computed. Then, given the molecular Prandtl number, the thermal sublayer thickness,
y

T
, is computed from the intersection of the linear and logarithmic proles, and stored.
During the iteration, depending on the y

value at the near-wall cell, either the linear or


the logarithmic prole in Equation 11.9-5 is applied to compute the wall temperature T
w
or heat ux q (depending on the type of the thermal boundary conditions).
The function for P given by equation Equation 11.9-6 is relevant for the smooth walls.
For the rough walls, however, this function is modied as follows:
P
rough
= 3.15Pr
0.695
_
1
E


1
E
_
0.359
+
_
E

E
_
0.6
P (11.9-7)
where E

is wall function constant modied for the rough walls (see Section 7.13.1: Mod-
eling Wall Roughness Eects in Turbulent Wall-Bounded Flows for the details).
Species
When using wall functions for species transport, FLUENT assumes that species transport
behaves analogously to heat transfer. Similarly to Equation 11.9-5, the law-of-the-wall
for species can be expressed for constant property ow with no viscous dissipation as
Y


(Y
i,w
Y
i
) C
1/4

k
1/2
P
J
i,w
=
_
Sc y

(y

< y

c
)
Sc
t
_
1

ln(Ey

) + P
c
_
(y

> y

c
)
(11.9-8)
where Y
i
is the local species mass fraction, Sc and Sc
t
are molecular and turbulent
Schmidt numbers, and J
i,w
is the diusion ux of species i at the wall. Note that P
c
and
y

c
are calculated in a similar way as P and y

T
, with the dierence being that the Prandtl
numbers are always replaced by the corresponding Schmidt numbers.
Turbulence
In the k- models and in the RSM (if the option to obtain wall boundary conditions from
the k equation is enabled), the k equation is solved in the whole domain including the
wall-adjacent cells. The boundary condition for k imposed at the wall is
k
n
= 0 (11.9-9)
11-56 c Fluent Inc. January 11, 2005
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
where n is the local coordinate normal to the wall.
The production of kinetic energy, G
k
, and its dissipation rate, , at the wall-adjacent
cells, which are the source terms in the k equation, are computed on the basis of the local
equilibrium hypothesis. Under this assumption, the production of k and its dissipation
rate are assumed to be equal in the wall-adjacent control volume.
Thus, the production of k is computed from
G
k

w
U
y
=
w

w
C
1/4
k
1/2
P
y
P
(11.9-10)
and is computed from

P
=
C
3/4

k
3/2
P
y
P
(11.9-11)
The equation is not solved at the wall-adjacent cells, but instead is computed using
Equation 11.9-11.
Note that, as shown here, the wall boundary conditions for the solution variables, in-
cluding mean velocity, temperature, species concentration, k, and , are all taken care of
by the wall functions. Therefore, you do not need to be concerned about the boundary
conditions at the walls.
The standard wall functions described so far are provided as a default option in FLUENT.
The standard wall functions work reasonably well for a broad range of wall-bounded ows.
However, they tend to become less reliable when the ow situations depart too much from
the ideal conditions that are assumed in their derivation. Among others, the constant-
shear and local equilibrium hypotheses are the ones that most restrict the universality
of the standard wall functions. Accordingly, when the near-wall ows are subjected to
severe pressure gradients, and when the ows are in strong non-equilibrium, the quality
of the predictions is likely to be compromised.
The non-equilibrium wall functions oered as an additional option can improve the results
in such situations.
c Fluent Inc. January 11, 2005 11-57
Modeling Turbulence
Non-Equilibrium Wall Functions
In addition to the standard wall function described above (which is the default near-wall
treatment) a two-layer-based, non-equilibrium wall function [170] is also available. The
key elements in the non-equilibrium wall functions are as follows:
Launder and Spaldings log-law for mean velocity is sensitized to pressure-gradient
eects.
The two-layer-based concept is adopted to compute the budget of turbulence kinetic
energy (G
k
,) in the wall-neighboring cells.
The law-of-the-wall for mean temperature or species mass fraction remains the same as
in the standard wall function described above.
The log-law for mean velocity sensitized to pressure gradients is

UC
1/4

k
1/2

w
/
=
1

ln
_
E
C
1/4

k
1/2
y

_
(11.9-12)
where

U = U
1
2
dp
dx
_
y
v

k
ln
_
y
y
v
_
+
y y
v

k
+
y
2
v

_
(11.9-13)
and y
v
is the physical viscous sublayer thickness, and is computed from
y
v

y

v
C
1/4
k
1/2
P
(11.9-14)
where y

v
= 11.225.
The non-equilibrium wall function employs the two-layer concept in computing the bud-
get of turbulence kinetic energy at the wall-adjacent cells, which is needed to solve the k
equation at the wall-neighboring cells. The wall-neighboring cells are assumed to consist
of a viscous sublayer and a fully turbulent layer. The following prole assumptions for
turbulence quantities are made:

t
=
_
0, y < y
v

w
, y > y
v
k =
_
_
_
_
y
yv
_
2
k
P
, y < y
v
k
P
, y > y
v
=
_
_
_
2k
y
2
, y < y
v
k
3/2
C

y
, y > y
v
(11.9-15)
where C

= C
3/4

, and y
v
is the dimensional thickness of the viscous sublayer, dened
in Equation 11.9-14.
11-58 c Fluent Inc. January 11, 2005
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
Using these proles, the cell-averaged production of k, G
k
, and the cell-averaged dissipa-
tion rate, , can be computed from the volume average of G
k
and of the wall-adjacent
cells. For quadrilateral and hexahedral cells for which the volume average can be ap-
proximated with a depth-average,
G
k

1
y
n
_
yn
0

t
U
y
dy =
1
y
n

2
w
C
1/4
k
1/2
P
ln
_
y
n
y
v
_
(11.9-16)
and

1
y
n
_
yn
0
dy =
1
y
n
_
_
2
y
v
+
k
1/2
P
C

ln
_
y
n
y
v
_
_
_
k
P
(11.9-17)
where y
n
is the height of the cell (y
n
= 2y
P
). For cells with other shapes (e.g., triangular
and tetrahedral grids), the appropriate volume averages are used.
In Equations 11.9-16 and 11.9-17, the turbulence kinetic energy budget for the wall-
neighboring cells is eectively sensitized to the proportions of the viscous sublayer and
the fully turbulent layer, which varies widely from cell to cell in highly non-equilibrium
ows. It eectively relaxes the local equilibrium assumption (production = dissipation)
that is adopted by the standard wall function in computing the budget of the turbulence
kinetic energy at wall-neighboring cells. Thus, the non-equilibrium wall functions, in
eect, partly account for non-equilibrium eects neglected in the standard wall function.
Standard Wall Functions vs. Non-Equilibrium Wall Functions
Because of the capability to partly account for the eects of pressure gradients and
departure from equilibrium, the non-equilibrium wall functions are recommended for use
in complex ows involving separation, reattachment, and impingement where the mean
ow and turbulence are subjected to severe pressure gradients and change rapidly. In
such ows, improvements can be obtained, particularly in the prediction of wall shear
(skin-friction coecient) and heat transfer (Nusselt or Stanton number).
c Fluent Inc. January 11, 2005 11-59
Modeling Turbulence
Limitations of the Wall Function Approach
The standard wall functions give reasonably accurate predictions for the majority of
high-Reynolds-number, wall-bounded ows. The non-equilibrium wall functions further
extend the applicability of the wall function approach by including the eects of pressure
gradient and strong non-equilibrium. However, the wall function approach becomes less
reliable when the ow conditions depart too much from the ideal conditions underlying
the wall functions. Examples are as follows:
Pervasive low-Reynolds-number or near-wall eects (e.g., ow through a small gap
or highly viscous, low-velocity uid ow).
Massive transpiration through the wall (blowing/suction).
Severe pressure gradients leading to boundary layer separations.
Strong body forces (e.g., ow near rotating disks, buoyancy-driven ows).
High three-dimensionality in the near-wall region (e.g., Ekman spiral ow, strongly
skewed 3D boundary layers).
If any of the items listed above is a prevailing feature of the ow you are modeling, and
if it is considered critically important to capture that feature for the success of your
simulation, you must employ the near-wall modeling approach combined with adequate
mesh resolution in the near-wall region. FLUENT provides the enhanced wall treatment
for such situations. This approach can be used with the three k- models and the RSM.
11.9.3 Enhanced Wall Treatment
Enhanced wall treatment is a near-wall modeling method that combines a two-layer
model with enhanced wall functions. If the near-wall mesh is ne enough to be able to
resolve the laminar sublayer (typically y
+
1), then the enhanced wall treatment will
be identical to the traditional two-layer zonal model (see below for details). However,
the restriction that the near-wall mesh must be suciently ne everywhere might impose
too large a computational requirement. Ideally, then, one would like to have a near-wall
formulation that can be used with coarse meshes (usually referred to as wall-function
meshes) as well as ne meshes (low-Reynolds-number meshes). In addition, excessive
error should not be incurred for intermediate meshes that are too ne for the near-wall
cell centroid to lie in the fully turbulent region, but also too coarse to properly resolve
the sublayer.
To achieve the goal of having a near-wall modeling approach that will possess the accuracy
of the standard two-layer approach for ne near-wall meshes and that, at the same time,
will not signicantly reduce accuracy for wall-function meshes, FLUENT can combine the
two-layer model with enhanced wall functions, as described in the following sections.
11-60 c Fluent Inc. January 11, 2005
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
Two-Layer Model for Enhanced Wall Treatment
In FLUENTs near-wall model, the viscosity-aected near-wall region is completely re-
solved all the way to the viscous sublayer. The two-layer approach is an integral part
of the enhanced wall treatment and is used to specify both and the turbulent vis-
cosity in the near-wall cells. In this approach, the whole domain is subdivided into a
viscosity-aected region and a fully-turbulent region. The demarcation of the two re-
gions is determined by a wall-distance-based, turbulent Reynolds number, Re
y
, dened
as
Re
y

y

(11.9-18)
where y is the normal distance from the wall at the cell centers. In FLUENT, y is
interpreted as the distance to the nearest wall:
y min
rww
r r
w
(11.9-19)
where r is the position vector at the eld point, and r
w
is the position vector on the
wall boundary.
w
is the union of all the wall boundaries involved. This interpretation
allows y to be uniquely dened in ow domains of complex shape involving multiple
walls. Furthermore, y dened in this way is independent of the mesh topology used, and
is denable even on unstructured meshes.
In the fully turbulent region (Re
y
> Re

y
; Re

y
= 200), the k- models or the RSM
(described in Sections 11.4 and 11.6) are employed.
In the viscosity-aected near-wall region (Re
y
< Re

y
), the one-equation model of Wolf-
stein [381] is employed. In the one-equation model, the momentum equations and the
k equation are retained as described in Sections 11.4 and 11.6. However, the turbulent
viscosity,
t
, is computed from

t,2layer
= C

k (11.9-20)
where the length scale that appears in Equation 11.9-20 is computed from [51]

= yc

_
1 e
Rey/A
_
(11.9-21)
The two-layer formulation for turbulent viscosity described above is used as a part of the
enhanced wall treatment, in which the two-layer denition is smoothly blended with the
high-Reynolds-number
t
denition from the outer region, as proposed by Jongen [154]:

t,enh
=

t
+ (1

)
t,2layer
(11.9-22)
c Fluent Inc. January 11, 2005 11-61
Modeling Turbulence
where
t
is the high-Reynolds-number denition as described in Section 11.4: The Stan-
dard, RNG, and Realizable k- Models or 11.6 for the k- models or the RSM. A blending
function,

, is dened in such a way that it is equal to unity far from walls and is zero
very near to walls. The blending function chosen is

=
1
2
_
1 + tanh
_
Re
y
Re

y
A
__
(11.9-23)
The constant A determines the width of the blending function. By dening a width such
that the value of

will be within 1% of its far-eld value given a variation of Re


y
, the
result is
A =
|Re
y
|
tanh(0.98)
(11.9-24)
Typically, Re
y
would be assigned a value that is between 5% and 20% of Re

y
. The
main purpose of the blending function

is to prevent solution convergence from being


impeded when the k- solution in the outer layer does not match with the two-layer
formulation.
The eld is computed from
=
k
3/2

(11.9-25)
The length scales that appear in Equation 11.9-25 are again computed from Chen and
Patel [51]:

= yc

_
1 e
Rey/A
_
(11.9-26)
If the whole ow domain is inside the viscosity-aected region (Re
y
< 200), is not
obtained by solving the transport equation; it is instead obtained algebraically from
Equation 11.9-25. FLUENT uses a procedure for the specication that is similar to the

t
blending in order to ensure a smooth transition between the algebraically-specied
in the inner region and the obtained from solution of the transport equation in the
outer region.
The constants in the length scale formulas, Equations 11.9-21 and 11.9-26, are taken
from [51]:
c

= C
3/4

, A

= 70, A

= 2c

(11.9-27)
11-62 c Fluent Inc. January 11, 2005
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
Enhanced Wall Functions
To have a method that can extend its applicability throughout the near-wall region
(i.e., laminar sublayer, buer region, and fully-turbulent outer region) it is necessary to
formulate the law-of-the wall as a single wall law for the entire wall region. FLUENT
achieves this by blending linear (laminar) and logarithmic (turbulent) laws-of-the-wall
using a function suggested by Kader [158]:
u
+
= e

u
+
lam
+ e
1

u
+
turb
(11.9-28)
where the blending function is given by:
=
a(y
+
)
4
1 + by
+
(11.9-29)
where a = 0.01 and b = 5.
Similarly, the general equation for the derivative
du
+
dy
+
is
du
+
dy
+
= e

du
+
lam
dy
+
+ e
1

du
+
turb
dy
+
(11.9-30)
This approach allows the fully turbulent law to be easily modied and extended to take
into account other eects such as pressure gradients or variable properties. This formula
also guarantees the correct asymptotic behavior for large and small values of y
+
and
reasonable representation of velocity proles in the cases where y
+
falls inside the wall
buer region (3 < y
+
< 10).
The enhanced wall functions were developed by smoothly blending an enhanced turbulent
wall law with the laminar wall law. The enhanced turbulent law-of-the-wall for compress-
ible ow with heat transfer and pressure gradients has been derived by combining the
approaches of White and Cristoph [377] and Huang et al. [137]:
du
+
turb
dy
+
=
1
y
+
_
S

(1 u
+
(u
+
)
2
)
_
1/2
(11.9-31)
where
S

=
_
1 + y
+
for y
+
< y
+
s
1 + y
+
s
for y
+
y
+
s
(11.9-32)
c Fluent Inc. January 11, 2005 11-63
Modeling Turbulence
and


w

w
u

dp
dx
=

2
(u

)
3
dp
dx
(11.9-33)


t
q
w
u

c
p

w
T
w
=

t
q
w
c
p
u

T
w
(11.9-34)


t
(u

)
2
2c
p
T
w
(11.9-35)
where y
+
s
is the location at which the log-law slope will remain xed. By default, y
+
s
= 60.
The coecient in Equation 11.9-31 represents the inuences of pressure gradients
while the coecients and represent thermal eects. Equation 11.9-31 is an ordinary
dierential equation and FLUENT will provide an appropriate analytical solution. If , ,
and all equal 0, an analytical solution would lead to the classical turbulent logarithmic
law-of-the-wall.
The laminar law-of-the-wall is determined from the following expression:
du
+
lam
dy
+
= 1 + y
+
(11.9-36)
Note that the above expression only includes eects of pressure gradients through ,
while the eects of variable properties due to heat transfer and compressibility on the
laminar wall law are neglected. These eects are neglected because they are thought to be
of minor importance when they occur close to the wall. Integration of Equation 11.9-36
results in
u
+
lam
= y
+
_
1 +

2
y
+
_
(11.9-37)
Enhanced thermal wall functions follow the same approach developed for the prole of
u
+
. The unied wall thermal formulation blends the laminar and logarithmic proles
according to the method of Kader [158]:
T
+
= e

T
+
lam
+ e
1

T
+
turb
(11.9-38)
where
=
a(Pr y
+
)
4
1 + bPr
3
y
+
(11.9-39)
where Pr is the molecular Prandtl number, and the coecients a and b are dened as
in Equation 11.9-29. Apart from the above formulation for T
+
, enhanced thermal wall
11-64 c Fluent Inc. January 11, 2005
11.9 Near-Wall Treatments for Wall-Bounded Turbulent Flows
functions follow the same logic as previously described for standard thermal wall functions
(see Section 11.9.2: Energy). A similar procedure is also used for species wall functions
when the enhanced wall treatment is used. See Section 11.9.2: Species for details about
species wall functions.
The boundary condition for turbulence kinetic energy is the same as for standard wall
functions (Equation 11.9-9). However, the production of turbulence kinetic energy G
k
is
computed using the velocity gradients that are consistent with the enhanced law-of-the-
wall (Equations 11.9-28 and 11.9-30), ensuring a formulation that is valid throughout the
near-wall region.
11.9.4 LES Near-Wall Treatment
When the mesh is ne enough to resolve the laminar sublayer, the wall shear stress is
obtained from the laminar stress-strain relationship:
u
u

=
u

(11.9-40)
If the mesh is too coarse to resolve the laminar sublayer, 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
u

=
1

ln E
_
u

_
(11.9-41)
where is the von Karm an constant and E = 9.793. If the mesh is a such that the rst
near wall point is within the buer region, then two above laws are blended in accordance
with equation Equation 11.9-28.
For the LES simulations in FLUENT, there is an alternative near wall approach based on
the work of Werner and Wengle [373], who proposed analytical integration of power-law
near-wall velocity distribution resulting in the following expressions for the wall shear
stress:
|
w
| =
_

_
2|up|
z
for |u
p
|

2z
A
2
1B

_
1B
2
A
1+B
1B
_

z
_
1+B
+
1+B
A
_

z
_
B
|u
p
|
_ 2
1+B
for |u
p
| >

2z
A
2
1B
(11.9-42)
where u
p
is velocity parallel to the wall, A = 8.3, B = 1/7 are the constants, and z is
the near-wall control volume length scale.
The Werner-Wengle wall functions can be enabled using the define/models/viscous/
near-wall-treatment/werner-wengle-wall-fn? text command.
c Fluent Inc. January 11, 2005 11-65
Modeling Turbulence
11.10 Grid Considerations for Turbulent Flow Simulations
Successful computations of turbulent ows require some consideration during the mesh
generation. Since turbulence (through the spatially-varying eective viscosity) plays a
dominant role in the transport of mean momentum and other parameters, you must
ascertain that turbulence quantities in complex turbulent ows are properly resolved if
high accuracy is required. Due to the strong interaction of the mean ow and turbulence,
the numerical results for turbulent ows tend to be more susceptible to grid dependency
than those for laminar ows.
It is therefore recommended that you resolve, with suciently ne meshes, the regions
where the mean ow changes rapidly and there are shear layers with a large mean rate
of strain.
You can check the near-wall mesh by displaying or plotting the values of y
+
, y

, and
Re
y
, which are all available in the postprocessing panels. It should be remembered that
y
+
, y

, and Re
y
are not xed, geometrical quantities. They are all solution-dependent.
For example, when you double the mesh (thereby halving the wall distance), the new y
+
does not necessarily become half of the y
+
for the original mesh.
For the mesh in the near-wall region, dierent strategies must be used depending on which
near-wall option you are using. In Sections 11.10.1 and 11.10.2 are general guidelines for
the near-wall mesh.
11.10.1 Near-Wall Mesh Guidelines for Wall Functions
The log-law, which is valid for equilibrium boundary layers and fully developed ows,
provides upper and lower bounds on the acceptable distance between the cell centroid
and the wall for wall-adjacent cells. The distance is usually measured in the wall unit,
y
+
( u

y/), or y

. Note that y
+
and y

have comparable values when the rst cell is


placed in the log-layer.
For standard or non-equilibrium wall functions, each wall-adjacent cells centroid
should be located within the log-law layer, 30 < y
+
< 300. A y
+
value close to the
lower bound (y
+
30) is most desirable.
Although FLUENT employs the linear (laminar) law when y

< 11.225, using an


excessively ne mesh near the walls should be avoided, because the wall functions
cease to be valid in the viscous sublayer.
As much as possible, the mesh should be made either coarse or ne enough to
prevent the wall-adjacent cells from being placed in the buer layer (y
+
= 5 30).
The upper bound of the log-layer depends on, among others, pressure gradients
and Reynolds number. As the Reynolds number increases, the upper bound tends
to also increase. y
+
values that are too large are not desirable, because the wake
component becomes substantially large above the log-layer.
11-66 c Fluent Inc. January 11, 2005
11.10 Grid Considerations for Turbulent Flow Simulations
Using excessive stretching in the direction normal to the wall should be avoided.
It is important to have at least a few cells inside the boundary layer.
11.10.2 Near-Wall Mesh Guidelines for the Enhanced Wall Treatment
Although the enhanced wall treatment is designed to extend the validity of near-wall
modeling beyond the viscous sublayer, it is still recommended that you construct a mesh
that will fully resolve the viscosity-aected near-wall region. In such a case, the two-layer
component of the enhanced wall treatment will be dominant and the following mesh
requirements are recommended (note that, here, the mesh requirements are in terms of
y
+
, not y

):
When the enhanced wall treatment is employed with the intention of resolving the
laminar sublayer, y
+
at the wall-adjacent cell should be on the order of y
+
= 1.
However, a higher y
+
is acceptable as long as it is well inside the viscous sublayer
(y
+
< 4 to 5).
You should have at least 10 cells within the viscosity-aected near-wall region
(Re
y
< 200) to be able to resolve the mean velocity and turbulent quantities in
that region.
11.10.3 Near-Wall Mesh Guidelines for the Spalart-Allmaras Model
The Spalart-Allmaras model in its complete implementation is a low-Reynolds-number
model. This means that it is designed to be used with meshes that properly resolve
the viscous-aected region, and damping functions have been built into the model in
order to properly attenuate the turbulent viscosity in the viscous sublayer. Therefore, to
obtain the full benet of the Spalart-Allmaras model, the near-wall mesh spacing should
be as described in Section 11.10.2: Near-Wall Mesh Guidelines for the Enhanced Wall
Treatment for the enhanced wall treatment.
However, as discussed in Section 11.3.6: Wall Boundary Conditions, the boundary condi-
tions for the Spalart-Allmaras model have been implemented so that the model will work
on coarser meshes, such as would be appropriate for the wall function approach. If you are
using a coarse mesh, you should follow the guidelines described in Section 11.10.1: Near-
Wall Mesh Guidelines for Wall Functions.
In summary, for best results with the Spalart-Allmaras model, you should use either a
very ne near-wall mesh spacing (on the order of y
+
= 1) or a mesh spacing such that
y
+
30.
c Fluent Inc. January 11, 2005 11-67
Modeling Turbulence
11.10.4 Near-Wall Mesh Guidelines for the k- Models
Both k- models available in FLUENT are available as low-Reynolds-number models as
well as high-Reynolds-number models. If the Transitional Flows option is enabled in the
Viscous Model panel, low-Reynolds-number variants will be used, and, in that case, mesh
guidelines should be the same as for the enhanced wall treatment. However, if this option
is not active, then the mesh guidelines should be the same as for the wall functions.
11.10.5 Near-Wall Mesh Guidelines for Large Eddy Simulation
For the LES implementation in FLUENT, the wall boundary conditions have been imple-
mented using a law-of-the-wall approach as described in Section 11.7.3: Inlet Boundary
Conditions for the LES Model. This means that there are no computational restrictions
on the near-wall mesh spacing. However, for best results, it might be necessary to use a
very ne near-wall mesh spacing (on the order of y
+
= 1).
11.11 Problem Setup for Turbulent Flows
When your FLUENT model includes turbulence you need to activate the relevant model
and options, and supply turbulent boundary conditions. These inputs are described in
this section.
The procedure for setting up a turbulent ow problem is described below. (Note that
this procedure includes only those steps necessary for the turbulence model itself; you
will need to set up other models, boundary conditions, etc. as usual.)
1. To activate the turbulence model, select Spalart-Allmaras, k-epsilon, k-omega, Reynolds
Stress, or (in 3D) Large Eddy Simulation under Model in the Viscous Model panel
(Figure 11.11.1).
Dene Models Viscous...
If you choose the k-epsilon model, select Standard, RNG, or Realizable under k-epsilon
Model. If you choose the k-omega model, select Standard or SST under k-omega
Model.
i
The Large Eddy Simulation model is available only for 3D cases.
2. If the ow involves walls, and you are using one of the k- models or the RSM, choose
one of the following options for the Near-Wall Treatment in the Viscous Model panel:
Standard Wall Functions
Non-Equilibrium Wall Functions
Enhanced Wall Treatment
11-68 c Fluent Inc. January 11, 2005
11.11 Problem Setup for Turbulent Flows
Figure 11.11.1: The Viscous Model Panel
These near-wall options are described in detail in Section 11.9: Near-Wall Treat-
ments for Wall-Bounded Turbulent Flows. By default, the standard wall function
is enabled.
The near-wall treatment for the Spalart-Allmaras, k-, and LES models is dened
automatically, as described in Sections 11.3.6, 11.5.1, and 11.7.3, respectively.
3. Enable the appropriate turbulence modeling options in the Viscous Model panel.
See Section 11.11.1: Turbulence Options for details.
4. Specify the boundary conditions for the solution variables.
Dene Boundary Conditions...
See Section 11.11.2: Dening Turbulence Boundary Conditions for details.
c Fluent Inc. January 11, 2005 11-69
Modeling Turbulence
5. Specify the initial guess for the solution variables.
Solve Initialize Initialize...
See Section 11.11.3: Providing an Initial Guess for k and (or k and ) for details.
Note that Reynolds stresses are automatically initialized using k, and therefore
need not be initialized.
11.11.1 Turbulence Options
The various options available for the turbulence models are described in detail in Sec-
tions 11.3 through 11.7. Instructions for activating these options are provided here.
If you choose the Spalart-Allmaras model, the following options are available:
Vorticity-based production
Strain/vorticity-based production
Viscous heating (always activated for the coupled solvers)
If you choose the standard k- model or the realizable k- model, the following options
are available:
Viscous heating (always activated for the coupled solvers)
Inclusion of buoyancy eects on
If you choose the RNG k- model, the following options are available:
Dierential viscosity model
Swirl modication
Viscous heating (always activated for the coupled solvers)
Inclusion of buoyancy eects on
If you choose the standard k- model, the following options are available:
Transitional ows
Shear ow corrections
Viscous heating (always activated for the coupled solvers)
11-70 c Fluent Inc. January 11, 2005
11.11 Problem Setup for Turbulent Flows
If you choose the shear-stress transport k- model, the following options are available:
Transitional ows
Viscous heating (always activated for the coupled solvers)
If you choose the RSM, the following options are available:
Wall reection eects on Reynolds stresses
Wall boundary conditions for the Reynolds stresses from the k equation
Quadratic pressure-strain model
Viscous heating (always activated for the coupled solvers)
Inclusion of buoyancy eects on
If you choose the enhanced wall treatment (available for the k- models and the RSM),
the following options are available:
Pressure gradient eects
Thermal eects
If you choose the DES model, the same options are available as for Spalart-Allmaras
model, since the Spalart-Allmaras model is used as the underlaying RANS model for DES
modeling approach. Additionally, you can perform the following DES-specic functions
by using the /define/models/viscous/detached-eddy-simulation? text command:
Use cell volume-based LES length scale (default is to use maximum cell edge).
Modify only the length scales that appear in the destruction term in
t
equation
(the default is to modify all length scales within the
t
equation).
If you choose the LES model, the following options are available:
Smagorinsky-Lilly model for the subgrid-scale viscosity
RNG model for the subgrid-scale viscosity
Viscous heating (always activated for the coupled solvers)
c Fluent Inc. January 11, 2005 11-71
Modeling Turbulence
It is also possible to modify the Model Constants, but this is not necessary for most
applications. See Sections 11.3 through 11.7 for details about these constants. Note that
C1-PS and C2-PS are the constants C
1
and C
2
in the linear pressure-strain approximation
of Equations 11.6-5 and 11.6-6, and C1-PS and C2-PS are the constants C

1
and C

2
in
Equation 11.6-7. C1-SSG-PS, C1-SSG-PS, C2-SSG-PS, C3-SSG-PS, C3-SSG-PS, C4-SSG-
PS, and C5-SSG-PS are the constants C
1
, C

1
, C
2
, C
3
, C

3
, C
4
, and C
5
in the quadratic
pressure-strain approximation of Equation 11.6-16.
Including the Viscous Heating Effects
See Sections 12.2.1 and 12.2.2 for information on including viscous heating eects in your
model.
Including Turbulence Generation Due to Buoyancy
If you specify a non-zero gravity force (in the Operating Conditions panel), and you are
modeling a non-isothermal ow, the generation of turbulent kinetic energy due to buoy-
ancy (G
b
in Equation 11.4-1) is, by default, always included in the k equation. However,
FLUENT does not, by default, include the buoyancy eects on .
To include the buoyancy eects on , you must turn on the Full Buoyancy Eects option
under Options in the Viscous Model panel.
This option is available for the three k- models and for the RSM.
Vorticity- and Strain/Vorticity-Based Production
For the Spalart-Allmaras model, you can choose either Vorticity-Based Production or
Strain/Vorticity-Based Production under Spalart-Allmaras Options in the Viscous Model
panel. If you choose Vorticity-Based Production, FLUENT will use Equation 11.3-8 to
compute the value of the deformation tensor S; if you choose Strain/Vorticity-Based Pro-
duction, it will use Equation 11.3-10.
(These options will not appear unless you have activated the Spalart-Allmaras model.)
Detached Eddy Simulation (DES) Modeling
If you enabled DES for the Spalart-Allmaras model as described at the beginning of this
section, FLUENT will use Equation 11.8-1 to compute the value of the length scale

d. By
default, the empirical constant C
des
is set to 0.65. You can change its value in the Cdes
eld under Model Constants.
11-72 c Fluent Inc. January 11, 2005
11.11 Problem Setup for Turbulent Flows
Differential Viscosity Modication
In the RNG turbulence model in FLUENT, you have an option to use a dierential formula
for eective viscosity
e
(Equation 11.4-6) to account for low-Reynolds-number eects.
To enable this option, turn on Dierential Viscosity Model under RNG Options in the
Viscous Model panel.
(This option will not appear unless you have activated the RNG k- model.)
Swirl Modication
Once you choose the RNG model, the swirl modication takes eect, by default, for all
three-dimensional ows and axisymmetric ows with swirl. The default swirl constant
(
s
in Equation 11.4-8) is set to 0.05, which works well for weakly to moderately swirling
ows. However, for strongly swirling ows, you may need to use a larger swirl constant.
To change the value of the swirl constant, you must rst turn on the Swirl Dominated
Flow option under RNG Options in the Viscous Model panel. (This option will not appear
unless you have activated the RNG k- model.)
Once you turn on the Swirl Dominated Flow option, the swirl constant
s
is increased to
0.07. You can change its value in the Swirl Factor eld under Model Constants.
Transitional Flows
If either of the k- models are used, you may enable a low-Reynolds-number correction to
the turbulent viscosity by enabling the Transitional Flows option under k-omega Options
in the Viscous Model panel. By default, this option is not enabled, and the damping
coecient (

in Equation 11.5-6) is equal to 1.


Shear Flow Corrections
In the standard k- model, you also have the option of including corrections to improve
the accuracy in predicting free shear ows. The Shear Flow Corrections option under
k-omega Options is enabled by default in the Viscous Model panel, as these corrections
are included in the standard k- model [378]. When this option is enabled, FLUENT will
calculate f

and f

using Equations 11.5-16 and 11.5-24, respectively. If this option is


disabled, f

and f

will be set equal to 1.


Including Pressure Gradient Effects
If the enhanced wall treatment is used, you may include the eects of pressure gradients
by enabling the Pressure Gradient Eects option under Enhanced Wall Treatment Options.
When this option is enabled, FLUENT will include the coecient in Equation 11.9-31.
c Fluent Inc. January 11, 2005 11-73
Modeling Turbulence
Including Thermal Effects
If the enhanced wall treatment is used, you may include thermal eects by enabling
the Thermal Eects option under Enhanced Wall Treatment Options. When this option
is enabled, FLUENT will include the coecient in Equation 11.9-31. will also be
included in Equation 11.9-31 when the Thermal Eects option is enabled if the ideal gas
law is selected for the uid density in the Materials panel.
Including the Wall Reection Term
If the RSM is used with the default model for pressure strain, FLUENT will, by default,
include the wall-reection eects in the pressure-strain term. That is, FLUENT will
calculate
ij,w
using Equation 11.6-7 and include it in Equation 11.6-4. Note that wall-
reection eects are not included if you have selected the quadratic pressure-strain model.
i
The empirical constants and the function f used in the calculation of
ij,w
are calibrated for simple canonical ows such as channel ows and at-plate
boundary layers involving a single wall. If the ow involves multiple walls
and the wall has signicant curvature (e.g., an axisymmetric pipe or curvi-
linear duct), the inclusion of the wall-reection term in Equation 11.6-7
may not improve the accuracy of the RSM predictions. In such cases,
you can disable the wall-reection eects by turning o the Wall Reection
Eects under Reynolds-Stress Options in the Viscous Model panel.
Solving the k Equation to Obtain Wall Boundary Conditions
In the RSM, FLUENT, by default, uses the explicit setting of boundary conditions for
the Reynolds stresses near the walls, with the values computed with Equation 11.6-28.
k is calculated by solving the k equation obtained by summing Equation 11.6-1 for
normal stresses. To disable this option and use the wall boundary conditions given in
Equation 11.6-29, turn o Wall B.C. from k Equation under Reynolds-Stress Options in the
Viscous Model panel. (This option will not appear unless you have activated the RSM.)
Quadratic Pressure-Strain Model
To use the quadratic pressure-strain model described in Section 11.6.3: Quadratic Pressure-
Strain Model, turn on the Quadratic Pressure-Strain Model option under Reynolds-Stress
Options in the Viscous Model panel. (This option will not appear unless you have activated
the RSM.) The following options are not available when the Quadratic Pressure-Strain
Model is enabled:
Wall Reection Eects under Reynolds-Stress Options
Enhanced Wall Treatment under Near-Wall Treatment
11-74 c Fluent Inc. January 11, 2005
11.11 Problem Setup for Turbulent Flows
Subgrid-Scale Model
If you have selected the Large Eddy Simulation model, you will be able to choose which of
the subgrid-scale models described in Section 11.7.2: Subgrid-Scale Models is to be used.
You can choose from the Smagorinsky-Lilly, WALE, or Kinetic-Energy Transport subgrid-
scale models. Note that the Dynamic Model is an option with the Smagorinsky-Lilly model,
while Kinetic-Energy Transport model is always run as a dynamic model.
(These options will not appear unless you have activated the LES model.)
Customizing the Turbulent Viscosity
If you are using the Spalart-Allmaras, k-, k-, or LES model, a user-dened function
can be used to customize the turbulent viscosity. This option will enable you to modify

t
in the case of the Spalart-Allmaras, k-, and k- models, and incorporate completely
new subgrid models in the case of the LES model. See the separate UDF Manual for
information about user-dened functions.
In the Viscous Model panel, under User-Dened Functions, select the appropriate user-
dened function in the Turbulent Viscosity drop-down list. For the LES model, select the
appropriate UDF in the Subgrid-Scale Turbulent Viscosity drop-down list.
Customizing the Turbulent Prandtl Numbers
If you are using the standard or realizable k- model or the standard k- model, a user-
dened function can be used to customize the turbulent Prandtl numbers. This option
will allow you to calculate
k
and either

or

(depending on if you have enabled the


appropriate k- or k- model) by using a UDF. You will also be able to calculate the value
of the energy Prandtl number (Pr
t
in Equation 11.4-23) and the Prandtl number at the
wall (Pr
t
in Equation 11.9-5) in this way. See the separate UDF Manual for information
about user-dened functions.
In the Viscous Model panel, under User-Dened Functions, select the appropriate user-
dened function from the drop-down lists under Prandtl Numbers. Options include: TKE
Prandtl Number, TDR Prandtl Number (k- models only), SDR Prandtl Number (k- model
only), Energy Prandtl Number, and Wall Prandtl Number.
Modeling Turbulence with Non-Newtonian Fluids
If the turbulent ow involves non-Newtonian uids, you can use the define/models/
viscous/turbulence-expert/turb-non-newtonian? text command to enable the se-
lection of non-Newtonian options for the material viscosity. See Section 8.4.5: Viscosity
for Non-Newtonian Fluids for details about these options.
c Fluent Inc. January 11, 2005 11-75
Modeling Turbulence
11.11.2 Dening Turbulence Boundary Conditions
k- Models and k- Models
When you are modeling turbulent ows in FLUENT using one of the k- models or one
of the k- models, you must provide the boundary conditions for k and (or k and )
in addition to other mean solution variables. The boundary conditions for k and (or k
and ) at the walls are internally taken care of by FLUENT, which obviates the need for
your inputs. The boundary condition inputs for k and (or k and ) you must supply
to FLUENT are the ones at inlet boundaries (velocity inlet, pressure inlet, etc.). In many
situations, it is important to specify correct or realistic boundary conditions at the inlets,
because the inlet turbulence can signicantly aect the downstream ow.
See Section 7.2.2: Determining Turbulence Parameters for details about specifying the
boundary conditions for k and (or k and ) at the inlets.
You may want to include the eects of the wall roughness on selected wall boundaries.
In such cases, you can specify the roughness parameters (roughness height and roughness
constant) in the panels for the corresponding wall boundaries (see Section 7.13.1: Setting
the Roughness Parameters).
Additionally, you can control whether or not to set the turbulent viscosity to zero within
a laminar zone. If the uid zone in question is laminar, the text command define/
boundary-conditions/fluid will contain an option called Set Turbulent Viscosity
to zero within laminar zone?. By setting this option to yes, FLUENT will set both
the production term in the turbulence transport equation and
t
to zero. In contrast,
when the Laminar Zone option is turned on in a Fluid boundary condition panel, only the
production term is set to zero. See Section 7.17.1: Specifying a Laminar Zone for details
about laminar zones.
i
Note that the laminar zone feature is also available for the Spalart-Allmaras
and RSM models.
The Spalart-Allmaras Model
When you are modeling turbulent ows in FLUENT using the Spalart-Allmaras model,
you must provide the boundary conditions for in addition to other mean solution vari-
ables. The boundary conditions for at the walls are internally taken care of by FLUENT,
which obviates the need for your inputs. The boundary condition input for you must
supply to FLUENT is the one at inlet boundaries (velocity inlet, pressure inlet, etc.). In
many situations, it is important to specify correct or realistic boundary conditions at the
inlets, because the inlet turbulence can signicantly aect the downstream ow.
See Section 7.2.2: Determining Turbulence Parameters for details about specifying the
boundary condition for at the inlets.
11-76 c Fluent Inc. January 11, 2005
11.11 Problem Setup for Turbulent Flows
You may want to include the eects of the wall roughness on selected wall boundaries.
In such cases, you can specify the roughness parameters (roughness height and roughness
constant) in the panels for the corresponding wall boundaries (see Section 7.13.1: Setting
the Roughness Parameters).
i
Note that if the DES model is enabled, the wall boundary conditions will be
treated the same as for the Spalart-Allmaras model. All other boundaries
will be treated the same as for the LES model (see below for details about
LES boundary conditions).
Reynolds Stress Model
The specication of turbulent boundary conditions for the RSM is the same as for the
other turbulence models for all boundaries except at boundaries where ow enters the
domain. Additional input methods are available for these boundaries and are described
here.
When you choose to use the RSM, the default inlet boundary condition inputs required
are identical to those required when the k- model is active. You can input the tur-
bulence quantities using any of the turbulence specication methods described in Sec-
tion 7.2.2: Determining Turbulence Parameters. FLUENT then uses the specied tur-
bulence quantities to derive the Reynolds stresses at the inlet from the assumption of
isotropy of turbulence:
u

2
i
=
2
3
k (i = 1, 2, 3) (11.11-1)
u

i
u

j
= 0.0 (11.11-2)
where u

2
i
is the normal Reynolds stress component in each direction. The boundary
condition for is determined in the same manner as for the k- turbulence models (see
Section 7.2.2: Determining Turbulence Parameters). To use this method, you will select
K or Turbulence Intensity as the Reynolds-Stress Specication Method in the appropriate
boundary condition panel.
Alternately, you can directly specify the Reynolds stresses by selecting Reynolds-Stress
Components as the Reynolds-Stress Specication Method in the boundary condition panel.
When this option is enabled, you should input the Reynolds stresses directly.
You can set the Reynolds stresses by using constant values, prole functions of coordinates
(see Section 7.26: Boundary Proles), or user-dened functions (see the separate UDF
Manual).
c Fluent Inc. January 11, 2005 11-77
Modeling Turbulence
Figure 11.11.2: Specifying Inlet Boundary Conditions for the Reynolds
Stresses
11-78 c Fluent Inc. January 11, 2005
11.11 Problem Setup for Turbulent Flows
Large Eddy Simulation Model
It is possible to specify the magnitude of random uctuations of the velocity components
at an inlet only if the velocity inlet boundary condition is selected. In this case, you must
specify a Turbulence Intensity that determines the magnitude of the random perturbations
on individual mean velocity components as described in Section 11.7.3: Inlet Boundary
Conditions for the LES Model. For all boundary types other than velocity inlets, the
boundary conditions for LES remain the same as for laminar ows.
11.11.3 Providing an Initial Guess for k and (or k and )
For ows using one of the k- models, one of the k- models, or the RSM, the converged
solutions or (for unsteady calculations) the solutions after a suciently long time has
elapsed should be independent of the initial values for k and (or k and ). For better
convergence, however, it is benecial to use a reasonable initial guess for k and (or k
and ).
In general, it is recommended that you start from a fully-developed state of turbulence.
When you use the enhanced wall treatment for the k- models or the RSM, it is critically
important to specify fully-developed turbulence elds. Guidelines are provided below.
If you were able to specify reasonable boundary conditions at the inlet, it may be
a good idea to compute the initial values for k and (or k and ) in the whole
domain from these boundary values. (See Section 26.15: Initializing the Solution
for details.)
For more complex ows (e.g., ows with multiple inlets with dierent conditions) it
may be better to specify the initial values in terms of turbulence intensity. 510%
is enough to represent fully-developed turbulence. k can then be computed from
the turbulence intensity and the characteristic mean velocity magnitude of your
problem (k = 1.5(Iu
avg
)
2
).
You should specify an initial guess for so that the resulting eddy viscosity (C

k
2

)
is suciently large in comparison to the molecular viscosity. In fully-developed
turbulence, the turbulent viscosity is roughly two orders of magnitude larger than
the molecular viscosity. From this, you can compute .
Note that, for the RSM, Reynolds stresses are initialized automatically using Equa-
tions 11.11-1 and 11.11-2.
c Fluent Inc. January 11, 2005 11-79
Modeling Turbulence
11.12 Solution Strategies for Turbulent Flow Simulations
Compared to laminar ows, simulations of turbulent ows are more challenging in many
ways. For the Reynolds-averaged approach, additional equations are solved for the tur-
bulence quantities. Since the equations for mean quantities and the turbulent quantities
(
t
, k, , , or the Reynolds stresses) are strongly coupled in a highly non-linear fashion,
it takes more computational eort to obtain a converged turbulent solution than to ob-
tain a converged laminar solution. The LES model, while embodying a simpler, algebraic
model for the subgrid-scale viscosity, requires a transient solution on a very ne mesh.
The delity of the results for turbulent ows is largely determined by the turbulence
model being used. Here are some guidelines that can enhance the quality of your turbulent
ow simulations.
11.12.1 Mesh Generation
The following are suggestions to follow when generating the mesh for use in your turbulent
ow simulation:
Picture in your mind the ow under consideration using your physical intuition or
any data for a similar ow situation, and identify the main ow features expected
in the ow you want to model. Generate a mesh that can resolve the major features
that you expect.
If the ow is wall-bounded, and the wall is expected to signicantly aect the ow,
take additional care when generating the mesh. You should avoid using a mesh
that is too ne (for the wall function approach) or too coarse (for the enhanced
wall treatment approach). See Section 11.10: Grid Considerations for Turbulent
Flow Simulations for details.
11.12.2 Accuracy
The suggestions below are provided to help you obtain better accuracy in your results:
Use the turbulence model that is better suited for the salient features you expect
to see in the ow (see Section 11.2: Choosing a Turbulence Model).
Because the mean quantities have larger gradients in turbulent ows than in laminar
ows, it is recommended that you use high-order schemes for the convection terms.
This is especially true if you employ a triangular or tetrahedral mesh. Note that
excessive numerical diusion adversely aects the solution accuracy, even with the
most elaborate turbulence model.
In some ow situations involving inlet boundaries, the ow downstream of the inlet
is dictated by the boundary conditions at the inlet. In such cases, you should
exercise care to make sure that reasonably realistic boundary values are specied.
11-80 c Fluent Inc. January 11, 2005
11.12 Solution Strategies for Turbulent Flow Simulations
11.12.3 Convergence
The suggestions below are provided to help you enhance convergence for turbulent ow
calculations:
Starting with excessively crude initial guesses for mean and turbulence quantities
may cause the solution to diverge. A safe approach is to start your calculation using
conservative (small) under-relaxation parameters and (for the coupled solvers) a
conservative Courant number, and increase them gradually as the iterations proceed
and the solution begins to settle down.
It is also helpful for faster convergence to start with reasonable initial guesses for
the k and (or k and ) elds. Particularly when the enhanced wall treatment
is used, it is important to start with a suciently developed turbulence eld, as
recommended in Section 11.11.3: Providing an Initial Guess for k and (or k and
), to avoid the need for an excessive number of iterations to develop the turbulence
eld.
When you are using the RNG k- model, an approach that might help you achieve
better convergence is to obtain a solution with the standard k- model before switch-
ing to the RNG model. Due to the additional non-linearities in the RNG model,
lower under-relaxation factors and (for the coupled solvers) a lower Courant number
might also be necessary.
Note that when you use the enhanced wall treatment, you may sometimes nd during
the calculation that the residual for is reported to be zero. This happens when your
ow is such that Re
y
is less than 200 in the entire ow domain, and is obtained from
the algebraic formula (Equation 11.9-25) instead of from its transport equation.
11.12.4 RSM-Specic Solution Strategies
Using the RSM creates a high degree of coupling between the momentum equations and
the turbulent stresses in the ow, and thus the calculation can be more prone to stability
and convergence diculties than with the k- models. When you use the RSM, therefore,
you may need to adopt special solution strategies in order to obtain a converged solution.
The following strategies are generally recommended:
Begin the calculations using the standard k- model. Turn on the RSM and use
the k- solution data as a starting point for the RSM calculation.
Use low under-relaxation factors (0.2 to 0.3) and (for the coupled solvers) a low
Courant number for highly swirling ows or highly complex ows. In these cases,
you may need to reduce the under-relaxation factors both for the velocities and for
all of the stresses.
c Fluent Inc. January 11, 2005 11-81
Modeling Turbulence
Instructions for setting these solution parameters are provided below. If you are applying
the RSM to prediction of a highly swirling ow, you will want to consider the solution
strategies discussed in Section 9.4: Swirling and Rotating Flows as well.
Under-Relaxation of the Reynolds Stresses
FLUENT applies under-relaxation to the Reynolds stresses. You can set under-relaxation
factors using the Solution Controls panel.
Solve Controls Solution...
The default settings of 0.5 are recommended for most cases. You may be able to increase
these settings and speed up the convergence when the RSM solution begins to converge.
Disabling Calculation Updates of the Reynolds Stresses
In some instances, you may wish to let the current Reynolds stress eld remain xed,
skipping the solution of the Reynolds transport equations while solving the other trans-
port equations. You can activate/deactivate all Reynolds stress equations in the Solution
Controls panel.
Solve Controls Solution...
Residual Reporting for the RSM
When you use the RSM for turbulence, FLUENT reports the equation residuals for the
individual Reynolds stress transport equations. You can apply the usual convergence
criteria to the Reynolds stress residuals: normalized residuals in the range of 10
3
usu-
ally indicate a practically-converged solution. However, you may need to apply tighter
convergence criteria (below 10
4
) to ensure full convergence.
11.12.5 LES-Specic Solution Strategies
Large eddy simulation involves running a transient solution from some initial condition,
on an appropriately ne grid, using an appropriate time step size. The solution must
be run long enough to become independent of the initial condition and to enable the
statistics of the ow eld to be determined.
The following are suggestions to follow when running a large eddy simulation:
1. Start by running a steady state ow simulation using a Reynolds-averaged tur-
bulence model such as standard k-, k-, Spalart-Allmaras, or even RSM. Run
until the ow eld is reasonably converged and then use the solve/initialize/
init-instantaneous-vel text command to generate the instantaneous velocity
11-82 c Fluent Inc. January 11, 2005
11.12 Solution Strategies for Turbulent Flow Simulations
eld out of the steady-state RANS results. This command must be executed be-
fore LES is enabled. This option is available for all RANS-based models and it will
create a much more realistic initial eld for the LES run. Additionally, it will help
in reducing the time needed for the LES simulation to reach a statistically stable
mode. This step is optional.
2. When you enable LES, FLUENT will automatically turn on the unsteady solver
option and choose the second-order implicit formulation. You will need to set
the appropriate time step size and all the needed solution parameters. (See Sec-
tion 26.18.1: User Inputs for Time-Dependent Problems for guidelines on setting
solution parameters for transient calculations in general.) The bounded central-
dierencing spatial discretization scheme will be automatically enabled for momen-
tum equations. Both the bounded central-dierencing and pure central-dierencing
schemes are available for all equations when running LES simulations.
3. Run LES until the ow becomes statistically steady. The best way to see if the ow
is fully developed and statistically steady is to monitor forces and solution variables
(e.g., velocity components or pressure) at selected locations in the ow.
4. Zero out the initial statistics using the solve/initialize/init-flow-statistics
text command. Before you restart the solution, enable Data Sampling for Time
Statistics in the Iterate panel, as described in Section 26.18.1: User Inputs for Time-
Dependent Problems.
5. Continue until you get statistically stable data. The duration of the simulation
can be determined beforehand by estimating the mean ow residence time in the
solution domain (L/U, where L is the characteristic length of the solution domain
and U is a characteristic mean ow velocity). The simulation should be run for at
least a few mean ow residence times.
Instructions for setting the solution parameters for LES are provided below.
Temporal Discretization
FLUENT provides both rst-order and second-order temporal discretizations. For LES,
the second-order discretization is recommended.
Dene Models Solver...
c Fluent Inc. January 11, 2005 11-83
Modeling Turbulence
Spatial Discretization
Overly diusive schemes such as the rst-order upwind or power law scheme should be
avoided, because they may unduly damp out the energy of the resolved eddies. The
central-dierencing based schemes are recommended for all equations when you use the
LES model. FLUENT provides two central-dierencing based schemes: pure central-
dierencing and bounded central-dierencing. The bounded scheme is the default option
when you select LES or DES.
Solve Controls Solution...
11.13 Postprocessing for Turbulent Flows
FLUENT provides postprocessing options for displaying, plotting, and reporting vari-
ous turbulence quantities, which include the main solution variables and other auxiliary
quantities.
Turbulence quantities that can be reported for the k- models are as follows:
Turbulent Kinetic Energy (k)
Turbulence Intensity
Turbulent Dissipation Rate (Epsilon)
Production of k
Turbulent Viscosity
Eective Viscosity
Turbulent Viscosity Ratio
Eective Thermal Conductivity
Eective Prandtl Number
Wall Yplus
Wall Ystar
Turbulent Reynolds Number (Re y) (only when the enhanced wall treatment is used
for the near-wall treatment)
11-84 c Fluent Inc. January 11, 2005
11.13 Postprocessing for Turbulent Flows
Turbulence quantities that can be reported for the k- models are as follows:
Turbulent Kinetic Energy (k)
Turbulence Intensity
Specic Dissipation Rate (Omega)
Production of k
Turbulent Viscosity
Eective Viscosity
Turbulent Viscosity Ratio
Eective Thermal Conductivity
Eective Prandtl Number
Wall Ystar
Wall Yplus
Turbulence quantities that can be reported for the Spalart-Allmaras model are as follows:
Modied Turbulent Viscosity
Turbulent Viscosity
Eective Viscosity
Turbulent Viscosity Ratio
Eective Thermal Conductivity
Eective Prandtl Number
Wall Yplus
Turbulence quantities that can be reported for the RSM are as follows:
Turbulent Kinetic Energy (k)
Turbulence Intensity
UU Reynolds Stress
c Fluent Inc. January 11, 2005 11-85
Modeling Turbulence
VV Reynolds Stress
WW Reynolds Stress
UV Reynolds Stress
VW Reynolds Stress
UW Reynolds Stress
Turbulent Dissipation Rate (Epsilon)
Production of k
Turbulent Viscosity
Eective Viscosity
Turbulent Viscosity Ratio
Eective Thermal Conductivity
Eective Prandtl Number
Wall Yplus
Wall Ystar
Turbulent Reynolds Number (Re y)
Turbulence quantities that can be reported for the DES model are as follows:
Modied Turbulent Viscosity
Turbulent Viscosity
Eective Viscosity
Turbulent Viscosity Ratio
Eective Thermal Conductivity
Eective Prandtl Number
Wall Yplus
Relative Length Scale (DES)
11-86 c Fluent Inc. January 11, 2005
11.13 Postprocessing for Turbulent Flows
Turbulence quantities that can be reported for the LES model are as follows:
Subgrid Turbulent Kinetic Energy
Subgrid Turbulent Viscosity
Subgrid Eective Viscosity
Subgrid Turbulent Viscosity Ratio
Eective Thermal Conductivity
Eective Prandtl Number
Wall Yplus
All of these variables can be found in the Turbulence... category of the variable selection
drop-down list that appears in postprocessing panels. See Chapter 31: Field Function
Denitions for their denitions.
11.13.1 Custom Field Functions for Turbulence
In addition to the quantities listed above, you can dene your own turbulence quantities
using the Custom Field Function Calculator panel.
Dene Custom Field Functions...
The following functions may be useful:
Ratio of production of k to its dissipation (G
k
/)
Ratio of the mean ow to turbulent time scale, ( Sk/)
Reynolds stresses derived from the Boussinesq formula (e.g., uv =
t
u
y
)
11.13.2 Postprocessing LES Statistics
As described in Section 11.7: The Large Eddy Simulation (LES) Model, LES involves
the solution of a transient ow eld, but it is the mean ow quantities that are of most
engineering interest. If you turn on the Data Sampling for Time Statistics option in the
Iterate panel, FLUENT will gather data for time statistics while performing a large eddy
simulation. You can then view both the mean and the root-mean-square (RMS) values in
FLUENT. See Section 26.18.4: Postprocessing for Time-Dependent Problems for details.
c Fluent Inc. January 11, 2005 11-87
Modeling Turbulence
11.13.3 Troubleshooting
You can use the postprocessing options not only for the purpose of interpreting your
results but also for investigating any anomalies that may appear in the solution. For
instance, you may want to plot contours of the k eld to check if there are any regions
where k is erroneously large or small. You should see a high k region in the region
where the production of k is large. You may want to display the turbulent viscosity
ratio eld in order to see whether or not turbulence takes full eect. Usually turbulent
viscosity is at least two orders of magnitude larger than molecular viscosity for fully-
developed turbulent ows modeled using the RANS approach (i.e., not using LES). You
may also want to see whether you are using a proper near-wall mesh for the enhanced
wall treatment. In this case, you can display lled contours of Re
y
(turbulent Reynolds
number) overlaid on the mesh.
11-88 c Fluent Inc. January 11, 2005

You might also like