You are on page 1of 82

Chapter 10.

Modeling Turbulence

This chapter provides details about the turbulence models available in FLUENT.
Information is presented in the following sections:

• Section 10.1: Introduction

• Section 10.2: Choosing a Turbulence Model

• Section 10.3: The Spalart-Allmaras Model

• Section 10.4: The Standard, RNG, and Realizable k- Models

• Section 10.5: The Standard and Shear-Stress Transport (SST) k-ω Models

• Section 10.6: The Reynolds Stress Model (RSM)

• Section 10.7: The Large Eddy Simulation (LES) Model

• Section 10.8: Near-Wall Treatments for Wall-Bounded Turbulent Flows

• Section 10.9: Grid Considerations for Turbulent Flow Simulations

• Section 10.10: Problem Setup for Turbulent Flows

• Section 10.11: Solution Strategies for Turbulent Flow Simulations

• Section 10.12: Postprocessing for Turbulent Flows

10.1 Introduction
Turbulent flows are characterized by fluctuating velocity fields. These fluctuations mix
transported quantities such as momentum, energy, and species concentration, and cause
the transported quantities to fluctuate as well. Since these fluctuations can be of small
scale and high frequency, they are too computationally expensive to simulate directly in
practical engineering calculations. Instead, the instantaneous (exact) governing equations
can be time-averaged, ensemble-averaged, or otherwise manipulated to remove the small
scales, resulting in a modified set of equations that are computationally less expensive
to solve. However, the modified equations contain additional unknown variables, and
turbulence models are needed to determine these variables in terms of known quantities.


c Fluent Inc. January 28, 2003 10-1
Modeling Turbulence

FLUENT provides the following choices of turbulence models:

• Spalart-Allmaras model
• k- models
– Standard k- model
– Renormalization-group (RNG) k- model
– Realizable k- model
• k-ω models
– Standard k-ω model
– Shear-stress transport (SST) k-ω model
• v 2 -f model
• Reynolds stress model (RSM)
• Large eddy simulation (LES) model

10.2 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 flow, the established practice for
a specific 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 effort and cost in terms of CPU time and
memory of the individual models is discussed. While it is impossible to state categorically
which model is best for a specific application, general guidelines are presented to help
you choose the appropriate turbulence model for the flow you want to model.

10.2.1 Reynolds-Averaged Approach vs. LES


A complete time-dependent solution of the exact Navier-Stokes equations for high-Reynolds-
number turbulent flows in complex geometries is unlikely to be attainable for some time
to come. Two alternative methods can be employed to transform the Navier-Stokes equa-
tions in such a way that the small-scale turbulent fluctuations do not have to be directly
simulated: Reynolds averaging and filtering. Both methods introduce additional terms in
the governing equations that need to be modeled in order to achieve “closure”. (Closure
implies that there are a sufficient number of equations for all the unknowns.)

10-2
c Fluent Inc. January 28, 2003
10.2 Choosing a Turbulence Model

The Reynolds-averaged Navier-Stokes (RANS) equations represent transport equations


for the mean flow quantities only, with all the scales of the turbulence being modeled. The
approach of permitting a solution for the mean flow variables greatly reduces the compu-
tational effort. If the mean flow is steady, the governing equations will not contain time
derivatives and a steady-state solution can be obtained economically. A computational
advantage is seen even in transient situations, since the time step will be determined by
the global unsteadiness in the mean flow rather than by the turbulence. The Reynolds-
averaged approach is generally adopted for practical engineering calculations, and uses
models such as Spalart-Allmaras, k- and its variants, k-ω and its variants, and the RSM.
LES provides an alternative approach in which the large eddies are computed in a time-
dependent simulation that uses a set of “filtered” equations. Filtering is essentially a
manipulation of the exact Navier-Stokes equations to remove only the eddies that are
smaller than the size of the filter, which is usually taken as the mesh size. Like Reynolds
averaging, the filtering process creates additional unknown terms that must be modeled
in order to achieve closure. Statistics of the mean flow quantities, which are generally
of most engineering interest, are gathered during the time-dependent simulation. The
attraction of LES is that, by modeling less of the turbulence (and solving more), the
error induced by the turbulence model will be reduced. One might also argue that it
ought to be easier to find a “universal” model for the small scales, which tend to be more
isotropic and less affected by the macroscopic flow features than the large eddies.
It should, however, be stressed that the application of LES to industrial fluid simulations
is in its infancy. As highlighted in a recent review publication [81], typical applications
to date have been for simple geometries. This is mainly because of the large computer
resources required to resolve the energy-containing turbulent eddies. Most successful
LES has been done using high-order spatial discretization, with great care being taken
to resolve all scales larger than the inertial subrange. The degradation of accuracy in the
mean flow quantities with poorly resolved LES is not well documented. In addition, the
use of wall functions with LES is an approximation that requires further validation.
As a general guideline, therefore, 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 10.7, has been made available for you to
try if you have the computational resources and are willing to invest the effort. The rest
of this section will deal with the choice of models using the Reynolds-averaged approach.

10.2.2 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 fluc-
tuating components. For the velocity components:

ui = ūi + u0i (10.2-1)


c Fluent Inc. January 28, 2003 10-3
Modeling Turbulence

where ūi and u0i are the mean and fluctuating velocity components (i = 1, 2, 3).
Likewise, for pressure and other scalar quantities:

φ = φ̄ + φ0 (10.2-2)

where φ denotes a scalar such as pressure, energy, or species concentration.


Substituting expressions of this form for the flow variables into the instantaneous conti-
nuity and momentum equations and taking a time (or ensemble) average (and dropping
the overbar on the mean velocity, ū) yields the ensemble-averaged momentum equations.
They can be written in Cartesian tensor form as:

∂ρ ∂
+ (ρui ) = 0 (10.2-3)
∂t ∂xi

" !#
∂ ∂ ∂p ∂ ∂ui ∂uj 2 ∂ul ∂
(ρui ) + (ρui uj ) = − + µ + − δij + (−ρu0i u0j )
∂t ∂xj ∂xi ∂xj ∂xj ∂xi 3 ∂xl ∂xj
(10.2-4)
Equations 10.2-3 and 10.2-4 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 effects of tur-
bulence. These Reynolds stresses, −ρu0i u0j , must be modeled in order to close Equa-
tion 10.2-4.
For variable-density flows, Equations 10.2-3 and 10.2-4 can be interpreted as Favre-
averaged Navier-Stokes equations [102], with the velocities representing mass-averaged
values. As such, Equations 10.2-3 and 10.2-4 can be applied to density-varying flows.

10.2.3 Boussinesq Approach vs. Reynolds Stress Transport


Models
The Reynolds-averaged approach to turbulence modeling requires that the Reynolds
stresses in Equation 10.2-4 be appropriately modeled. A common method employs the
Boussinesq hypothesis [102] to relate the Reynolds stresses to the mean velocity gradients:
! !
∂ui ∂uj 2 ∂ui
− ρu0i u0j = µt + − ρk + µt δij (10.2-5)
∂xj ∂xi 3 ∂xi

The Boussinesq hypothesis is used in the Spalart-Allmaras model, the k- models, and
the k-ω models. The advantage of this approach is the relatively low computational

10-4
c Fluent Inc. January 28, 2003
10.2 Choosing a Turbulence Model

cost associated with the computation of the turbulent viscosity, µt . In the case of the
Spalart-Allmaras model, only one additional transport equation (representing turbulent
viscosity) is solved. In the case of the k- and k-ω models, two additional transport
equations (for the turbulence kinetic energy, k, and either the turbulence dissipation
rate, , or the specific dissipation rate, ω) are solved, and µt is computed as a function of
k and . The disadvantage of the Boussinesq hypothesis as presented is that it assumes
µt is an isotropic scalar quantity, which is not strictly true.
The alternative approach, embodied in the RSM, is to solve transport equations for each
of the terms in the Reynolds stress tensor. An additional scale-determining equation
(normally for ) is also required. This means that five additional transport equations are
required in 2D flows and seven additional transport equations must be solved in 3D.
In many cases, models based on the Boussinesq hypothesis perform very well, and the
additional computational expense of the Reynolds stress model is not justified. However,
the RSM is clearly superior for situations in which the anisotropy of turbulence has a
dominant effect on the mean flow. Such cases include highly swirling flows and stress-
driven secondary flows.

10.2.4 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 specifically for aerospace applications involving wall-bounded flows and has been
shown to give good results for boundary layers subjected to adverse pressure gradients.
It is also gaining popularity for turbomachinery applications.
In its original form, the Spalart-Allmaras model is effectively a low-Reynolds-number
model, requiring the viscous-affected 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 sufficiently fine. This might make it the best
choice for relatively crude simulations on coarse meshes where accurate turbulent flow
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-ω models. This might make the model less sensitive to numerical error
when non-layered meshes are used near walls. See Section 5.1.2 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 flows. 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 flow changes


c Fluent Inc. January 28, 2003 10-5
Modeling Turbulence

abruptly from a wall-bounded to a free shear flow.

The Detached Eddy Simulation (DES) Model


The detached eddy simulation (DES) model is a modified version of the Spalart-Allmaras
model and can be considered a more practical alternative to LES for predicting the
flow around high-Reynolds-number, high-lift airfoils. The DES approach combines an
unsteady RANS version of the Spalart-Allmaras model with a filtered version of the
same model to create two separate regions inside the flow 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 effects prevail, the standard RANS model is recovered.

10.2.5 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 flow
calculations in the time since it was proposed by Launder and Spalding [140]. Robust-
ness, economy, and reasonable accuracy for a wide range of turbulent flows explain its
popularity in industrial flow and heat transfer simulations. It is a semi-empirical model,
and the derivation of the model equations relies on phenomenological considerations and
empiricism.
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 [299] and the realizable k- model [232].

10.2.6 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 refinements:

• The RNG model has an additional term in its  equation that significantly improves
the accuracy for rapidly strained flows.
• The effect of swirl on turbulence is included in the RNG model, enhancing accuracy
for swirling flows.
• The RNG theory provides an analytical formula for turbulent Prandtl numbers,
while the standard k- model uses user-specified, constant values.

10-6
c Fluent Inc. January 28, 2003
10.2 Choosing a Turbulence Model

• While the standard k- model is a high-Reynolds-number model, the RNG theory
provides an analytically-derived differential formula for effective viscosity that ac-
counts for low-Reynolds-number effects. Effective use of this feature does, however,
depend on an appropriate treatment of the near-wall region.

These features make the RNG k- model more accurate and reliable for a wider class of
flows than the standard k- model.

10.2.7 The Realizable k- Model


The realizable k- model is a relatively recent development and differs 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 fluctuation.

The term “realizable” means that the model satisfies certain mathematical constraints
on the Reynolds stresses, consistent with the physics of turbulent flows. Neither the
standard k- model nor the RNG k- model is realizable.
An immediate benefit of the realizable k- model is that it more accurately predicts
the spreading rate of both planar and round jets. It is also likely to provide superior
performance for flows involving rotation, boundary layers under strong adverse pressure
gradients, separation, and recirculation.
Both the realizable and RNG k- models have shown substantial improvements over the
standard k- model where the flow features include strong streamline curvature, vortices,
and rotation. Since the model is still relatively new, it is not clear in exactly which
instances the realizable k- model consistently outperforms the RNG model. However,
initial studies have shown that the realizable model provides the best performance of all
the k- model versions for several validations of separated flows and flows with complex
secondary flow features.
One limitation of the realizable k- model is that it produces non-physical turbulent
viscosities in situations when the computational domain contains both rotating and sta-
tionary fluid zones (e.g., multiple reference frames, rotating sliding meshes). This is due
to the fact that the realizable k- model includes the effects of mean rotation in the
definition of the turbulent viscosity (see Equations 10.4-17–10.4-19). This extra rotation
effect has been tested on single rotating reference frame systems and showed superior
behavior over the standard k- model. However, due to the nature of this modification,
its application to multiple reference frame systems should be taken with some caution.
See Section 10.4.3 for information about how to include or exclude this term from the
model.


c Fluent Inc. January 28, 2003 10-7
Modeling Turbulence

10.2.8 The Standard k-ω Model


The standard k-ω model in FLUENT is based on the Wilcox k-ω model [294], which
incorporates modifications for low-Reynolds-number effects, compressibility, and shear
flow spreading. The Wilcox model predicts free shear flow 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 flows and free shear flows. A variation
of the standard k-ω model called the SST k-ω model is also available in FLUENT, and is
described in Section 10.2.9.

10.2.9 The Shear-Stress Transport (SST) k-ω Model


The shear-stress transport (SST) k-ω model was developed by Menter [166] to effectively
blend the robust and accurate formulation of the k-ω model in the near-wall region with
the free-stream independence of the k- model in the far field. To achieve this, the k-
model is converted into a k-ω formulation. The SST k-ω model is similar to the standard
k-ω model, but includes the following refinements:

• The standard k-ω model and the transformed k- model are both multiplied by a
blending function and both models are added together. The blending function is
designed to be one in the near-wall region, which activates the standard k-ω model,
and zero away from the surface, which activates the transformed k- model.

• The SST model incorporates a damped cross-diffusion derivative term in the ω


equation.

• The definition of the turbulent viscosity is modified to account for the transport of
the turbulent shear stress.

• The modeling constants are different.

These features make the SST k-ω model more accurate and reliable for a wider class
of flows (e.g., adverse pressure gradient flows, airfoils, transonic shock waves) than the
standard k-ω model.

10.2.10 The v 2 -f Model


The v 2 -f model is similar to the standard k- model, but incorporates near-wall tur-
bulence anisotropy and non-local pressure-strain effects. It 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 [65], it also accurately simulates flows
dominated by separation [17].

10-8
c Fluent Inc. January 28, 2003
10.2 Choosing a Turbulence Model

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 fluctuation 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).

10.2.11 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
five additional transport equations are required in 2D flows and seven additional trans-
port equations must be solved in 3D.
Since the RSM accounts for the effects 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 flows. However,
the fidelity of RSM predictions is still limited by the closure assumptions employed to
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 flows to warrant the additional computational expense. However, use
of the RSM is a must when the flow features of interest are the result of anisotropy in
the Reynolds stresses. Among the examples are cyclone flows, highly swirling flows in
combustors, rotating flow passages, and the stress-induced secondary flows in ducts.

10.2.12 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 effort than the Spalart-
Allmaras model since an additional transport equation is solved. The realizable k-
model requires only slightly more computational effort 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 10–15% more
CPU time than with the standard k- model. Like the k- models, the k-ω models are
also two-equation models, and thus require about the same computational effort.


c Fluent Inc. January 28, 2003 10-9
Modeling Turbulence

Compared with the k- and k-ω models, the RSM requires additional memory and CPU
time due to the increased number of the transport equations for Reynolds stresses. How-
ever, efficient programming in FLUENT has reduced the CPU time per iteration signifi-
cantly. On average, the RSM in FLUENT requires 50–60% more CPU time per iteration
compared to the k- and k-ω models. Furthermore, 15–20% more memory is needed.
Aside from the time per iteration, the choice of turbulence model can affect the ability of
FLUENT to obtain a converged solution. For example, the standard k- model is known
to be slightly over-diffusive 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
diffusion has a stabilizing effect 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-ω models
due to the strong coupling between the Reynolds stresses and the mean flow.

10.3 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 [251] solves
a transport equation for a quantity that is a modified form of the turbulent kinematic
viscosity.

10.3.1 Transport Equation for the Spalart-Allmaras Model


The transported variable in the Spalart-Allmaras model, ν̃, is identical to the turbu-
lent kinematic viscosity except in the near-wall (viscous-affected) region. The transport
equation for ν̃ is

 ( ) !2 
∂ ∂ 1 ∂ ∂ ν̃ ∂ ν̃  −Yν +Sν̃ (10.3-1)
(ρν̃)+ (ρν̃ui ) = Gν +  (µ + ρν̃) + Cb2 ρ
∂t ∂xi σν̃ ∂xj ∂xj ∂xj

where Gν is the production of turbulent viscosity and Yν is the destruction of turbulent


viscosity that occurs in the near-wall region due to wall blocking and viscous damping.
σν̃ and Cb2 are constants and ν is the molecular kinematic viscosity. Sν̃ is a user-defined
source term. Note that since the turbulence kinetic energy k is not calculated in the
Spalart-Allmaras model, the last term in Equation 10.2-5 is ignored when estimating the
Reynolds stresses.

10-10
c Fluent Inc. January 28, 2003
10.3 The Spalart-Allmaras Model

10.3.2 Modeling the Turbulent Viscosity


The turbulent viscosity, µt , is computed from

µt = ρν̃fv1 (10.3-2)

where the viscous damping function, fv1 , is given by

χ3
fv1 = 3
(10.3-3)
χ3 + Cv1

and

ν̃
χ≡ (10.3-4)
ν

10.3.3 Modeling the Turbulent Production


The production term, Gν , is modeled as

Gν = Cb1 ρS̃ ν̃ (10.3-5)

where

ν̃
S̃ ≡ S + fv2 (10.3-6)
κ2 d2
and

χ
fv2 = 1 − (10.3-7)
1 + χfv1

Cb1 and κ are constants, d is the distance from the wall, and S is a scalar measure of the
deformation tensor. By default in FLUENT, as in the original model proposed by Spalart
and Allmaras, S is based on the magnitude of the vorticity:

q
S≡ 2Ωij Ωij (10.3-8)

where Ωij is the mean rate-of-rotation tensor and is defined by


!
1 ∂ui ∂uj
Ωij = − (10.3-9)
2 ∂xj ∂xi


c Fluent Inc. January 28, 2003 10-11
Modeling Turbulence

The justification for the default expression for S is that, for the wall-bounded flows 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 effect of mean strain on the turbulence production, and
a modification to the model has been proposed [49] and incorporated into FLUENT.
This modification combines measures of both rotation and strain tensors in the definition
of S:

S ≡ |Ωij | + Cprod min (0, |Sij | − |Ωij |) (10.3-10)

where

q q
Cprod = 2.0, |Ωij | ≡ 2Ωij Ωij , |Sij | ≡ 2Sij Sij

with the mean strain rate, Sij , defined as


!
1 ∂uj ∂ui
Sij = + (10.3-11)
2 ∂xi ∂xj

Including both the rotation and strain tensors reduces the production of eddy viscosity
and consequently reduces the eddy viscosity itself in regions where the measure of vortic-
ity exceeds that of strain rate. One such example can be found in vortical flows, i.e., flow
near the core of a vortex subjected to a pure rotation where turbulence is known to be
suppressed. Including both the rotation and strain tensors more correctly accounts for
the effects of rotation on turbulence. The default option (including the rotation tensor
only) tends to overpredict the production of eddy viscosity and hence overpredicts the
eddy viscosity itself in certain circumstances.
You can select the modified form for calculating production in the Viscous Model panel.

10.3.4 Modeling the Turbulent Destruction


The destruction term is modeled as

 2
ν̃
Yν = Cw1 ρfw (10.3-12)
d
where

#1/6
6
"
1 + Cw3
fw = g 6 6
(10.3-13)
g + Cw3

10-12
c Fluent Inc. January 28, 2003
10.3 The Spalart-Allmaras Model

 
g = r + Cw2 r6 − r (10.3-14)

ν̃
r≡ (10.3-15)
S̃κ2 d2

Cw1 , Cw2 , and Cw3 are constants, and S̃ is given by Equation 10.3-6. Note that the
modification described above to include the effects of mean strain on S will also affect
the value of S̃ used to compute r.

10.3.5 The DES Model


The standard Spalart-Allmaras model uses the distance to the closest wall as the defini-
tion for the length scale d, which plays a major role in determining the level of production
and destruction of turbulent viscosity (Equations 10.3-6, 10.3-12, and 10.3-15). The DES
model, as proposed by Shur et al. [233] replaces d everywhere with a new length scale d, ˜
defined as

d˜ = min(d, Cdes ∆) (10.3-16)

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 Cdes has a value of 0.65.

10.3.6 Model Constants


The model constants Cb1 , Cb2 , σν̃ , Cv1 , Cw1 , Cw2 , Cw3 , and κ have the following default
values [251]:

2
Cb1 = 0.1355, Cb2 = 0.622, σν̃ = , Cv1 = 7.1
3

Cb1 (1 + Cb2 )
Cw1 = + , Cw2 = 0.3, Cw3 = 2.0, κ = 0.4187
κ2 σν̃

10.3.7 Wall Boundary Conditions


At walls, the modified turbulent kinematic viscosity, ν̃, is set to zero.
When the mesh is fine enough to resolve the laminar sublayer, the wall shear stress is
obtained from the laminar stress-strain relationship:

u ρuτ y
= (10.3-17)
uτ µ


c Fluent Inc. January 28, 2003 10-13
Modeling Turbulence

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 1 ρuτ y
= ln E (10.3-18)
uτ κ µ

where u is the velocity parallel to the wall, uτ is the shear velocity, y is the distance from
the wall, κ is the von Kármán constant (0.4187), and E = 9.793.

10.3.8 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:
" #
∂ ∂ ∂ c p µt ∂T

(ρE) + [ui (ρE + p)] = k+ + ui (τij )eff + Sh (10.3-19)
∂t ∂xi ∂xj Prt ∂xj

where k, in this case, is the thermal conductivity, E is the total energy, and (τij )eff is the
deviatoric stress tensor, defined as
!
∂uj ∂ui 2 ∂ui
(τij )eff = µeff + − µeff δij
∂xi ∂xj 3 ∂xi

The term involving (τij )eff 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 Prt 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”.

10-14
c Fluent Inc. January 28, 2003
10.4 The Standard, RNG, and Realizable k- Models

10.4 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 differences in the
models are as follows:

• the method of calculating turbulent viscosity


• the turbulent Prandtl numbers governing the turbulent diffusion 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 effects of compressibility, and modeling heat and mass transfer.

10.4.1 The Standard k- Model


The standard k- model [140] 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 flow is fully turbulent, and
the effects of molecular viscosity are negligible. The standard k- model is therefore valid
only for fully turbulent flows.

Transport Equations for the Standard k- Model


The turbulence kinetic energy, k, and its rate of dissipation, , are obtained from the
following transport equations:
" #
∂ ∂ ∂ µt ∂k

(ρk) + (ρkui ) = µ+ + Gk + Gb − ρ − YM + Sk (10.4-1)
∂t ∂xi ∂xj σk ∂xj

and

2
" #
∂ ∂ ∂ µt ∂ 

(ρ) + (ρui ) = µ+ + C1 (Gk + C3 Gb ) − C2 ρ + S (10.4-2)
∂t ∂xi ∂xj σ ∂xj k k

In these equations, Gk represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 10.4.4. Gb is the generation


c Fluent Inc. January 28, 2003 10-15
Modeling Turbulence

of turbulence kinetic energy due to buoyancy, calculated as described in Section 10.4.5.


YM represents the contribution of the fluctuating dilatation in compressible turbulence
to the overall dissipation rate, calculated as described in Section 10.4.6. C1 , C2 , and
C3 are constants. σk and σ are the turbulent Prandtl numbers for k and , respectively.
Sk and S are user-defined source terms.

Modeling the Turbulent Viscosity


The turbulent (or eddy) viscosity, µt , is computed by combining k and  as follows:

k2
µt = ρCµ (10.4-3)

where Cµ is a constant.

Model Constants
The model constants C1 , C2 , Cµ , σk , and σ have the following default values [140]:

C1 = 1.44, C2 = 1.92, Cµ = 0.09, σk = 1.0, σ = 1.3

These default values have been determined from experiments with air and water for funda-
mental turbulent shear flows including homogeneous shear flows and decaying isotropic
grid turbulence. They have been found to work fairly well for a wide range of wall-
bounded and free shear flows.
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.

10.4.2 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 different from those in
the standard k- model, and additional terms and functions in the transport equations
for k and . A more comprehensive description of RNG theory and its application to
turbulence can be found in [40].

Transport Equations for the RNG k- Model


The RNG k- model has a similar form to the standard k- model:
!
∂ ∂ ∂ ∂k
(ρk) + (ρkui ) = αk µeff + Gk + Gb − ρ − YM + Sk (10.4-4)
∂t ∂xi ∂xj ∂xj

10-16
c Fluent Inc. January 28, 2003
10.4 The Standard, RNG, and Realizable k- Models

and

2
!
∂ ∂ ∂ ∂ 
(ρ) + (ρui ) = α µeff + C1 (Gk + C3 Gb ) − C2 ρ − R + S (10.4-5)
∂t ∂xi ∂xj ∂xj k k

In these equations, Gk represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 10.4.4. Gb is the generation
of turbulence kinetic energy due to buoyancy, calculated as described in Section 10.4.5.
YM represents the contribution of the fluctuating dilatation in compressible turbulence
to the overall dissipation rate, calculated as described in Section 10.4.6. The quantities
αk and α are the inverse effective Prandtl numbers for k and , respectively. Sk and S
are user-defined source terms.

Modeling the Effective Viscosity


The scale elimination procedure in RNG theory results in a differential equation for
turbulent viscosity:

ρ2 k
!
ν̂
d √ = 1.72 √ dν̂ (10.4-6)
µ ν̂ 3 − 1 + Cν

where

ν̂ = µeff /µ
Cν ≈ 100

Equation 10.4-6 is integrated to obtain an accurate description of how the effective tur-
bulent transport varies with the effective Reynolds number (or eddy scale), allowing the
model to better handle low-Reynolds-number and near-wall flows.
In the high-Reynolds-number limit, Equation 10.4-6 gives

k2
µt = ρCµ (10.4-7)

with Cµ = 0.0845, derived using RNG theory. It is interesting to note that this value
of Cµ is very close to the empirically-determined value of 0.09 used in the standard k-
model.
In FLUENT, by default, the effective viscosity is computed using the high-Reynolds-
number form in Equation 10.4-7. However, there is an option available that allows you
to use the differential relation given in Equation 10.4-6 when you need to include low-
Reynolds-number effects.


c Fluent Inc. January 28, 2003 10-17
Modeling Turbulence

RNG Swirl Modification


Turbulence, in general, is affected by rotation or swirl in the mean flow. The RNG model
in FLUENT provides an option to account for the effects of swirl or rotation by modifying
the turbulent viscosity appropriately. The modification takes the following functional
form:
!
k
µt = µt0 f αs , Ω, (10.4-8)


where µt0 is the value of turbulent viscosity calculated without the swirl modification
using either Equation 10.4-6 or Equation 10.4-7. Ω is a characteristic swirl number eval-
uated within FLUENT, and αs is a swirl constant that assumes different values depending
on whether the flow is swirl-dominated or only mildly swirling. This swirl modification
always takes effect for axisymmetric, swirling flows and three-dimensional flows when the
RNG model is selected. For mildly swirling flows (the default in FLUENT), αs is set to
0.05 and cannot be modified. For strongly swirling flows, however, a higher value of αs
can be used.

Calculating the Inverse Effective Prandtl Numbers


The inverse effective Prandtl numbers, αk and α , are computed using the following
formula derived analytically by the RNG theory:

0.6321 0.3679
α − 1.3929 α + 2.3929 µmol


= (10.4-9)

α0 − 1.3929
α0 + 2.3929

µeff

where α0 = 1.0. In the high-Reynolds-number limit (µmol /µeff  1), αk = α ≈ 1.393.

The R Term in the  Equation


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

Cµ ρη 3 (1 − η/η0 ) 2
R = (10.4-10)
1 + βη 3 k

where η ≡ Sk/, η0 = 4.38, β = 0.012.


The effects of this term in the RNG  equation can be seen more clearly by rearranging
Equation 10.4-5. Using Equation 10.4-10, the third and fourth terms on the right-hand
side of Equation 10.4-5 can be merged, and the resulting  equation can be rewritten as

10-18
c Fluent Inc. January 28, 2003
10.4 The Standard, RNG, and Realizable k- Models

2
!
∂ ∂ ∂ ∂  ∗ 
(ρ) + (ρui ) = α µeff + C1 (Gk + C3 Gb ) − C2 ρ (10.4-11)
∂t ∂xi ∂xj ∂xj k k


where C2 is given by

∗ Cµ ρη 3 (1 − η/η0 )
C2 ≡ C2 + (10.4-12)
1 + βη 3

In regions where η < η0 , the R term makes a positive contribution, and C2 becomes
larger than C2 . In the logarithmic layer, for instance, it can be shown that η ≈ 3.0,

giving C2 ≈ 2.0, which is close in magnitude to the value of C2 in the standard k-
model (1.92). As a result, for weakly to moderately strained flows, the RNG model tends
to give results largely comparable to the standard k- model.
In regions of large strain rate (η > η0 ), however, the R term makes a negative contribu-

tion, making the value of C2 less than C2 . In comparison with the standard k- model,
the smaller destruction of  augments , reducing k and, eventually, the effective viscosity.
As a result, in rapidly strained flows, the RNG model yields a lower turbulent viscosity
than the standard k- model.
Thus, the RNG model is more responsive to the effects of rapid strain and streamline
curvature than the standard k- model, which explains the superior performance of the
RNG model for certain classes of flows.

Model Constants
The model constants C1 and C2 in Equation 10.4-5 have values derived analytically by
the RNG theory. These values, used by default in FLUENT, are

C1 = 1.42, C2 = 1.68

10.4.3 The Realizable k- Model


In addition to the standard and RNG-based k- models described in Sections 10.4.1 and
10.4.2, FLUENT also provides the so-called realizable k- model [232]. The term “real-
izable” means that the model satisfies certain mathematical constraints on the normal
stresses, consistent with the physics of turbulent flows. To understand this, consider
combining the Boussinesq relationship (Equation 10.2-5) and the eddy viscosity defini-
tion (Equation 10.4-3) to obtain the following expression for the normal Reynolds stress
in an incompressible strained mean flow:

2 ∂U
u2 = k − 2 νt (10.4-13)
3 ∂x


c Fluent Inc. January 28, 2003 10-19
Modeling Turbulence

Using Equation 10.4-3 for νt ≡ µt /ρ, one obtains the result that the normal stress, u2 ,
which by definition is a positive quantity, becomes negative, i.e., “non-realizable”, when
the strain is large enough to satisfy

k ∂U 1
> ≈ 3.7 (10.4-14)
 ∂x 3Cµ

Similarly, it can also be shown that the Schwarz inequality for shear stresses (uα uβ 2 ≤
u2α u2β ; no summation over α and β) can be violated when the mean strain rate is large.
The most straightforward way to ensure the realizability (positivity of normal stresses
and Schwarz inequality for shear stresses) is to make Cµ variable by sensitizing it to
the mean flow (mean deformation) and the turbulence (k, ). The notion of variable
Cµ is suggested by many modelers including Reynolds [212], and is well substantiated
by experimental evidence. For example, Cµ is found to be around 0.09 in the inertial
sublayer of equilibrium boundary layers, and 0.05 in a strong homogeneous shear flow.
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 finding that the spreading rate in planar jets is predicted reasonably well, but
prediction of the spreading rate for axisymmetric jets is unexpectedly poor) is considered
to be mainly due to the modeled dissipation equation.
The realizable k- model proposed by Shih et al. [232] was intended to address these
deficiencies of traditional k- models by adopting the following:

• a new eddy-viscosity formula involving a variable Cµ originally proposed by


Reynolds [212]
• a new model equation for dissipation () based on the dynamic equation of the
mean-square vorticity fluctuation

Transport Equations for the Realizable k- Model


The modeled transport equations for k and  in the realizable k- model are

" #
∂ ∂ ∂ µt ∂k

(ρk) + (ρkuj ) = µ+ + Gk + Gb − ρ − YM + Sk (10.4-15)
∂t ∂xi ∂xi σk ∂xj

and

2
" #
∂ ∂ ∂ µt ∂ 

(ρ) + (ρuj ) = µ+ + ρ C1 S − ρ C2 √ + C1 C3 Gb + S
∂t ∂xj ∂xj σ ∂xj k + ν k
(10.4-16)

10-20
c Fluent Inc. January 28, 2003
10.4 The Standard, RNG, and Realizable k- Models

where
" #
η k
C1 = max 0.43, , η=S
η+5 

In these equations, Gk represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 10.4.4. Gb is the generation
of turbulence kinetic energy due to buoyancy, calculated as described in Section 10.4.5.
YM represents the contribution of the fluctuating dilatation in compressible turbulence
to the overall dissipation rate, calculated as described in Section 10.4.6. C2 and C1 are
constants. σk and σ are the turbulent Prandtl numbers for k and , respectively. Sk and
S are user-defined source terms.
Note that the k equation (Equation 10.4-15) is the same as that in the standard k-
 model (Equation 10.4-1) and the RNG k- model (Equation 10.4-4), except for the
model constants. However, the form of the  equation is quite different from those in
the standard and RNG-based k- models (Equations 10.4-2 and 10.4-5). One of the
noteworthy features is that the production term in the  equation (the second term on
the right-hand side of Equation 10.4-16) does not involve the production of k; i.e., it does
not contain the same Gk term as the other k- models. It is believed that the present
form better represents the spectral energy transfer. Another desirable feature is that
the destruction term (the next to last term on the right-hand side of Equation 10.4-16)
does not have any singularity; i.e., its denominator never vanishes, even if k vanishes or
becomes smaller than zero. This feature is contrasted with traditional k- models, which
have a singularity due to k in the denominator.
This model has been extensively validated for a wide range of flows [129, 232], including
rotating homogeneous shear flows, free flows including jets and mixing layers, channel
and boundary layer flows, and separated flows. For all these cases, the performance of
the model has been found to be substantially better than that of the standard k- model.
Especially noteworthy is the fact that the realizable k- model resolves the round-jet
anomaly; i.e., it predicts the spreading rate for axisymmetric jets as well as that for
planar jets.

Modeling the Turbulent Viscosity


As in other k- models, the eddy viscosity is computed from

k2
µt = ρCµ (10.4-17)

The difference between the realizable k- model and the standard and RNG k- models
is that Cµ is no longer constant. It is computed from


c Fluent Inc. January 28, 2003 10-21
Modeling Turbulence

1
Cµ = ∗ (10.4-18)
A0 + As kU

where
q

U ≡ Sij Sij + Ω̃ij Ω̃ij (10.4-19)

and

Ω̃ij = Ωij − 2ijk ωk


Ωij = Ωij − ijk ωk

where Ωij is the mean rate-of-rotation tensor viewed in a rotating reference frame with
the angular velocity ωk . The model constants A0 and As are given by


A0 = 4.04, As = 6 cos φ

where


!
1 Sij Sjk Ski q 1 ∂uj ∂ui
φ = cos−1 ( 6W ), W = , S̃ = Sij Sij , Sij = +
3 S̃ 2 ∂xi ∂xj

It can be seen that Cµ is a function of the mean strain and rotation rates, the angular ve-
locity of the system rotation, and the turbulence fields (k and ). Cµ in Equation 10.4-17
can be shown to recover the standard value of 0.09 for an inertial sublayer in an equilib-
rium boundary layer.
! In FLUENT, the term −2ijk ωk is, by default, not included in the calculation of Ω̃ij . This
is an extra rotation term that is not compatible 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 C2 , σk , and σ have been established to ensure that the model
performs well for certain canonical flows. The model constants are

C1 = 1.44, C2 = 1.9, σk = 1.0, σ = 1.2

10-22
c Fluent Inc. January 28, 2003
10.4 The Standard, RNG, and Realizable k- Models

10.4.4 Modeling Turbulent Production in the k- Models


The term Gk , 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 defined as

∂uj
Gk = −ρu0i u0j (10.4-20)
∂xi

To evaluate Gk in a manner consistent with the Boussinesq hypothesis,

G k = µt S 2 (10.4-21)

where S is the modulus of the mean rate-of-strain tensor, defined as

q
S≡ 2Sij Sij (10.4-22)

10.4.5 Effects of Buoyancy on Turbulence in the k- Models


When a non-zero gravity field and temperature gradient are present simultaneously, the
k- models in FLUENT account for the generation of k due to buoyancy (Gb in Equa-
tions 10.4-1, 10.4-4, and 10.4-15), and the corresponding contribution to the production
of  in Equations 10.4-2, 10.4-5, and 10.4-16.
The generation of turbulence due to buoyancy is given by

µt ∂T
Gb = βgi (10.4-23)
Prt ∂xi

where Prt is the turbulent Prandtl number for energy and gi is the component of the
gravitational vector in the ith direction. For the standard and realizable k- models, the
default value of Prt is 0.85. In the case of the RNG k- model, Prt = 1/α, where α
is given by Equation 10.4-9, but with α0 = 1/Pr = k/µcp . The coefficient of thermal
expansion, β, is defined as
!
1 ∂ρ
β=− (10.4-24)
ρ ∂T p

For ideal gases, Equation 10.4-23 reduces to

µt ∂ρ
Gb = −gi (10.4-25)
ρPrt ∂xi


c Fluent Inc. January 28, 2003 10-23
Modeling Turbulence

It can be seen from the transport equations for k (Equations 10.4-1, 10.4-4, and 10.4-15)
that turbulence kinetic energy tends to be augmented (Gb > 0) in unstable stratification.
For stable stratification, buoyancy tends to suppress the turbulence (Gb < 0). In FLU-
ENT, the effects of buoyancy on the generation of k are always included when you have
both a non-zero gravity field and a non-zero temperature (or density) gradient.
While the buoyancy effects on the generation of k are relatively well understood, the
effect on  is less clear. In FLUENT, by default, the buoyancy effects on  are neglected
simply by setting Gb to zero in the transport equation for  (Equation 10.4-2, 10.4-5, or
10.4-16).
However, you can include the buoyancy effects on  in the Viscous Model panel. In this
case, the value of Gb given by Equation 10.4-25 is used in the transport equation for 
(Equation 10.4-2, 10.4-5, or 10.4-16).
The degree to which  is affected by the buoyancy is determined by the constant C3 .
In FLUENT, C3 is not specified, but is instead calculated according to the following
relation [101]:

v

C3 = tanh (10.4-26)
u
where v is the component of the flow velocity parallel to the gravitational vector and
u is the component of the flow velocity perpendicular to the gravitational vector. In
this way, C3 will become 1 for buoyant shear layers for which the main flow direction is
aligned with the direction of gravity. For buoyant shear layers that are perpendicular to
the gravitational vector, C3 will become zero.

10.4.6 Effects of Compressibility on Turbulence in the k- Models


For high-Mach-number flows, compressibility affects turbulence through so-called “di-
latation dissipation”, which is normally neglected in the modeling of incompressible
flows [294]. 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 effects in the k- models in FLUENT, the dilatation
dissipation term, YM , is included in the k equation. This term is modeled according to
a proposal by Sarkar [220]:

YM = 2ρM2t (10.4-27)

where Mt is the turbulent Mach number, defined as


s
k
Mt = (10.4-28)
a2

10-24
c Fluent Inc. January 28, 2003
10.4 The Standard, RNG, and Realizable k- Models


where a (≡ γRT ) is the speed of sound.
This compressibility modification always takes effect when the compressible form of the
ideal gas law is used.

10.4.7 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
(ρE) + [ui (ρE + p)] = keff + ui (τij )eff + Sh (10.4-29)
∂t ∂xi ∂xj ∂xj

where E is the total energy, keff is the effective thermal conductivity, and
(τij )eff is the deviatoric stress tensor, defined as
!
∂uj ∂ui 2 ∂ui
(τij )eff = µeff + − µeff δij
∂xi ∂xj 3 ∂xi

The term involving (τij )eff 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 11.2.1 for more details.
For the standard and realizable k- models, the effective thermal conductivity is given
by

c p µt
keff = k +
Prt

where k, in this case, is the thermal conductivity. The default value of the turbulent
Prandtl number is 0.85. You can change the value of the turbulent Prandtl number in
the Viscous Model panel.
For the RNG k- model, the effective thermal conductivity is

keff = αcp µeff

where α is calculated from Equation 10.4-9, but with α0 = 1/Pr = k/µcp .


c Fluent Inc. January 28, 2003 10-25
Modeling Turbulence

The fact that α varies with µmol /µeff , as in Equation 10.4-9, 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 [124]. Equation 10.4-9
works well across a very broad range of molecular Prandtl numbers, from liquid metals
(Pr ≈ 10−2 ) to paraffin oils (Pr ≈ 103 ), which allows heat transfer to be calculated in low-
Reynolds-number regions. Equation 10.4-9 smoothly predicts the variation of effective
Prandtl number from the molecular value (α = 1/Pr) in the viscosity-dominated region
to the fully turbulent value (α = 1.393) in the fully turbulent regions of the flow.
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 effective turbulent diffusivity for mass
transfer is calculated in a manner that is analogous to the method used for the heat
transport. The value of α0 in Equation 10.4-9 is α0 = 1/Sc, where Sc is the molecular
Schmidt number.

10.5 The Standard and SST k-ω Models


This section presents the standard and shear-stress transport (SST) k-ω models. Both
models have similar forms, with transport equations for k and ω. The major ways in
which the SST model differs from the standard model are as follows:

• gradual change from the standard k-ω model in the inner region of the boundary
layer to a high-Reynolds-number version of the k- model in the outer part of the
boundary layer

• modified turbulent viscosity formulation to account for the transport effects of the
principal turbulent shear stress

The transport equations, methods of calculating turbulent viscosity, and methods of


calculating model constants and other terms are presented separately for each model.

10.5.1 The Standard k-ω Model


The standard k-ω model is an empirical model based on model transport equations for
the turbulence kinetic energy (k) and the specific dissipation rate (ω), which can also be
thought of as the ratio of  to k [294].
As the k-ω model has been modified over the years, production terms have been added
to both the k and ω equations, which have improved the accuracy of the model for
predicting free shear flows.

10-26
c Fluent Inc. January 28, 2003
10.5 The Standard and SST k-ω Models

Transport Equations for the Standard k-ω Model


The turbulence kinetic energy, k, and the specific dissipation rate, ω, are obtained from
the following transport equations:
!
∂ ∂ ∂ ∂k
(ρk) + (ρkui ) = Γk + G k − Yk + S k (10.5-1)
∂t ∂xi ∂xj ∂xj

and
!
∂ ∂ ∂ ∂ω
(ρω) + (ρωui ) = Γω + G ω − Yω + S ω (10.5-2)
∂t ∂xi ∂xj ∂xj

In these equations, Gk represents the generation of turbulence kinetic energy due to mean
velocity gradients. Gω represents the generation of ω. Γk and Γω represent the effective
diffusivity of k and ω, respectively. Yk and Yω represent the dissipation of k and ω due
to turbulence. All of the above terms are calculated as described below. Sk and Sω are
user-defined source terms.

Modeling the Effective Diffusivity


The effective diffusivities for the k-ω model are given by

µt
Γk = µ + (10.5-3)
σk
µt
Γω = µ+ (10.5-4)
σω

where σk and σω are the turbulent Prandtl numbers for k and ω, respectively. The
turbulent viscosity, µt , is computed by combining k and ω as follows:

ρk
µt = α ∗ (10.5-5)
ω

Low-Reynolds-Number Correction

The coefficient α∗ damps the turbulent viscosity causing a low-Reynolds-number correc-


tion. It is given by

α0∗ + Ret /Rk


!
∗ ∗
α = α∞ (10.5-6)
1 + Ret /Rk

where


c Fluent Inc. January 28, 2003 10-27
Modeling Turbulence

ρk
Ret = (10.5-7)
µω
Rk = 6 (10.5-8)
βi
α0∗ = (10.5-9)
3
βi = 0.072 (10.5-10)

Note that, in the high-Reynolds-number form of the k-ω model, α∗ = α∞



= 1.

Modeling the Turbulence Production


Production of k

The term Gk represents the production of turbulence kinetic energy. From the exact
equation for the transport of k, this term may be defined as

∂uj
Gk = −ρu0i u0j (10.5-11)
∂xi

To evaluate Gk in a manner consistent with the Boussinesq hypothesis,

G k = µt S 2 (10.5-12)

where S is the modulus of the mean rate-of-strain tensor, defined in the same way as for
the k- model (see Equation 10.4-22).

Production of ω

The production of ω is given by

ω
Gω = α Gk (10.5-13)
k
where Gk is given by Equation 10.5-11.
The coefficient α is given by
!
α∞ α0 + Ret /Rω
α= ∗ (10.5-14)
α 1 + Ret /Rω

where Rω = 2.95. α∗ and Ret are given by Equations 10.5-6 and 10.5-7, respectively.
Note that, in the high-Reynolds-number form of the k-ω model, α = α∞ = 1.

10-28
c Fluent Inc. January 28, 2003
10.5 The Standard and SST k-ω Models

Modeling the Turbulence Dissipation


Dissipation of k

The dissipation of k is given by

Yk = ρ β ∗ fβ ∗ k ω (10.5-15)

where

 1 χk ≤ 0
fβ ∗ =  1+680χ2k (10.5-16)
1+400χ2k
χk > 0

where

1 ∂k ∂ω
χk ≡ (10.5-17)
ω 3 ∂xj ∂xj

and

β ∗ = βi∗ [1 + ζ ∗ F (Mt )] (10.5-18)


4
!
4/15 + (Re t /Rβ )
βi∗ = β∞∗
(10.5-19)
1 + (Ret /Rβ )4
ζ ∗ = 1.5 (10.5-20)
Rβ = 8 (10.5-21)

β∞ = 0.09 (10.5-22)

where Ret is given by Equation 10.5-7.

Dissipation of ω

The dissipation of ω is given by

Yω = ρ β fβ ω 2 (10.5-23)

where

1 + 70χω
fβ = (10.5-24)
1 + 80χω


c Fluent Inc. January 28, 2003 10-29
Modeling Turbulence


Ω Ω S
ij jk ki
χω = (10.5-25)
(β ∗ ω)3


!
1 ∂ui ∂uj
Ωij = − (10.5-26)
2 ∂xj ∂xi

The strain rate tensor, Sij is defined in Equation 10.3-11. Also,

β∗
" #
β = βi 1 − i ζ ∗ F (Mt ) (10.5-27)
βi

βi∗ and F (Mt ) are defined by Equations 10.5-19 and 10.5-28, respectively.

Compressibility Correction

The compressibility function, F (Mt ), is given by


(
0 Mt ≤ Mt0
F (Mt ) = (10.5-28)
M2t − M2t0 Mt > Mt0

where

2k
M2t ≡ (10.5-29)
a2
Mt0 = 0.25 (10.5-30)
q
a = γRT (10.5-31)

Note that, in the high-Reynolds-number form of the k-ω model, βi∗ = β∞



. In the incom-
∗ ∗
pressible form, β = βi .

Model Constants
∗ 1 ∗
α∞ = 1, α∞ = 0.52, α0 = , β∞ = 0.09, βi = 0.072, Rβ = 8
9
Rk = 6, Rω = 2.95, ζ ∗ = 1.5, Mt0 = 0.25, σk = 2.0, σω = 2.0

Wall Boundary Conditions


The wall boundary conditions for the k equation in the k-ω 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 fine meshes, the appropriate
low-Reynolds-number boundary conditions will be applied.

10-30
c Fluent Inc. January 28, 2003
10.5 The Standard and SST k-ω Models

In FLUENT the value of ω at the wall is specified as

ρ (u∗ )2 +
ωw = ω (10.5-32)
µ

The asymptotic value of ω + in the laminar sublayer is given by


!
+ 6
ω = min ωw+ , ∗ (y + )2
(10.5-33)
β∞

where

  2
50

 ks+
 ks+ < 25
ωw+ =  (10.5-34)
100
ks+ ≥ 25
 
ks+

where

ρks u∗
!
ks+ = max 1.0, (10.5-35)
µ

and ks is the roughness height.


In the logarithmic (or turbulent) region, the value of ω + is

1 du+turb
ω+ = q +
(10.5-36)

β∞ dy

which leads to the value of ω in the wall cell as

u∗
ω=q (10.5-37)
β∞∗ κy

Note that in the case of a wall cell being placed in the buffer region, FLUENT will blend
ω + between the logarithmic and laminar sublayer values.

10.5.2 The Shear-Stress Transport (SST) k-ω Model


In addition to the standard k-ω model described in Section 10.5.1, FLUENT also provides
a variation called the shear-stress transport (SST) k-ω model, so named because the def-
inition of the turbulent viscosity is modified to account for the transport of the principal
turbulent shear stress. It is this feature that gives the SST k-ω model an advantage in


c Fluent Inc. January 28, 2003 10-31
Modeling Turbulence

terms of performance over both the standard k-ω model and the standard k- model.
Other modifications include the addition of a cross-diffusion term in the ω equation and
a blending function to ensure that the model equations behave appropriately in both the
near-wall and far-field zones.

Transport Equations for the SST k-ω Model


The SST k-ω model has a similar form to the standard k-ω model:
!
∂ ∂ ∂ ∂k
(ρk) + (ρkui ) = Γk + G k − Yk + S k (10.5-38)
∂t ∂xi ∂xj ∂xj

and
!
∂ ∂ ∂ ∂ω
(ρω) + (ρωui ) = Γω + G ω − Y ω + Dω + S ω (10.5-39)
∂t ∂xi ∂xj ∂xj

In these equations, Gk represents the generation of turbulence kinetic energy due to mean
velocity gradients, calculated as described in Section 10.5.1. Gω represents the generation
of ω, calculated as described in Section 10.5.1. Γk and Γω represent the effective diffusivity
of k and ω, respectively, which are calculated as described below. Yk and Yω represent
the dissipation of k and ω due to turbulence, calculated as described in Section 10.5.1.
Dω represents the cross-diffusion term, calculated as described below. Sk and Sω are
user-defined source terms.

Modeling the Effective Diffusivity


The effective diffusivities for the SST k-ω model are given by

µt
Γk = µ + (10.5-40)
σk
µt
Γω = µ+ (10.5-41)
σω

where σk and σω are the turbulent Prandtl numbers for k and ω, respectively. The
turbulent viscosity, µt , is computed as follows:

ρk 1
µt = h i (10.5-42)
ω max 1∗ , ΩF2
α a1 ω

where

10-32
c Fluent Inc. January 28, 2003
10.5 The Standard and SST k-ω Models

q
Ω ≡ 2Ωij Ωij (10.5-43)
1
σk = (10.5-44)
F1 /σk,1 + (1 − F1 )/σk,2
1
σω = (10.5-45)
F1 /σω,1 + (1 − F1 )/σω,2

Ωij is the mean rate-of-rotation tensor and α∗ is defined in Equation 10.5-6. The blending
functions, F1 and F2 , are given by

 
F1 = tanh Φ41 (10.5-46)
" √ ! #
k 500µ 4ρk
Φ1 = min max , 2 , (10.5-47)
0.09ωy ρy ω σω,2 Dω+ y 2
" #
1 1 ∂k ∂ω −20
Dω+ = max 2ρ , 10 (10.5-48)
σω,2 ω ∂xj ∂xj

 
F2 = tanh Φ22 (10.5-49)
" √ #
k 500µ
Φ2 = max 2 , (10.5-50)
0.09ωy ρy 2 ω

where y is the distance to the next surface and Dω+ is the positive portion of the cross-
diffusion term (see Equation 10.5-60).

Modeling the Turbulence Production


Production of k

The term Gk represents the production of turbulence kinetic energy, and is defined in
the same manner as in the standard k-ω model. See Section 10.5.1 for details.

Production of ω

The term Gω represents the production of ω and is given by

α
Gω = Gk (10.5-51)
νt
Note that this formulation differs from the standard k-ω model. The difference between
the two models also exists in the way the term α∞ is evaluated. In the standard k-ω
model, α∞ is defined as a constant (0.52). For the SST k-ω model, α∞ is given by


c Fluent Inc. January 28, 2003 10-33
Modeling Turbulence

α∞ = F1 α∞,1 + (1 − F1 )α∞,2 (10.5-52)

where

βi,1 κ2
α∞,1 = ∗
− q (10.5-53)
β∞ ∗
σw,1 β∞
βi,2 κ2
α∞,2 = ∗
− q (10.5-54)
β∞ ∗
σw,2 β∞

where κ is 0.41. βi,1 and βi,2 are given by Equations 10.5-58 and 10.5-59, respectively.

Modeling the Turbulence Dissipation


Dissipation of k

The term Yk represents the dissipation of turbulence kinetic energy, and is defined in a
similar manner as in the standard k-ω model (see Section 10.5.1). The difference is in the
way the term fβ ∗ is evaluated. In the standard k-ω model, fβ ∗ is defined as a piecewise
function. For the SST k-ω model, fβ ∗ is a constant equal to 1. Thus,

Yk = ρβ ∗ kω (10.5-55)

Dissipation of ω

The term Yω represents the dissipation of ω, and is defined in a similar manner as in the
standard k-ω model (see Section 10.5.1). The difference is in the way the terms βi and
fβ are evaluated. In the standard k-ω model, βi is defined as a constant (0.072) and fβ is
defined in Equation 10.5-24. For the SST k-ω model, fβ is a constant equal to 1. Thus,

Yk = ρβω 2 (10.5-56)

Instead of a having a constant value, βi is given by

βi = F1 βi,1 + (1 − F1 )βi,2 (10.5-57)

where

βi,1 = 0.075 (10.5-58)


βi,2 = 0.0828 (10.5-59)

10-34
c Fluent Inc. January 28, 2003
10.6 The Reynolds Stress Model (RSM)

and F1 is obtained from Equation 10.5-46.

Cross-Diffusion Modification
The SST k-ω model is based on both the standard k-ω model and the standard k- model.
To blend these two models together, the standard k- model has been transformed into
equations based on k and ω, which leads to the introduction of a cross-diffusion term
(Dω in Equation 10.5-39). Dω is defined as

1 ∂k ∂ω
Dω = 2 (1 − F1 ) ρσω,2 (10.5-60)
ω ∂xj ∂xj

For details about the various k- models, see Section 10.4.

Model Constants
σk,1 = 1.176, σω,1 = 2.0, σk,2 = 1.0, σω,2 = 1.168

a1 = 0.31, βi,1 = 0.075 βi,2 = 0.0828


∗ ∗
All additional model constants (α∞ , α∞ , α0 , β∞ , Rβ , Rk , Rω , ζ ∗ , and Mt0 ) have the same
values as for the standard k-ω model (see Section 10.5.1).

10.6 The Reynolds Stress Model (RSM)


The Reynolds stress model [85, 137, 138] involves calculation of the individual Reynolds
stresses, u0i u0j , using differential transport equations. The individual Reynolds stresses
are then used to obtain closure of the Reynolds-averaged momentum equation (Equa-
tion 10.2-4).
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 fluctuating 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.

10.6.1 The Reynolds Stress Transport Equations


The exact transport equations for the transport of the Reynolds stresses, ρu0i u0j , may be
written as follows:


c Fluent Inc. January 28, 2003 10-35
Modeling Turbulence

∂ ∂ ∂

  
(ρ u0i u0j ) + (ρuk u0i u0j ) = − ρ u0i u0j u0k + p δkj u0i + δik u0j
|∂t {z } ∂x
| k {z } |
∂xk {z }
Local Time Derivative Cij ≡ Convection DT,ij ≡ Turbulent Diffusion

" # !
∂ ∂ ∂uj ∂ui
+ µ (u0 u0 ) − ρ u0i u0k + u0j u0k − ρβ(gi u0j θ + gj u0i θ)
∂xk ∂xk i j ∂xk ∂xk | {z }
| {z } | {z }
Gij ≡ Buoyancy Production
DL,ij ≡ Molecular Diffusion Pij ≡ Stress Production

∂u0i ∂u0j ∂u0i ∂u0j


!
+ p + − 2µ
∂xj ∂xi ∂xk ∂xk
| {z } | {z }
φij ≡ Pressure Strain ij ≡ Dissipation
 
−2ρΩk u0j u0m ikm + u0i u0m jkm + Suser (10.6-1)
| {z } | {z }
Fij ≡ Production by System Rotation User-Defined Source Term

Of the various terms in these exact equations, Cij , DL,ij , Pij , and Fij do not require any
modeling. However, DT,ij , Gij , φij , and ij need to be modeled to close the equations.
The following sections describe the modeling assumptions required to close the equation
set.

10.6.2 Modeling Turbulent Diffusive Transport


DT,ij can be modeled by the generalized gradient-diffusion model of Daly and Harlow [51]:

ku0 u0 ∂u0 u0
!

DT,ij = Cs ρ k ` i j (10.6-2)
∂xk  ∂x`

However, this equation can result in numerical instabilities, so it has been simplified in
FLUENT to use a scalar turbulent diffusivity as follows [150]:

µt ∂u0i u0j
!

DT,ij = (10.6-3)
∂xk σk ∂xk

The turbulent viscosity, µt , is computed using Equation 10.6-27.


Lien and Leschziner [150] derived a value of σk = 0.82 by applying the generalized
gradient-diffusion model, Equation 10.6-2, to the case of a planar homogeneous shear
flow. Note that this value of σk is different from that in the standard and realizable k-
models, in which σk = 1.0.

10-36
c Fluent Inc. January 28, 2003
10.6 The Reynolds Stress Model (RSM)

10.6.3 Modeling the Pressure-Strain Term


Linear Pressure-Strain Model
By default in FLUENT, the pressure-strain term, φij , in Equation 10.6-1 is modeled
according to the proposals by Gibson and Launder [85], Fu et al. [80], and Launder [136,
137].
The classical approach to modeling φij uses the following decomposition:

φij = φij,1 + φij,2 + φij,w (10.6-4)

where φij,1 is the slow pressure-strain term, also known as the return-to-isotropy term,
φij,2 is called the rapid pressure-strain term, and φij,w is the wall-reflection term.
The slow pressure-strain term, φij,1 , is modeled as

 0 0 2
 
φij,1 ≡ −C1 ρ u u − δij k (10.6-5)
k i j 3
with C1 = 1.8.
The rapid pressure-strain term, φij,2 , is modeled as

2
 
φij,2 ≡ −C2 (Pij + Fij + Gij − Cij ) − δij (P + G − C) (10.6-6)
3

where C2 = 0.60, Pij , Fij , Gij , and Cij are defined as in Equation 10.6-1, P = 12 Pkk ,
G = 12 Gkk , and C = 12 Ckk .
The wall-reflection term, φij,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

 3 3 k 3/2
 
φij,w ≡ C10
u0k u0m nk nm δij − u0i u0k nj nk − u0j u0k ni nk
k 2 2 C` d
 3/2
3 3 k

+ C20 φkm,2 nk nm δij − φik,2 nj nk − φjk,2 ni nk
2 2 C` d
(10.6-7)

where C10 = 0.5, C20 = 0.3, nk is the xk component of the unit normal to the wall, d is
the normal distance to the wall, and C` = Cµ3/4 /κ, where Cµ = 0.09 and κ is the von
Kármán constant (= 0.4187).
φij,w is included by default in the Reynolds stress model.


c Fluent Inc. January 28, 2003 10-37
Modeling Turbulence

Low-Re Modifications to the Linear Pressure-Strain Model


When the RSM is applied to near-wall flows using the enhanced wall treatment described
in Section 10.8.3, the pressure-strain model needs to be modified. The modification used
in FLUENT specifies the values of C1 , C2 , C10 , and C20 as functions of the Reynolds stress
invariants and the turbulent Reynolds number, according to the suggestion of Launder
and Shima [139]:

q n h io
C1 = 1 + 2.58A A2 1 − exp −(0.0067Ret )2 (10.6-8)

C2 = 0.75 A (10.6-9)
2
C10 = − C1 + 1.67 (10.6-10)
3 "
2
C − 16
#
0 3 2
C2 = max ,0 (10.6-11)
C2

with the turbulent Reynolds number defined as Ret = (ρk 2 /µ). The parameter A and
tensor invariants, A2 and A3 , are defined as

9
 
A ≡ 1 − (A2 − A3 ) (10.6-12)
8
A2 ≡ aik aki (10.6-13)
A3 ≡ aik akj aji (10.6-14)

aij is the Reynolds-stress anisotropy tensor, defined as

−ρu0i u0j + 23 ρkδij


!
aij = − (10.6-15)
ρk

The modifications detailed above are employed only when the enhanced wall treatment
is selected in the Viscous Model panel.

Quadratic Pressure-Strain Model


An optional pressure-strain model proposed by Speziale, Sarkar, and Gatski [254] is
provided in FLUENT. This model has been demonstrated to give superior performance in a
range of basic shear flows, including plane strain, rotating plane shear, and axisymmetric
expansion/contraction. This improved accuracy should be beneficial for a wider class of
complex engineering flows, 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:

10-38
c Fluent Inc. January 28, 2003
10.6 The Reynolds Stress Model (RSM)

1
  q 
φij = − (C1 ρ + C1∗ P ) bij + C2 ρ bik bkj − bmn bmn δij + C3 − C3∗ bij bij ρkSij
3

2
 
+ C4 ρk bik Sjk + bjk Sik − bmn Smn δij + C5 ρk (bik Ωjk + bjk Ωik ) (10.6-16)
3
where bij is the Reynolds-stress anisotropy tensor defined as

−ρu0i u0j + 23 ρkδij


!
bij = − (10.6-17)
2ρk

The mean strain rate, Sij , is defined as


!
1 ∂uj ∂ui
Sij = + (10.6-18)
2 ∂xi ∂xj

The mean rate-of-rotation tensor, Ωij , is defined by


!
1 ∂ui ∂uj
Ωij = − (10.6-19)
2 ∂xj ∂xi

The constants are

C1 = 3.4, C1∗ = 1.8, C2 = 4.2, C3 = 0.8, C3∗ = 1.3, C4 = 1.25, C5 = 0.4

The quadratic pressure-strain model does not require a correction to account for the
wall-reflection effect 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.

10.6.4 Effects of Buoyancy on Turbulence


The production terms due to buoyancy are modeled as
!
µt ∂T ∂T
Gij = β gi + gj (10.6-20)
Prt ∂xj ∂xi

where Prt is the turbulent Prandtl number for energy, with a default value of 0.85.


c Fluent Inc. January 28, 2003 10-39
Modeling Turbulence

Using the definition of the coefficient of thermal expansion, β, given by Equation 10.4-24,
the following expression is obtained for Gij for ideal gases:
!
µt ∂ρ ∂ρ
Gij = − gi + gj (10.6-21)
ρPrt ∂xj ∂xi

10.6.5 Modeling the Turbulence Kinetic Energy


In general, when the turbulence kinetic energy is needed for modeling a specific term, it
is obtained by taking the trace of the Reynolds stress tensor:

1
k = u0i u0i (10.6-22)
2
As described in Section 10.6.8, 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:

" #
∂ ∂ ∂ µt ∂k 1

(ρk) + (ρkui ) = µ+ + (Pii + Gii ) − ρ(1 + 2M2t ) + Sk (10.6-23)
∂t ∂xi ∂xj σk ∂xj 2

where σk = 0.82 and Sk is a user-defined source term. Equation 10.6-23 is obtainable


by contracting the modeled equation for the Reynolds stresses (Equation 10.6-1). As
one might expect, it is essentially identical to Equation 10.4-1 used in the standard k-
model.
Although Equation 10.6-23 is solved globally throughout the flow domain, the values of
k obtained are used only for boundary conditions. In every other case, k is obtained from
Equation 10.6-22. This is a minor point, however, since the values of k obtained with
either method should be very similar.

10.6.6 Modeling the Dissipation Rate


The dissipation tensor, ij , is modeled as

2
ij = δij (ρ + YM ) (10.6-24)
3
where YM = 2ρM2t is an additional “dilatation dissipation” term according to the model
by Sarkar [220]. The turbulent Mach number in this term is defined as
s
k
Mt = (10.6-25)
a2

10-40
c Fluent Inc. January 28, 2003
10.6 The Reynolds Stress Model (RSM)


where a (≡ γRT ) is the speed of sound. This compressibility modification always takes
effect 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:

2
" #
∂ ∂ ∂ µt ∂ 1 

(ρ) + (ρui ) = µ+ C1 [Pii + C3 Gii ] − C2 ρ + S (10.6-26)
∂t ∂xi ∂xj σ ∂xj 2 k k

where σ = 1.0, C1 = 1.44, C2 = 1.92, C3 is evaluated as a function of the local flow
direction relative to the gravitational vector, as described in Section 10.4.5, and S is a
user-defined source term.

10.6.7 Modeling the Turbulent Viscosity


The turbulent viscosity, µt , is computed similarly to the k- models:

k2
µt = ρCµ (10.6-27)

where Cµ = 0.09.

10.6.8 Boundary Conditions for the Reynolds Stresses


Whenever flow enters the domain, FLUENT requires values for individual Reynolds
stresses, u0i u0j , 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 10.10.2.
At walls, FLUENT computes the near-wall values of the Reynolds stresses and  from wall
functions (see Section 10.8.2). FLUENT applies explicit wall boundary conditions for the
Reynolds stresses by using the log-law and the assumption of equilibrium, disregarding
convection and diffusion in the transport equations for the stresses (Equation 10.6-1).
Using a local coordinate system, where τ is the tangential coordinate, η is the normal
coordinate, and λ is the binormal coordinate, the Reynolds stresses at the wall-adjacent
cells are computed from

u0τ2 u0η2 0
uλ2 u0 u0
= 1.098, = 0.247, = 0.655, − τ η = 0.255 (10.6-28)
k k k k
To obtain k, FLUENT solves the transport equation of Equation 10.6-23. 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 field k is obtained directly from the


c Fluent Inc. January 28, 2003 10-41
Modeling Turbulence

normal Reynolds stresses using Equation 10.6-22. By default, the values of the Reynolds
stresses near the wall are fixed using the values computed from Equation 10.6-28, and
the transport equations in Equation 10.6-1 are solved only in the bulk flow region.
Alternatively, the Reynolds stresses can be explicitly specified in terms of wall-shear
stress, instead of k:

u0τ2 u0η2 0
uλ2 u0τ u0η
= 5.1, = 1.0, = 2.3, − = 1.0 (10.6-29)
u2τ u2τ u2τ u2τ
q
where uτ is the friction velocity defined by uτ ≡ τw /ρ, where τw is the wall-shear stress.
When this option is chosen, the k transport equation is not solved.

10.6.9 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:
" #
∂ ∂ ∂ c p µt ∂T

(ρE) + [ui (ρE + p)] = k+ + ui (τij )eff + Sh (10.6-30)
∂t ∂xi ∂xj Prt ∂xj

where E is the total energy and (τij )eff is the deviatoric stress tensor, defined as
!
∂uj ∂ui 2 ∂ui
(τij )eff = µeff + − µeff δij
∂xi ∂xj 3 ∂xi

The term involving (τij )eff 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 Prt 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.

10-42
c Fluent Inc. January 28, 2003
10.7 The Large Eddy Simulation (LES) Model

10.7 The Large Eddy Simulation (LES) Model


Turbulent flows 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 flow. The smallest scales are responsible for the dissipation of turbulence kinetic
energy.
It is theoretically possible to directly resolve the whole spectrum of turbulent scales
using an approach known as direct numerical simulation (DNS). DNS is not, however,
feasible for practical engineering problems. To understand the large computational cost of
DNS, consider that the ratio of the large (energy-containing) scales to the small (energy-
3/4
dissipating) scales is proportional to Ret , where Ret is the turbulent Reynolds number.
Therefore, to resolve all the scales, the mesh size in three dimensions will be proportional
9/4
to Ret . Simple arithmetic shows that, for high Reynolds numbers, the mesh sizes
required for DNS are prohibitive. Adding to the computational cost is the fact that
the simulation will be a transient one with very small time steps, since the temporal
resolution requirements are governed by the dissipating scales, rather than the mean flow
or the energy-containing eddies.
As explained in Section 10.2.1, the conventional approach to flow simulations employs
the solution of the Reynolds-averaged Navier-Stokes (RANS) equations. In the RANS
approach, all the turbulent motions are modeled, resulting in a significant savings in
computational effort.
Conceptually, large eddy simulation (LES) is situated somewhere between DNS and the
RANS approach. Basically large eddies are resolved directly in LES, while small eddies
are modeled. 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 flow involved.

• Small eddies are less dependent on the geometry, tend to be more isotropic, and
are consequently more universal.

• The chance of finding a universal model is much higher when only small eddies are
modeled.

Solving only for the large eddies and modeling the smaller scales results in mesh resolution
requirements that are much less restrictive than with DNS. Typically, mesh sizes can be
at least one order of magnitude smaller than with DNS. Furthermore, the time step
sizes will be proportional to the eddy-turnover time, which is much less restrictive than
with DNS. In practical terms, however, extremely fine meshes are still required. It is
only due to the explosive increases in computer hardware performance coupled with the


c Fluent Inc. January 28, 2003 10-43
Modeling Turbulence

availability of parallel processing that LES can even be considered as a possibility for
engineering calculations.
The following sections give details of the governing equations for LES, present the two op-
tions for modeling the subgrid-scale stresses (necessary to achieve closure of the governing
equations), and discuss the relevant boundary conditions.

10.7.1 Filtered Navier-Stokes Equations


The governing equations employed for LES are obtained by filtering the time-dependent
Navier-Stokes equations in either Fourier (wave-number) space or configuration (physical)
space. The filtering process effectively filters out the eddies whose scales are smaller than
the filter width or grid spacing used in the computations. The resulting equations thus
govern the dynamics of large eddies.
A filtered variable (denoted by an overbar) is defined by
Z
φ(x) = φ(x0 )G(x, x0 )dx0 (10.7-1)
D

where D is the fluid domain, and G is the filter function that determines the scale of the
resolved eddies.
In FLUENT, the finite-volume discretization itself implicitly provides the filtering opera-
tion:

1 Z
φ(x) = φ(x0 ) dx0 , x0 ∈ V (10.7-2)
V V

where V is the volume of a computational cell. The filter function, G(x, x0 ), implied here
is then
(
0 1/V, x0 ∈ V
G(x, x ) (10.7-3)
0, x0 otherwise

Since the application of LES to compressible flows is still in its infancy, the theory is
presented here for incompressible flows. It is assumed that the LES model in FLUENT
will be applied to essentially incompressible (but not necessarily constant-density) flows.
Filtering the incompressible Navier-Stokes equations, one obtains

∂ρ ∂
+ (ρui ) = 0 (10.7-4)
∂t ∂xi

and

10-44
c Fluent Inc. January 28, 2003
10.7 The Large Eddy Simulation (LES) Model

!
∂ ∂ ∂ ∂ui ∂p ∂τij
(ρui ) + (ρui uj ) = µ − − (10.7-5)
∂t ∂xj ∂xj ∂xj ∂xi ∂xj

where τij is the subgrid-scale stress defined by

τij ≡ ρui uj − ρui uj (10.7-6)

The similarity between the filtered equations, 10.7-4 through 10.7-6, and the incompress-
ible form of the RANS equations, Equations 10.2-3 and 10.2-4, is obvious. The major
difference is that the dependent variables are now filtered quantities rather than mean
quantities, and the expressions for the turbulent stresses differ.

10.7.2 Subgrid-Scale Models


The subgrid-scale stresses resulting from the filtering operation are unknown, and require
modeling. The majority of subgrid-scale models in use today are eddy viscosity models
of the following form:

1
τij − τkk δij = −2µt S ij (10.7-7)
3

where µt is the subgrid-scale turbulent viscosity, and S ij is the rate-of-strain tensor for
the resolved scale defined by
!
1 ∂ui ∂uj
S ij ≡ + (10.7-8)
2 ∂xj ∂xi

FLUENT contains two models for µt : the Smagorinsky-Lilly model and the RNG-based
subgrid-scale model.

Smagorinsky-Lilly Model
The most basic of subgrid-scale models was proposed by Smagorinsky [239] and further
developed by Lilly [152]. In the Smagorinsky-Lilly model, the eddy viscosity is modeled
by


µt = ρL2s S (10.7-9)
q
where Ls is the mixing length for subgrid scales and S ≡ 2S ij S ij . Cs is the Smagorin-
sky constant. In FLUENT, Ls is computed using


c Fluent Inc. January 28, 2003 10-45
Modeling Turbulence

 
Ls = min κd, Cs V 1/3 (10.7-10)

where κ is the von Kármán constant, d is the distance to the closest wall, and V is the
volume of the computational cell.
Lilly derived a value of 0.23 for Cs from homogeneous isotropic turbulence in the inertial
subrange. However, this value was found to cause excessive damping of large-scale fluc-
tuations in the presence of mean shear or in transitional flows. Cs =0.1 has been found
to yield the best results for a wide range of flows, and is the default value in FLUENT.

RNG-Based Subgrid-Scale Model


Renormalization group (RNG) theory can be used to derive a model for the subgrid-
scale eddy viscosity [298]. The RNG procedure results in an effective subgrid viscosity,
µeff = µ + µt , given by

µeff = µ [1 + H(x)]1/3 (10.7-11)

H(x) is the Heaviside function:


(
x, x > 0
H(x) = (10.7-12)
0, x ≤ 0

where

µ2s µeff
x= −C (10.7-13)
µ3

and
q
µs = ρ(Crng V 1/3 )2 2S ij S ij (10.7-14)

where V is the volume of the computational cell. The theory gives Crng = 0.157 and
C = 100.
In highly turbulent regions of the flow (µt  µ), µeff ≈ µs , and the RNG-based subgrid-
scale model reduces to the Smagorinsky-Lilly model with a different model constant. In
low-Reynolds-number regions of the flow, the argument of the ramp function becomes
negative and the effective viscosity recovers molecular viscosity. This enables the RNG-
based subgrid-scale eddy viscosity to model the low-Reynolds-number effects encountered
in transitional flows and near-wall regions.

10-46
c Fluent Inc. January 28, 2003
10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

10.7.3 Boundary Conditions for the LES Model


The stochastic components of the flow at the velocity-specified inlet boundaries are ac-
counted for by superposing random perturbations on individual velocity components as

ui =< ui > +I ψ |u| (10.7-15)

where I is the
q intensity of the fluctuation, ψ is a Gaussian random number satisfying
ψ = 0, and ψ 0 = 1.
When the mesh is fine enough to resolve the laminar sublayer, the wall shear stress is
obtained from the laminar stress-strain relationship:

u ρuτ y
= (10.7-16)
uτ µ

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 1 ρuτ y
= ln E (10.7-17)
uτ κ µ

where κ is the von Kármán constant and E = 9.793.

10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows


10.8.1 Overview
Turbulent flows are significantly affected by the presence of walls. Obviously, the mean
velocity field is affected through the no-slip condition that has to be satisfied 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 fluctuations, while
kinematic blocking reduces the normal fluctuations. 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 significantly impacts the fidelity 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 flow in
the near-wall region determines successful predictions of wall-bounded turbulent flows.
The k- models, the RSM, and the LES model are primarily valid for turbulent core
flows (i.e., the flow in the regions somewhat far from walls). Consideration therefore


c Fluent Inc. January 28, 2003 10-47
Modeling Turbulence

needs to be given as to how to make these models suitable for wall-bounded flows. The
Spalart-Allmaras and k-ω models were designed to be applied throughout the boundary
layer, provided that the near-wall mesh resolution is sufficient.
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 flow 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 effects of molecular viscosity and turbulence are equally
important. Figure 10.8.1 illustrates these subdivisions of the near-wall region, plotted in
semi-log coordinates.

U/Uτ = 2.5 ln(U τ y/ ν) + 5.45

inner layer

U/Uτ = Uτ y/ν
U/Uτ

outer layer

fully turbulent region Upper limit


buffer layer depends on
or
or
log-law region Reynolds no.
blending
viscous sublayer region

y+ ∼
−5 ∼ 60
y+ −
ln Uτ y/ ν

Figure 10.8.1: Subdivisions of the Near-Wall Region

Wall Functions vs. Near-Wall Model


Traditionally, there are two approaches to modeling the near-wall region. In one ap-
proach, the viscosity-affected inner region (viscous sublayer and buffer layer) is not re-
solved. Instead, semi-empirical formulas called “wall functions” are used to bridge the
viscosity-affected 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 modified to enable the viscosity-affected
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 10.8.2.

10-48
c Fluent Inc. January 28, 2003
10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

turbulent core
? buffer &
sublayer

Wall Function Approach Near-Wall Model Approach

The viscosity-affected region is not The near-wall region is resolved


resolved, instead is bridged by the all the way down to the wall.
wall functions.

High-Re turbulence models can be The turbulence models ought to be valid


used. throughout the near-wall region.

Figure 10.8.2: Near-Wall Treatments in FLUENT

In most high-Reynolds-number flows, the wall function approach substantially saves com-
putational resources, because the viscosity-affected 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 flow simulations.
The wall function approach, however, is inadequate in situations where the low-Reynolds-
number effects are pervasive in the flow 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-affected 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-ω, and LES


Models
See Sections 10.3.7, 10.5.1, and 10.7.3, respectively, for a description of the near-wall
treatments applied by the Spalart-Allmaras, k-ω, and LES models.

10.8.2 Wall Functions


Wall functions are a collection of semi-empirical formulas and functions that in effect
“bridge” or “link” the solution variables at the near-wall cells and the corresponding
quantities on the wall. The wall functions comprise


c Fluent Inc. January 28, 2003 10-49
Modeling Turbulence

• laws-of-the-wall for mean velocity and temperature (or other scalars)

• formulas for near-wall turbulent quantities

FLUENT offers 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 [141], and have been most widely used for industrial flows. They are provided as a
default option in FLUENT.

Momentum

The law-of-the-wall for mean velocity yields

1
U∗ = ln(Ey ∗ ) (10.8-1)
κ
where

1/2
UP Cµ1/4 kP

U ≡ (10.8-2)
τw /ρ

1/2
ρCµ1/4 kP yP

y ≡ (10.8-3)
µ

and κ = von Kármán constant (= 0.42)


E = empirical constant (= 9.793)
UP = mean velocity of the fluid at point P
kP = turbulence kinetic energy at point P
yP = distance from point P to the wall
µ = dynamic viscosity of the fluid
The logarithmic law for mean velocity is known to be valid for y ∗ > about 30 to 60. 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

10-50
c Fluent Inc. January 28, 2003
10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

U ∗ = y∗ (10.8-4)

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 + (≡ ρuτ y/µ). 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 different laws:

• linear law for the thermal conduction sublayer where conduction is important

• logarithmic law for the turbulent region where effects of turbulence dominate con-
duction

The thickness of the thermal conduction layer is, in general, different from the thickness
of the (momentum) viscous sublayer, and changes from fluid to fluid. For example, the
thickness of the thermal sublayer for a high-Prandtl-number fluid (e.g., oil) is much less
than its momentum sublayer thickness. For fluids of low Prandtl numbers (e.g., liquid
metal), on the contrary, it is much larger than the momentum sublayer thickness.
In highly compressible flows, the temperature distribution in the near-wall region can
be significantly different from that of low subsonic flows, due to the heating by viscous
dissipation. In FLUENT, the temperature wall functions include the contribution from
the viscous heating [273].
The law-of-the-wall implemented in FLUENT has the following composite form:

1/4 1/2

C k



 Pr y ∗ + 12 ρPr µ q̇ P UP2 (y ∗ < yT∗ )
1/2 h i
(Tw − TP ) ρcp Cµ1/4 kP

 Pr 1 ln(Ey ∗ ) + P +

∗ t κ
T ≡ = 1/4 1/2
q̇ 

 1 C µ kP
ρ {Prt UP2 + (Pr − Prt )Uc2 } (y ∗ > yT∗ )


 2 q̇

(10.8-5)
where P is computed by using the formula given by Jayatilleke [117]:

" 3/4 #
Pr h i
P = 9.24 −1 1 + 0.28e−0.007Pr/Prt (10.8-6)
Prt

and


c Fluent Inc. January 28, 2003 10-51
Modeling Turbulence

kf = thermal conductivity of fluid


ρ = density of fluid
cp = specific heat of fluid
q̇ = wall heat flux
TP = temperature at the cell adjacent to wall
Tw = temperature at the wall
Pr = molecular Prandtl number (µcp /kf )
Prt = turbulent Prandtl number (0.85 at the wall)
A = 26 (Van Driest constant)
κ = 0.4187 (von Kármán constant)
E = 9.793 (wall function constant)
Uc = mean velocity magnitude at y ∗ = yT∗
Note that, for the segregated solver, the terms

1/2
1 C 1/4 k
ρPr µ P UP2
2 q̇
and
1/2
1 Cµ1/4 kP n o
ρ Prt UP2 + (Pr − Prt )Uc2
2 q̇

will be included in Equation 10.8-5 only for compressible flow calculations.


The non-dimensional thermal sublayer thickness, yT∗ , in Equation 10.8-5 is computed as
the y ∗ value at which the linear law and the logarithmic law intersect, given the molecular
Prandtl number of the fluid being modeled.
The procedure of applying the law-of-the-wall for temperature is as follows. Once the
physical properties of the fluid being modeled are specified, its molecular Prandtl number
is computed. Then, given the molecular Prandtl number, the thermal sublayer thickness,
yT∗ , is computed from the intersection of the linear and logarithmic profiles, and stored.
During the iteration, depending on the y ∗ value at the near-wall cell, either the linear or
the logarithmic profile in Equation 10.8-5 is applied to compute the wall temperature Tw
or heat flux q̇ (depending on the type of the thermal boundary conditions).

Species

When using wall functions for species transport, FLUENT assumes that species transport
behaves analogously to heat transfer. Similarly to Equation 10.8-5, the law-of-the-wall
for species can be expressed for constant property flow with no viscous dissipation as

1/2
(Yi,w − Yi ) ρCµ1/4 kP Sc yh∗ (y ∗ < yc∗ )
(

Y ≡ = i (10.8-7)
Ji,w Sct κ1 ln(Ey ∗ ) + Pc (y ∗ > yc∗ )

10-52
c Fluent Inc. January 28, 2003
10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

where Yi is the local species mass fraction, Sc and Sct are molecular and turbulent
Schmidt numbers, and Ji,w is the diffusion flux of species i at the wall. Note that Pc and
yc∗ are calculated in a similar way as P and yT∗ , with the difference 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
=0 (10.8-8)
∂n
where n is the local coordinate normal to the wall.
The production of kinetic energy, Gk , 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

∂U τw
G k ≈ τw = τw 1/4 1/2
(10.8-9)
∂y κρCµ kP yP

and  is computed from

3/2
C 3/4 k
P = µ P (10.8-10)
κyP

The  equation is not solved at the wall-adjacent cells, but instead is computed using
Equation 10.8-10.
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 flows.
However, they tend to become less reliable when the flow 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 flows are subjected to


c Fluent Inc. January 28, 2003 10-53
Modeling Turbulence

severe pressure gradients, and when the flows are in strong non-equilibrium, the quality
of the predictions is likely to be compromised.
The non-equilibrium wall functions offered as an additional option can improve the results
in such situations.

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 [128] is also available. The
key elements in the non-equilibrium wall functions are as follows:

• Launder and Spalding’s log-law for mean velocity is sensitized to pressure-gradient


effects.

• The two-layer-based concept is adopted to compute the budget of turbulence kinetic


energy (Gk ,) 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

Ũ Cµ1/4 k 1/2 ρCµ1/4 k 1/2 y


!
1
= ln E (10.8-11)
τw /ρ κ µ

where

y − yv y 2
" ! #
1 dp yv y
Ũ = U − √ ln + √ + v (10.8-12)
2 dx ρκ k yv ρκ k µ

and yv is the physical viscous sublayer thickness, and is computed from

µyv∗
yv ≡ 1/4 1/2
(10.8-13)
ρCµ kP

where yv∗ = 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 profile assumptions for
turbulence quantities are made:

10-54
c Fluent Inc. January 28, 2003
10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

  2 
( 2νk
0, y < yv  y
k P , y < yv 
y2
, y < yv
τt = k =  yv = k3/2 (10.8-14)
τw , y > yv kP , y > yv C` y
, y > yv

where C` = κCµ−3/4 , and yv is the dimensional thickness of the viscous sublayer, defined
in Equation 10.8-13.
Using these profiles, the cell-averaged production of k, Gk , and the cell-averaged dissipa-
tion rate, , can be computed from the volume average of Gk and  of the wall-adjacent
cells. For quadrilateral and hexahedral cells for which the volume average can be ap-
proximated with a depth-average,

τw2
!
1 Z yn ∂U 1 yn
Gk ≡ τt dy = ln (10.8-15)
yn 0 ∂y κyn ρCµ1/4 kP1/2 yv

and
 
1/2 !
1 Z yn 1  2ν kP yn 
≡  dy = + ln kP (10.8-16)
yn 0 yn yv C` yv

where yn is the height of the cell (yn = 2yP ). For cells with other shapes (e.g., triangular
and tetrahedral grids), the appropriate volume averages are used.
In Equations 10.8-15 and 10.8-16, the turbulence kinetic energy budget for the wall-
neighboring cells is effectively sensitized to the proportions of the viscous sublayer and
the fully turbulent layer, which varies widely from cell to cell in highly non-equilibrium
flows. It effectively 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
effect, partly account for non-equilibrium effects neglected in the standard wall function.

Standard Wall Functions vs. Non-Equilibrium Wall Functions


Because of the capability to partly account for the effects of pressure gradients and
departure from equilibrium, the non-equilibrium wall functions are recommended for use
in complex flows involving separation, reattachment, and impingement where the mean
flow and turbulence are subjected to severe pressure gradients and change rapidly. In
such flows, improvements can be obtained, particularly in the prediction of wall shear
(skin-friction coefficient) and heat transfer (Nusselt or Stanton number).

Limitations of the Wall Function Approach


The standard wall functions give reasonably accurate predictions for the majority of
high-Reynolds-number, wall-bounded flows. The non-equilibrium wall functions further


c Fluent Inc. January 28, 2003 10-55
Modeling Turbulence

extend the applicability of the wall function approach by including the effects of pressure
gradient and strong non-equilibrium. However, the wall function approach becomes less
reliable when the flow conditions depart too much from the ideal conditions underlying
the wall functions. Examples are as follows:

• Pervasive low-Reynolds-number or near-wall effects (e.g., flow through a small gap


or highly viscous, low-velocity fluid flow)

• Massive transpiration through the wall (blowing/suction)

• Severe pressure gradients leading to boundary layer separations

• Strong body forces (e.g., flow near rotating disks, buoyancy-driven flows)

• High three-dimensionality in the near-wall region (e.g., Ekman spiral flow, strongly
skewed 3D boundary layers)

If any of the items listed above is a prevailing feature of the flow 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.

10.8.3 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 fine 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 sufficiently fine 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 fine meshes (low-Reynolds-number meshes). In addition, excessive
error should not be incurred for intermediate meshes that are too fine 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 fine near-wall meshes and that, at the same time,
will not significantly reduce accuracy for wall-function meshes, FLUENT can combine the
two-layer model with enhanced wall functions, as described in the following sections.

10-56
c Fluent Inc. January 28, 2003
10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

Two-Layer Model for Enhanced Wall Treatment


In FLUENT’s near-wall model, the viscosity-affected 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-affected region and a fully-turbulent region. The demarcation of the two re-
gions is determined by a wall-distance-based, turbulent Reynolds number, Rey , defined
as

ρy k
Rey ≡ (10.8-17)
µ

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 k~r − ~rw k (10.8-18)


rw ∈Γw
~

where ~r is the position vector at the field point, and ~rw is the position vector on the
wall boundary. Γw is the union of all the wall boundaries involved. This interpretation
allows y to be uniquely defined in flow domains of complex shape involving multiple
walls. Furthermore, y defined in this way is independent of the mesh topology used, and
is definable even on unstructured meshes.
In the fully turbulent region (Rey > Re∗y ; Re∗y = 200), the k- models or the RSM
(described in Sections 10.4 and 10.6) are employed.
In the viscosity-affected near-wall region (Rey < Re∗y ), the one-equation model of Wolf-
stein [296] is employed. In the one-equation model, the momentum equations and the
k equation are retained as described in Sections 10.4 and 10.6. However, the turbulent
viscosity, µt , is computed from


µt,2layer = ρ Cµ `µ k (10.8-19)

where the length scale that appears in Equation 10.8-19 is computed from [37]

 
`µ = yc` 1 − e−Rey /Aµ (10.8-20)

The two-layer formulation for turbulent viscosity described above is used as a part of the
enhanced wall treatment, in which the two-layer definition is smoothly blended with the
high-Reynolds-number µt definition from the outer region, as proposed by Jongen [119]:

µt,enh = λ µt + (1 − λ )µt,2layer (10.8-21)


c Fluent Inc. January 28, 2003 10-57
Modeling Turbulence

where µt is the high-Reynolds-number definition as described in Section 10.4 or 10.6 for


the k- models or the RSM. A blending function, λ , is defined 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

Rey − Re∗y
" !#
1
λ = 1 + tanh (10.8-22)
2 A

The constant A determines the width of the blending function. By defining a width such
that the value of λ will be within 1% of its far-field value given a variation of ∆Rey , the
result is

|∆Rey |
A= (10.8-23)
tanh(0.98)

Typically, ∆Rey would be assigned a value that is between 5% and 20% of Re∗y . The
main purpose of the blending function λ 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  field is computed from

k 3/2
= (10.8-24)
`

The length scales that appear in Equation 10.8-24 are again computed from Chen and
Patel [37]:

 
` = yc` 1 − e−Rey /A (10.8-25)

If the whole flow domain is inside the viscosity-affected region (Rey < 200),  is not
obtained by solving the transport equation; it is instead obtained algebraically from
Equation 10.8-24. FLUENT uses a procedure for the  specification that is similar to the
µt blending in order to ensure a smooth transition between the algebraically-specified
 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 10.8-20 and 10.8-25, are taken
from [37]:

c` = κCµ−3/4 , Aµ = 70, A = 2c` (10.8-26)

10-58
c Fluent Inc. January 28, 2003
10.8 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, buffer 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 [121]:

1
u+ = eΓ u+ +
lam + e uturb
Γ (10.8-27)

where the blending function is given by:

a(y + )4
Γ = − (10.8-28)
1 + by +
E
 
c = exp − 1.0 (10.8-29)
E 00
a = 0.01c (10.8-30)
5
b = (10.8-31)
c

where E = 9.793 and E 00 is equal to E/fr , where fr is a roughness function. See Sec-
tion 6.13.1 for more information about wall roughness effects.
du+
Similarly, the general equation for the derivative dy +
is

+ +
du+ Γ dulam 1 du
turb
+
=e +
+e Γ (10.8-32)
dy dy dy +

This approach allows the fully turbulent law to be easily modified and extended to take
into account other effects 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 profiles in the cases where y + falls inside the wall
buffer 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 flow with heat transfer and pressure gradients has been derived by combining the
approaches of White and Cristoph [293] and Huang et al. [107]:

du+
turb 1 h 0 + + 2 1/2
i
= S (1 − βu − γ(u ) ) (10.8-33)
dy + κy +

where


c Fluent Inc. January 28, 2003 10-59
Modeling Turbulence

(
0 1 + αy + for y + < ys+
S = (10.8-34)
1 + αys+ for y + ≥ ys+

and

νw dp µ dp
α ≡ ∗
= 2 ∗ 3 (10.8-35)
τw u dx ρ (u ) dx

σt qw u σt qw
β ≡ = (10.8-36)
cp τw Tw ρcp u∗ Tw
σt (u∗ )2
γ ≡ (10.8-37)
2cp Tw

where ys+ is the location at which the log-law slope will remain fixed. By default, ys+ = 60.
The coefficient α in Equation 10.8-33 represents the influences of pressure gradients
while the coefficients β and γ represent thermal effects. Equation 10.8-33 is an ordinary
differential equation and FLUENT will provide an appropriate analytical solution. If α, β,
and γ 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
= 1 + αy + (10.8-38)
dy +
Note that the above expression only includes effects of pressure gradients through α,
while the effects of variable properties due to heat transfer and compressibility on the
laminar wall law are neglected. These effects are neglected because they are thought to be
of minor importance when they occur close to the wall. Integration of Equation 10.8-38
results in

α
 
u+
lam =y +
1 + y+ (10.8-39)
2
Enhanced thermal wall functions follow the same approach developed for the profile of
u+ . The unified wall thermal formulation blends the laminar and logarithmic profiles
according to the method of Kader [121]:

1
+ +
T + = eΓ Tlam + e Γ Tturb (10.8-40)
where
a(Pr y + )4
Γ = − (10.8-41)
1 + bPr3 y +

10-60
c Fluent Inc. January 28, 2003
10.9 Grid Considerations for Turbulent Flow Simulations

where Pr is the molecular Prandtl number, and the coefficients a and b are defined as
in Equations 10.8-30 and 10.8-31. Apart from the above formulation for T + , enhanced
thermal wall functions follow the same logic as previously described for standard thermal
wall functions (see Section 10.8.2). A similar procedure is also used for species wall
functions when the enhanced wall treatment is used. See Section 10.8.2 for details about
species wall functions.
The boundary condition for turbulence kinetic energy is the same as for standard wall
functions (Equation 10.8-8). However, the production of turbulence kinetic energy Gk is
computed using the velocity gradients that are consistent with the enhanced law-of-the-
wall (Equations 10.8-27 and 10.8-32), ensuring a formulation that is valid throughout the
near-wall region.

10.9 Grid Considerations for Turbulent Flow Simulations


Successful computations of turbulent flows require some consideration during the mesh
generation. Since turbulence (through the spatially-varying effective viscosity) plays a
dominant role in the transport of mean momentum and other scalars for the majority
of complex turbulent flows, you must ascertain that turbulence quantities are properly
resolved, if high accuracy is required. Due to the strong interaction of the mean flow and
turbulence, the numerical results for turbulent flows tend to be more susceptible to grid
dependency than those for laminar flows.
It is therefore recommended that you resolve, with sufficiently fine meshes, the regions
where the mean flow 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
Rey , which are all available in the postprocessing panels. It should be remembered that
y + , y ∗ , and Rey are not fixed, 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, different strategies must be used depending on
which near-wall option you are using. In Sections 10.9.1 and 10.9.2 are general guidelines
for the near-wall mesh.

10.9.1 Near-Wall Mesh Guidelines for Wall Functions


The distance from the wall at the wall-adjacent cells must be determined by considering
the range over which the log-law is valid. The distance is usually measured in the wall
unit, y + (≡ ρuτ y/µ), or y ∗ . Note that y + and y ∗ have comparable values when the first
cell is placed in the log-layer.


c Fluent Inc. January 28, 2003 10-61
Modeling Turbulence

• It is known that the log-law is valid for y + > 30 to 60.

• Although FLUENT employs the linear (laminar) law when y + < 11.225, using an
excessively fine mesh near the walls should be avoided, because the wall functions
cease to be valid in the viscous sublayer.

• 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.

• A y + value close to the lower bound (y + ≈ 30) is most desirable.

• 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.

10.9.2 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-affected 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-affected near-wall region
(Rey < 200) to be able to resolve the mean velocity and turbulent quantities in
that region.

10.9.3 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-affected 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

10-62
c Fluent Inc. January 28, 2003
10.9 Grid Considerations for Turbulent Flow Simulations

the full benefit of the Spalart-Allmaras model, the near-wall mesh spacing should be as
described in Section 10.9.2 for the enhanced wall treatment.
However, as discussed in Section 10.3.7, the boundary conditions 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 10.9.1.
In summary, for best results with the Spalart-Allmaras model, you should use either a
very fine near-wall mesh spacing (on the order of y + = 1) or a mesh spacing such that
y + ≥ 30.

10.9.4 Near-Wall Mesh Guidelines for the k-ω Models


Both k-ω 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.

10.9.5 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 10.7.3. 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 fine near-wall mesh spacing (on the order of
y + = 1).


c Fluent Inc. January 28, 2003 10-63
Modeling Turbulence

10.10 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 flow 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 10.10.1).
Define −→ Models −→Viscous...

Figure 10.10.1: The Viscous Model Panel

10-64
c Fluent Inc. January 28, 2003
10.10 Problem Setup for Turbulent Flows

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.
! The Large Eddy Simulation model is available only for 3D cases.

2. If the flow 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
These near-wall options are described in detail in Section 10.8. By default, the
standard wall function is enabled.
The near-wall treatment for the Spalart-Allmaras, k-ω, and LES models is defined
automatically, as described in Sections 10.3.7, 10.5.1, and 10.7.3, respectively.

3. Enable the appropriate turbulence modeling options in the Viscous Model panel.
See Section 10.10.1 for details.

4. Specify the boundary conditions for the solution variables.


Define −→Boundary Conditions...
See Section 10.10.2 for details.

5. Specify the initial guess for the solution variables.


Solve −→ Initialize −→Initialize...
See Section 10.10.3 for details. Note that Reynolds stresses are automatically ini-
tialized using k, and therefore need not be initialized.

10.10.1 Turbulence Options


The various options available for the turbulence models are described in detail in Sec-
tions 10.3 through 10.7. 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)


c Fluent Inc. January 28, 2003 10-65
Modeling Turbulence

Once you have enabled the Spalart-Allmaras model, you can enable the DES model
using the define/models/viscous/turbulence-expert/detached-eddy-simulation?
text command.
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 effects on 

If you choose the RNG k- model, the following options are available:

• Differential viscosity model

• Swirl modification

• Viscous heating (always activated for the coupled solvers)

• Inclusion of buoyancy effects on 

If you choose the standard k-ω model, the following options are available:

• Transitional flows

• Shear flow corrections

• Viscous heating (always activated for the coupled solvers)

If you choose the shear-stress transport k-ω model, the following options are available:

• Transitional flows

• Viscous heating (always activated for the coupled solvers)

If you choose the RSM, the following options are available:

• Wall reflection effects 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)

10-66
c Fluent Inc. January 28, 2003
10.10 Problem Setup for Turbulent Flows

• Inclusion of buoyancy effects on 

If you choose the enhanced wall treatment (available for the k- models and the RSM),
the following options are available:

• Pressure gradient effects

• Thermal effects

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)

It is also possible to modify the Model Constants, but this is not necessary for most
applications. See Sections 10.3 through 10.7 for details about these constants. Note that
C1-PS and C2-PS are the constants C1 and C2 in the linear pressure-strain approximation
of Equations 10.6-5 and 10.6-6, and C1’-PS and C2’-PS are the constants C10 and C20 in
Equation 10.6-7. 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 C1 , C1∗ , C2 , C3 , C3∗ , C4 , and C5 in the quadratic
pressure-strain approximation of Equation 10.6-16.

Including the Viscous Heating Effects


See Sections 11.2.1 and 11.2.2 for information on including viscous heating effects 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 flow, the generation of turbulent kinetic energy due to buoy-
ancy (Gb in Equation 10.4-1) is, by default, always included in the k equation. However,
FLUENT does not, by default, include the buoyancy effects on .
To include the buoyancy effects on , you must turn on the Full Buoyancy Effects option
under Options in the Viscous Model panel.
This option is available for the three k- models and for the RSM.


c Fluent Inc. January 28, 2003 10-67
Modeling Turbulence

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 10.3-8 to
compute the value of the deformation tensor S; if you choose Strain/Vorticity-Based Pro-
duction, it will use Equation 10.3-10.
(These options will not appear unless you have activated the Spalart-Allmaras model.)

Detached Eddy Simulation (DES) Modeling


If you enabled DES for the Sparlart-Allmaras model as described at the beginning of this
section, FLUENT will use Equation 10.3-16 to compute the value of the length scale d. ˜
By default, the empirical constant Cdes is set to 0.65. You can change its value in the
Cdes field under Model Constants.

Differential Viscosity Modification


In the RNG turbulence model in FLUENT, you have an option to use a differential formula
for effective viscosity µeff (Equation 10.4-6) to account for low-Reynolds-number effects.
To enable this option, turn on Differential Viscosity Model under RNG Options in the
Viscous Model panel.
(This option will not appear unless you have activated the RNG k- model.)

Swirl Modification
Once you choose the RNG model, the swirl modification takes effect, by default, for all
three-dimensional flows and axisymmetric flows with swirl. The default swirl constant
(αs in Equation 10.4-8) is set to 0.05, which works well for weakly to moderately swirling
flows. However, for strongly swirling flows, you may need to use a larger swirl constant.
To change the value of the swirl constant, you must first 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 αs is increased to
0.07. You can change its value in the Swirl Factor field under Model Constants.

Transitional Flows
If either of the k-ω 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
coefficient (α∗ in Equation 10.5-6) is equal to 1.

10-68
c Fluent Inc. January 28, 2003
10.10 Problem Setup for Turbulent Flows

Shear Flow Corrections


In the standard k-ω model, you also have the option of including corrections to improve
the accuracy in predicting free shear flows. 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-ω model [294]. When this option is enabled, FLUENT will
calculate fβ∗ and fβ using Equations 10.5-16 and 10.5-24, respectively. If this option is
disabled, fβ∗ and fβ will be set equal to 1.

Including Pressure Gradient Effects


If the enhanced wall treatment is used, you may include the effects of pressure gradients
by enabling the Pressure Gradient Effects option under Enhanced Wall Treatment Options.
When this option is enabled, FLUENT will include the coefficient α in Equation 10.8-33.

Including Thermal Effects


If the enhanced wall treatment is used, you may include thermal effects by enabling
the Thermal Effects option under Enhanced Wall Treatment Options. When this option
is enabled, FLUENT will include the coefficient β in Equation 10.8-33. γ will also be
included in Equation 10.8-33 when the Thermal Effects option is enabled if the ideal gas
law is selected for the fluid density in the Materials panel.

Including the Wall Reflection Term


If the RSM is used with the default model for pressure strain, FLUENT will, by default,
include the wall-reflection effects in the pressure-strain term. That is, FLUENT will
calculate φij,w using Equation 10.6-7 and include it in Equation 10.6-4. Note that wall-
reflection effects are not included if you have selected the quadratic pressure-strain model.
! The empirical constants and the function f used in the calculation of φij,w are calibrated
for simple canonical flows such as channel flows and flat-plate boundary layers involving
a single wall. If the flow involves multiple walls and the wall has significant curvature
(e.g., an axisymmetric pipe or curvilinear duct), the inclusion of the wall-reflection term
in Equation 10.6-7 may not improve the accuracy of the RSM predictions. In such cases,
you can disable the wall-reflection effects by turning off the Wall Reflection Effects 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 10.6-28.
k is calculated by solving the k equation obtained by summing Equation 10.6-1 for
normal stresses. To disable this option and use the wall boundary conditions given in


c Fluent Inc. January 28, 2003 10-69
Modeling Turbulence

Equation 10.6-29, turn off 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 10.6.3, 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 Reflection Effects under Reynolds-Stress Options

• Enhanced Wall Treatment under Near-Wall Treatment

Subgrid-Scale Model
If you have selected the Large Eddy Simulation model, you will be able to choose which of
the two subgrid-scale models described in Section 10.7.2 is to be used. You can choose
either the Smagorinsky-Lilly or the RNG subgrid-scale 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-ω, or LES model, a user-defined function
can be used to customize the turbulent viscosity. This option will enable you to modify
µt in the case of the Spalart-Allmaras, k-, and k-ω models, and incorporate completely
new subgrid models in the case of the LES model. See the separate UDF Manual for
information about user-defined functions.
In the Viscous Model panel, under User-Defined Functions, select the appropriate user-
defined 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-ω model, a user-
defined function can be used to customize the turbulent Prandtl numbers. This option
will allow you to calculate σk and either σ or σω (depending on if you have enabled the
appropriate k- or k-ω model) by using a UDF. You will also be able to calculate the value
of the energy Prandtl number (Prt in Equation 10.4-23) and the Prandtl number at the
wall (Prt in Equation 10.8-5) in this way. See the separate UDF Manual for information
about user-defined functions.

10-70
c Fluent Inc. January 28, 2003
10.10 Problem Setup for Turbulent Flows

In the Viscous Model panel, under User-Defined Functions, select the appropriate user-
defined function from the drop-down lists under Prandtl Numbers. Options include: TKE
Prandtl Number, TDR Prandtl Number (k- models only), SDR Prandtl Number (k-ω model
only), Energy Prandtl Number, and Wall Prandtl Number.

Modeling Turbulence with Non-Newtonian Fluids


If the turbulent flow involves non-Newtonian fluids, 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 7.3.5 for details
about these options.

10.10.2 Defining Turbulence Boundary Conditions


k- Models and k-ω Models
When you are modeling turbulent flows in FLUENT using one of the k- models or one
of the k-ω models, you must provide the boundary conditions for k and  (or k and ω)
in addition to other mean solution variables. The boundary conditions for k and  (or k
and ω) 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 ω) 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 significantly affect the downstream flow.
See Section 6.2.2 for details about specifying the boundary conditions for k and  (or k
and ω) at the inlets.
You may want to include the effects 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 6.13.1).
Additionally, you can control whether or not to set the turbulent viscosity to zero within
a laminar zone. If the fluid 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 no, FLUENT will set only
the production term in the turbulence transport equation to zero, allowing µt inside the
laminar zone to be calculated. In contrast, when the Laminar Zone option is turned on in
a Fluid boundary condition panel, both µt and the production term are set to zero. See
Section 6.17.1 for details about laminar zones.
! Note that this laminar zone feature is also available for the Spalart-Allmaras model.


c Fluent Inc. January 28, 2003 10-71
Modeling Turbulence

The Spalart-Allmaras Model


When you are modeling turbulent flows in FLUENT using the Spalart-Allmaras model,
you must provide the boundary conditions for ν̃ in addition to other mean solution vari-
ables. The boundary conditions for ν̃ at the walls are internally taken care of by FLUENT,
which obviates the need for your inputs. The boundary condition input for ν̃ 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 significantly affect the downstream flow.
See Section 6.2.2 for details about specifying the boundary condition for ν̃ at the inlets.
You may want to include the effects 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 6.13.1).
! 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 specification of turbulent boundary conditions for the RSM is the same as for the
other turbulence models for all boundaries except at boundaries where flow 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 turbulence
quantities using any of the turbulence specification methods described in Section 6.2.2.
FLUENT then uses the specified turbulence quantities to derive the Reynolds stresses at
the inlet from the assumption of isotropy of turbulence:

0 2
ui2 = k (i = 1, 2, 3) (10.10-1)
3
u0i u0j = 0.0 (10.10-2)

0
where ui2 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 6.2.2). To use this method, you will select K or Turbulence Intensity as the
Reynolds-Stress Specification Method in the appropriate boundary condition panel.
Alternately, you can directly specify the Reynolds stresses by selecting Reynolds-Stress
Components as the Reynolds-Stress Specification Method in the boundary condition panel.
When this option is enabled, you should input the Reynolds stresses directly.

10-72
c Fluent Inc. January 28, 2003
10.10 Problem Setup for Turbulent Flows

You can set the Reynolds stresses by using constant values, profile functions of coordinates
(see Section 6.26), or user-defined functions (see the separate UDF Manual).

Large Eddy Simulation Model


It is possible to specify the magnitude of random fluctuations 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 10.7.3. For all boundary
types other than velocity inlets, the boundary conditions for LES remain the same as for
laminar flows.

10.10.3 Providing an Initial Guess for k and  (or k and ω)


For flows using one of the k- models, one of the k-ω models, or the RSM, the converged
solutions or (for unsteady calculations) the solutions after a sufficiently long time has
elapsed should be independent of the initial values for k and  (or k and ω). For better
convergence, however, it is beneficial to use a reasonable initial guess for k and  (or k
and ω).
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 fields. 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 ω) in the whole
domain from these boundary values. (See Section 24.13 for details.)

• For more complex flows (e.g., flows with multiple inlets with different conditions) it
may be better to specify the initial values in terms of turbulence intensity. 5–10%
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(Iuavg )2 ).
2
You should specify an initial guess for  so that the resulting eddy viscosity (Cµ k )
is sufficiently 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 10.10-1 and 10.10-2.


c Fluent Inc. January 28, 2003 10-73
Modeling Turbulence

Figure 10.10.2: Specifying Inlet Boundary Conditions for the Reynolds Stresses

10-74
c Fluent Inc. January 28, 2003
10.11 Solution Strategies for Turbulent Flow Simulations

10.11 Solution Strategies for Turbulent Flow Simulations


Compared to laminar flows, simulations of turbulent flows 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
(µt , k, , ω, or the Reynolds stresses) are strongly coupled in a highly non-linear fashion,
it takes more computational effort 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 fine mesh.
The fidelity of the results for turbulent flows is largely determined by the turbulence
model being used. Here are some guidelines that can enhance the quality of your turbulent
flow simulations.

10.11.1 Mesh Generation


The following are suggestions to follow when generating the mesh for use in your turbulent
flow simulation:

• Picture in your mind the flow under consideration using your physical intuition or
any data for a similar flow situation, and identify the main flow features expected
in the flow you want to model. Generate a mesh that can resolve the major features
that you expect.
• If the flow is wall-bounded, and the wall is expected to significantly affect the flow,
take additional care when generating the mesh. You should avoid using a mesh
that is too fine (for the wall function approach) or too coarse (for the enhanced
wall treatment approach). See Section 10.9 for details.

10.11.2 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 flow (see Section 10.2).
• Because the mean quantities have larger gradients in turbulent flows than in laminar
flows, 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 diffusion adversely affects the solution accuracy, even with the
most elaborate turbulence model.
• In some flow situations involving inlet boundaries, the flow 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 specified.


c Fluent Inc. January 28, 2003 10-75
Modeling Turbulence

10.11.3 Convergence
The suggestions below are provided to help you enhance convergence for turbulent flow
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 ω) fields. Particularly when the enhanced wall treatment
is used, it is important to start with a sufficiently developed turbulence field, as
recommended in Section 10.10.3, to avoid the need for an excessive number of
iterations to develop the turbulence field.

• 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 find during
the calculation that the residual for  is reported to be zero. This happens when your
flow is such that Rey is less than 200 in the entire flow domain, and  is obtained from
the algebraic formula (Equation 10.8-24) instead of from its transport equation.

10.11.4 RSM-Specific Solution Strategies


Using the RSM creates a high degree of coupling between the momentum equations and
the turbulent stresses in the flow, and thus the calculation can be more prone to stability
and convergence difficulties 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 flows or highly complex flows. In these cases,
you may need to reduce the under-relaxation factors both for the velocities and for
all of the stresses.

10-76
c Fluent Inc. January 28, 2003
10.11 Solution Strategies for Turbulent Flow Simulations

Instructions for setting these solution parameters are provided below. If you are applying
the RSM to prediction of a highly swirling flow, you will want to consider the solution
strategies discussed in Section 8.4 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 field remain fixed,
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.

10.11.5 LES-Specific Solution Strategies


Large eddy simulation involves running a transient solution from some initial condition,
on an appropriately fine 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 flow field to be determined.
The following are suggestions to follow when running a large eddy simulation:

1. Start by running a flow simulation assuming laminar flow or using a simple Reynolds-
averaged turbulence model such as standard k- or Spalart-Allmaras. Since this is
only an initial condition, you need run only until the flow field is somewhat con-
verged. This step is optional.


c Fluent Inc. January 28, 2003 10-77
Modeling Turbulence

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 24.15.1 for guidelines on setting solution parameters for transient calculations
in general.) Use the central-differencing spatial discretization scheme for all equa-
tions.

3. Run LES until the flow becomes statistically steady. The best way to see if the flow
is fully developed and statistically steady is to monitor forces and solution variables
(e.g., velocity components or pressure) at selected locations in the flow.

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 24.15.1.

5. Continue until you get statistically stable data. The duration of the simulation
can be determined beforehand by estimating the mean flow residence time in the
solution domain (L/U , where L is the characteristic length of the solution domain
and U is a characteristic mean flow velocity). The simulation should be run for at
least a few mean flow residence times.

Instructions for setting the solution parameters for LES are provided below.

Temporal Discretization
FLUENT provides both first-order and second-order temporal discretizations. For LES,
the second-order discretization is recommended.
Define −→ Models −→Solver...

Spatial Discretization
Overly diffusive schemes such as the first-order upwind or power law scheme should be
avoided, because they may unduly damp out the energy of the resolved eddies. The
central-differencing scheme is recommended for all equations when you use the LES
model.
Solve −→ Controls −→Solution...

10-78
c Fluent Inc. January 28, 2003
10.12 Postprocessing for Turbulent Flows

10.12 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

• Effective Viscosity

• Turbulent Viscosity Ratio

• Effective Thermal Conductivity

• Effective Prandtl Number

• Wall Yplus

• Wall Ystar

• Turbulent Reynolds Number (Re y) (only when the enhanced wall treatment is used
for the near-wall treatment)

Turbulence quantities that can be reported for the k-ω models are as follows:

• Turbulent Kinetic Energy (k)

• Turbulence Intensity

• Specific Dissipation Rate (Omega)

• Production of k

• Turbulent Viscosity

• Effective Viscosity

• Turbulent Viscosity Ratio


c Fluent Inc. January 28, 2003 10-79
Modeling Turbulence

• Effective Thermal Conductivity

• Effective Prandtl Number

• Wall Ystar

• Wall Yplus

Turbulence quantities that can be reported for the Spalart-Allmaras model are as follows:

• Modified Turbulent Viscosity

• Turbulent Viscosity

• Effective Viscosity

• Turbulent Viscosity Ratio

• Effective Thermal Conductivity

• Effective 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

• VV Reynolds Stress

• WW Reynolds Stress

• UV Reynolds Stress

• VW Reynolds Stress

• UW Reynolds Stress

• Turbulent Dissipation Rate (Epsilon)

• Production of k

• Turbulent Viscosity

• Effective Viscosity

10-80
c Fluent Inc. January 28, 2003
10.12 Postprocessing for Turbulent Flows

• Turbulent Viscosity Ratio

• Effective Thermal Conductivity

• Effective Prandtl Number

• Wall Yplus

• Wall Ystar

• Turbulent Reynolds Number (Re y)

Turbulence quantities that can be reported for the LES model are as follows:

• Subgrid Turbulent Kinetic Energy

• Subgrid Turbulent Viscosity

• Subgrid Effective Viscosity

• Subgrid Turbulent Viscosity Ratio

• Effective Thermal Conductivity

• Effective 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 29 for their definitions.

10.12.1 Custom Field Functions for Turbulence


In addition to the quantities listed above, you can define your own turbulence quantities
using the Custom Field Function Calculator panel.
Define −→Custom Field Functions...
The following functions may be useful:

• Ratio of production of k to its dissipation (Gk /ρ)

• Ratio of the mean flow to turbulent time scale, η (≡ Sk/)

• Reynolds stresses derived from the Boussinesq formula (e.g., −uv = νt ∂u


∂y
)


c Fluent Inc. January 28, 2003 10-81
Modeling Turbulence

10.12.2 Postprocessing LES Statistics


As described in Section 10.7, LES involves the solution of a transient flow field, but it is
the mean flow 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 24.15.3 for details.

10.12.3 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 field 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 field in order to see whether or not turbulence takes full effect. Usually turbulent
viscosity is at least two orders of magnitude larger than molecular viscosity for fully-
developed turbulent flows 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 filled contours of Rey (turbulent Reynolds
number) overlaid on the mesh.

10-82
c Fluent Inc. January 28, 2003

You might also like