Sie sind auf Seite 1von 112

Chapter 12.

Modeling Turbulence

This chapter provides details about the turbulence models available in FLUENT.
Information is presented in the following sections:
Section 12.1: Introduction
Section 12.2: Choosing a Turbulence Model
Section 12.3: Spalart-Allmaras Model Theory
Section 12.4: Standard, RNG, and Realizable k- Models Theory
Section 12.5: Standard and SST k- Models Theory
Section 12.6: The v 2 -f Model Theory
Section 12.7: Reynolds Stress Model (RSM) Theory
Section 12.8: Detached Eddy Simulation (DES) Model Theory
Section 12.9: Large Eddy Simulation (LES) Model Theory
Section 12.10: Near-Wall Treatments for Wall-Bounded Turbulent Flows
Section 12.11: Grid Considerations for Turbulent Flow Simulations
Section 12.12: Steps in Using a Turbulence Model
Section 12.13: Setting Up the Spalart-Allmaras Model
Section 12.14: Setting Up the k- Model
Section 12.15: Setting Up the k- Model
Section 12.16: Setting Up the Reynolds Stress Model
Section 12.17: Setting Up the Detached Eddy Simulation Model
Section 12.18: Setting Up the Large Eddy Simulation Model
Section 12.19: Setup Options for all Turbulence Modeling
Section 12.20: Defining Turbulence Boundary Conditions

c Fluent Inc. September 29, 2006


12-1

Modeling Turbulence

Section 12.21: Providing an Initial Guess for k and  (or k and )


Section 12.22: Solution Strategies for Turbulent Flow Simulations
Section 12.23: Postprocessing for Turbulent Flows

12.1

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.
FLUENT provides the following choices of turbulence models:
Spalart-Allmaras model
k- models
Standard k- model
Renormalization-group (RNG) k- model
Realizable k- model
k- models
Standard k- model
Shear-stress transport (SST) k- model
v 2 -f model (addon)
Reynolds stress model (RSM)
Linear pressure-strain RSM model
Quadratic pressure-strain RSM model
Low-Re stress-omega RSM model

12-2

c Fluent Inc. September 29, 2006


12.2 Choosing a Turbulence Model

Detached eddy simulation (DES) model


Spalart-Allmaras RANS model
Realizable k- RANS model
SST k- RANS model
Large eddy simulation (LES) model
Smagorinsky-Lilly subgrid-scale model
WALE subgrid-scale model
Kinetic-energy transport subgrid-scale model

12.2

Choosing a Turbulence Model

It is an unfortunate fact that no single turbulence model is universally accepted as being 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 appropriate 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.

12.2.1

Reynolds-Averaged Approach of the DES Model vs. LES

Time-dependent solutions of the Navier-Stokes equations for high Reynolds-number turbulent flows in complex geometries which set out to resolve all the way down to the
smallest scales of the motions are unlikely to be attainable for some time to come. Two
alternative methods can be employed to render the Navier-Stokes equations tractable
so that the small-scale turbulent fluctuations do not have to be directly simulated:
Reynolds-averaging (or ensemble-averaging) and filtering. Both methods introduce additional terms in the governing equations that need to be modeled in order to achieve a
closure for the unknowns.
The Reynolds-averaged Navier-Stokes (RANS) equations govern the transport of the averaged flow quantities, with the whole range of the scales of turbulence being modeled. The
RANS-based modeling approach therefore greatly reduces the required computational effort and resources, and is widely adopted for practical engineering applications. An entire
hierarchy of closure models are available in FLUENT including Spalart-Allmaras, k- and
its variants, k- and its variants, and the RSM. The RANS equations are often used

c Fluent Inc. September 29, 2006


12-3

Modeling Turbulence

to compute time-dependent flows, whose unsteadiness may be externally imposed (e.g.,


time-dependent boundary conditions or sources) or self-sustained (e.g., vortex-shedding,
flow instabilities).
LES provides an alternative approach in which large eddies are explicitly computed (resolved) in a time-dependent simulation using the filtered Navier-Stokes equations. The
rationale behind LES is that by modeling less of turbulence (and resolving more), the
error introduced by turbulence modeling can be reduced. It is also believed to be easier
to find a universal model for the small scales, since they tend to be more isotropic and
less affected by the macroscopic features like boundary conditions, than the large eddies.
Filtering is essentially a mathematical manipulation of the exact Navier-Stokes equations
to remove the eddies that are smaller than the size of the filter, which is usually taken as
the mesh size when spatial filtering is employed as in FLUENT. Like Reynolds-averaging,
the filtering process creates additional unknown terms that must be modeled to achieve
closure. Statistics of the time-varying flow-fields such as time-averages and r.m.s. values
of the solution variables, which are generally of most engineering interest, can be collected during the time-dependent simulation. LES for high Reynolds number industrial
flows requires a significant amount of compute resources. This is mainly because of the
need to accurately resolve the energy-containing turbulent eddies in both space and time
domains, which becomes most acute in near-wall regions where the scales to be resolved
become increasingly smaller. Wall functions in combination with a coarse near wall mesh
can be employed, often with some success, to reduce the cost of LES for wall-bounded
flows. However, one needs to carefully consider the ramification of using wall functions
for the flow in question. For the same reason (to accurately resolve the eddies), LES also
requires highly accurate spatial and temporal discretizations.

12.2.2

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 fluctuating components. For the velocity components:
ui = ui + u0i

(12.2-1)

where ui and u0i are the mean and fluctuating velocity components (i = 1, 2, 3).
Likewise, for pressure and other scalar quantities:
= + 0

(12.2-2)

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


Substituting expressions of this form for the flow variables into the instantaneous continuity and momentum equations and taking a time (or ensemble) average (and dropping

12-4

c Fluent Inc. September 29, 2006


12.2 Choosing a Turbulence Model

the overbar on the mean velocity, u) yields the ensemble-averaged momentum equations.
They can be written in Cartesian tensor form as:

+
(ui ) = 0
t xi

(12.2-3)

"

ui uj
2 ul

(ui ) +
(ui uj ) =
+

+
ij
t
xj
xi xj
xj
xi
3 xl

!#

(u0i u0j )
xj
(12.2-4)

Equations 12.2-3 and 12.2-4 are called Reynolds-averaged Navier-Stokes (RANS) equations. They have the same general form as the instantaneous Navier-Stokes equations,
with the velocities and other solution variables now representing ensemble-averaged (or
time-averaged) values. Additional terms now appear that represent the effects of turbulence. These Reynolds stresses, u0i u0j , must be modeled in order to close Equation 12.2-4.
For variable-density flows, Equations 12.2-3 and 12.2-4 can be interpreted as Favreaveraged Navier-Stokes equations [142], with the velocities representing mass-averaged
values. As such, Equations 12.2-3 and 12.2-4 can be applied to density-varying flows.

12.2.3

Boussinesq Approach vs. Reynolds Stress Transport Models

The Reynolds-averaged approach to turbulence modeling requires that the Reynolds


stresses in Equation 12.2-4 be appropriately modeled. A common method employs the
Boussinesq hypothesis [142] to relate the Reynolds stresses to the mean velocity gradients:

u0i u0j

= t

ui uj
+
xj
xi

2
uk

k + t
ij
3
xk

(12.2-5)

The Boussinesq hypothesis is used in the Spalart-Allmaras model, the k- models, and
the k- models. The advantage of this approach is the relatively low computational
cost associated with the computation of the turbulent viscosity, t . In the case of the
Spalart-Allmaras model, only one additional transport equation (representing turbulent
viscosity) is solved. In the case of the k- and k- models, two additional transport
equations (for the turbulence kinetic energy, k, and either the turbulence dissipation
rate, , or the specific dissipation rate, ) are solved, and t is computed as a function of
k and . The disadvantage of the Boussinesq hypothesis as presented is that it assumes
t is an isotropic scalar quantity, which is not strictly true.

c Fluent Inc. September 29, 2006


12-5

Modeling Turbulence

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 stressdriven secondary flows.

12.2.4

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 SpalartAllmaras 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 1015% more
CPU time than with the standard k- model. Like the k- models, the k- models are
also two-equation models, and thus require about the same computational effort.
Compared with the k- and k- models, the RSM requires additional memory and CPU
time due to the increased number of the transport equations for Reynolds stresses. However, efficient programming in FLUENT has reduced the CPU time per iteration significantly. On average, the RSM in FLUENT requires 5060% more CPU time per iteration
compared to the k- and k- models. Furthermore, 1520% more memory is needed.
Aside from the time per iteration, the choice of turbulence model can 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- models
due to the strong coupling between the Reynolds stresses and the mean flow.

12-6

c Fluent Inc. September 29, 2006


12.3 Spalart-Allmaras Model Theory

12.3

Spalart-Allmaras Model Theory

12.3.1

Overview

The Spalart-Allmaras model is a relatively simple one-equation model that solves a modeled 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- models. This might make the model less sensitive to numerical error
when non-layered meshes are used near walls. See Section 6.1.3: Numerical Diffusion 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 turbulence. 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
abruptly from a wall-bounded to a free shear flow.
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 [349] solves
a transport equation for a quantity that is a modified form of the turbulent kinematic
viscosity.

12.3.2

Transport Equation for the Spalart-Allmaras Model

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

(
)+
(
ui ) = G +
t
xi
xj

c Fluent Inc. September 29, 2006



( +
)
xj


+ Cb2
xj

!2
Y +S (12.3-1)

12-7

Modeling Turbulence

where G is the production of turbulent viscosity and Y is the destruction of turbulent


viscosity that occurs in the near-wall region due to wall blocking and viscous damping.
and Cb2 are constants and is the molecular kinematic viscosity. S 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 12.2-5 is ignored when estimating the
Reynolds stresses.

12.3.3

Modeling the Turbulent Viscosity

The turbulent viscosity, t , is computed from


t =
fv1

(12.3-2)

where the viscous damping function, fv1 , is given by


fv1 =

3
3
3 + Cv1

(12.3-3)

(12.3-4)

and

12.3.4

Modeling the Turbulent Production

The production term, G , is modeled as


G = Cb1 S

(12.3-5)

S S + 2 2 fv2
d

(12.3-6)

where

and
fv2 = 1

12-8

1 + fv1

(12.3-7)

c Fluent Inc. September 29, 2006


12.3 Spalart-Allmaras Model Theory

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

2ij ij

(12.3-8)

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


1
ij =
2

ui
uj

xj
xi

(12.3-9)

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 [73] and incorporated into FLUENT.
This modification combines measures of both rotation and strain tensors in the definition
of S:
S |ij | + Cprod min (0, |Sij | |ij |)

(12.3-10)

where
Cprod = 2.0, |ij |

2ij ij , |Sij |

2Sij Sij

with the mean strain rate, Sij , defined as


1
Sij =
2

uj
ui
+
xi xj

(12.3-11)

Including both the rotation and strain tensors reduces the production of eddy viscosity
and consequently reduces the eddy viscosity itself in regions where the measure of vorticity 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.

c Fluent Inc. September 29, 2006


12-9

Modeling Turbulence

12.3.5

Modeling the Turbulent Destruction

The destruction term is modeled as

Y = Cw1 fw

 2

(12.3-12)

where
6
1 + Cw3
fw = g 6
6
g + Cw3

"

#1/6

(12.3-13)

(12.3-14)

g = r + Cw2 r6 r

S2 d2

(12.3-15)

Cw1 , Cw2 , and Cw3 are constants, and S is given by Equation 12.3-6. Note that the
modification described above to include the effects of mean strain on S will also affect
the value of S used to compute r.

12.3.6

Model Constants

The model constants Cb1 , Cb2 , , Cv1 , Cw1 , Cw2 , Cw3 , and have the following default
values [349]:
2
Cb1 = 0.1355, Cb2 = 0.622, = , Cv1 = 7.1
3
Cw1 =

12.3.7

Cb1 (1 + Cb2 )
+
, Cw2 = 0.3, Cw3 = 2.0, = 0.4187
2

Wall Boundary Conditions

At walls, the modified turbulent kinematic viscosity, , 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
u y
=
u

12-10

(12.3-16)

c Fluent Inc. September 29, 2006


12.3 Spalart-Allmaras Model Theory

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
u y
= ln E
u

(12.3-17)

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

12.3.8

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:

(E) +
[ui (E + p)] =
t
xi
xj

"

c p t
k+
Prt

T
+ ui (ij )eff + Sh
xj

(12.3-18)

where k, in this case, is the thermal conductivity, E is the total energy, and (ij )eff is the
deviatoric stress tensor, defined as

(ij )eff = eff

uj
ui
+
xi xj

2
uk
eff
ij
3
xk

The term involving (ij )eff represents the viscous heating, and is always computed in the
density-based solvers. It is not computed by default in the pressure-based solver, but it
can be enabled in the Viscous Model 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.

c Fluent Inc. September 29, 2006


12-11

Modeling Turbulence

12.4

Standard, RNG, and Realizable k- Models Theory

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.

12.4.1

Standard k- Model

Overview
The simplest complete models of turbulence are two-equation models in which the solution 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 [196]. Robustness, 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, improvements have been made to the model to improve its performance. Two of these variants
are available in FLUENT: the RNG k- model [408] and the realizable k- model [330].
The standard k- model [196] is a semi-empirical model based on model transport equations for the turbulence kinetic energy (k) and its dissipation rate (). The model transport 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 mathematically exact counterpart.
In the derivation of the k- model, the assumption is 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.

12-12

c Fluent Inc. September 29, 2006


12.4 Standard, RNG, and Realizable k- Models Theory

Transport Equations for the Standard k- Model


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

(k) +
(kui ) =
t
xi
xj

"

t
+
k

k
+ Gk + Gb  YM + Sk
xj

(12.4-1)

and

() +
(ui ) =
t
xi
xj

"

t
+




2
+ C1 (Gk + C3 Gb ) C2 + S (12.4-2)
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 12.4.4: Modeling Turbulent
Production in the k- Models. Gb is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 12.4.5: Effects of Buoyancy on Turbulence in the k- Models. YM represents the contribution of the fluctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Section 12.4.6: Effects of Compressibility on Turbulence in the k- Models. C1 , C2 , and C3
are constants. k and  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, t , is computed by combining k and  as follows:
k2
t = C


(12.4-3)

where C is a constant.

Model Constants
The model constants C1 , C2 , C , k , and  have the following default values [196]:
C1 = 1.44, C2 = 1.92, C = 0.09, k = 1.0,  = 1.3

c Fluent Inc. September 29, 2006


12-13

Modeling Turbulence

These default values have been determined from experiments with air and water for fundamental 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 wallbounded 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.

12.4.2

RNG k- Model

Overview
The RNG k- model was derived using a rigorous statistical technique (called renormalization 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.
While the standard k- model is a high-Reynolds-number model, the RNG theory
provides an analytically-derived differential formula for effective viscosity that accounts for low-Reynolds-number effects. Effective use of this feature does, however,
depend on an appropriate treatment of the near-wall region.
These features make the RNG k- model more accurate and reliable for a wider class of
flows than the standard k- model.
The RNG-based k- turbulence model is derived from the instantaneous Navier-Stokes
equations, using a mathematical technique called renormalization group (RNG) methods. 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 [59].

12-14

c Fluent Inc. September 29, 2006


12.4 Standard, RNG, and Realizable k- Models Theory

Transport Equations for the RNG k- Model


The RNG k- model has a similar form to the standard k- model:

(k) +
(kui ) =
t
xi
xj

k
k eff
xj

+ Gk + Gb  YM + Sk

(12.4-4)

and

() +
(ui ) =
t
xi
xj



2
 eff
+ C1 (Gk + C3 Gb ) C2 R + S (12.4-5)
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 12.4.4: Modeling Turbulent
Production in the k- Models. Gb is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 12.4.5: Effects of Buoyancy on Turbulence in the k- Models. YM represents the contribution of the fluctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Section 12.4.6: Effects of Compressibility on Turbulence in the k- Models. The quantities
k and  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:
2 k
d


= 1.72

1 + C

(12.4-6)

where

= eff /
C 100
Equation 12.4-6 is integrated to obtain an accurate description of how the effective turbulent transport varies with the effective Reynolds number (or eddy scale), allowing the
model to better handle low-Reynolds-number and near-wall flows.

c Fluent Inc. September 29, 2006


12-15

Modeling Turbulence

In the high-Reynolds-number limit, Equation 12.4-6 gives


t = C

k2


(12.4-7)

with C = 0.0845, derived using RNG theory. It is interesting to note that this value
of C is very close to the empirically-determined value of 0.09 used in the standard k-
model.
In FLUENT, by default, the effective viscosity is computed using the high-Reynoldsnumber form in Equation 12.4-7. However, there is an option available that allows you
to use the differential relation given in Equation 12.4-6 when you need to include lowReynolds-number effects.

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:

t = t0

k
f s , ,


(12.4-8)

where t0 is the value of turbulent viscosity calculated without the swirl modification
using either Equation 12.4-6 or Equation 12.4-7. is a characteristic swirl number evaluated within FLUENT, and s 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), s is set to
0.07. For strongly swirling flows, however, a higher value of s can be used.

Calculating the Inverse Effective Prandtl Numbers


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

1.3929

0 1.3929

0.6321

+ 2.3929



0 + 2.3929

0.3679

mol

=

eff

(12.4-9)

where 0 = 1.0. In the high-Reynolds-number limit (mol /eff  1), k =  1.393.

12-16

c Fluent Inc. September 29, 2006


12.4 Standard, RNG, and Realizable k- Models Theory

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
R =

C 3 (1 /0 ) 2
1 + 3
k

(12.4-10)

where Sk/, 0 = 4.38, = 0.012.


The effects of this term in the RNG  equation can be seen more clearly by rearranging
Equation 12.4-5. Using Equation 12.4-10, the third and fourth terms on the right-hand
side of Equation 12.4-5 can be merged, and the resulting  equation can be rewritten as

() +
(ui ) =
t
xi
xj


 eff
xj

2


+ C1 (Gk + C3 Gb ) C2

k
k

(12.4-11)

where C2
is given by

C2
C2 +

C 3 (1 /0 )
1 + 3

(12.4-12)

In regions where < 0 , the R term makes a positive contribution, and C2
becomes
larger than C2 . In the logarithmic layer, for instance, it can be shown that 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 ( > 0 ), 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 12.4-5 have values derived analytically by
the RNG theory. These values, used by default in FLUENT, are
C1 = 1.42, C2 = 1.68

c Fluent Inc. September 29, 2006


12-17

Modeling Turbulence

12.4.3

Realizable k- Model

Overview
The realizable k- model [330] 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.
To understand the mathematics behind the realizable k- model, consider combining
the Boussinesq relationship (Equation 12.2-5) and the eddy viscosity definition (Equation 12.4-3) to obtain the following expression for the normal Reynolds stress in an
incompressible strained mean flow:
2
U
u2 = k 2 t
3
x

(12.4-13)

Using Equation 12.4-3 for t t /, 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
 x
3C

(12.4-14)

Similarly, it can also be shown that the Schwarz inequality for shear stresses (u u 2
u2 u2 ; no summation over and ) can be violated when the mean strain rate is large.
The most straightforward way to ensure the realizability (positivity of normal stresses
and Schwarz inequality for shear stresses) is to make C variable by sensitizing it to
the mean flow (mean deformation) and the turbulence (k, ). The notion of variable
C is suggested by many modelers including Reynolds [303], and is well substantiated
by experimental evidence. For example, C is found to be around 0.09 in the inertial
sublayer of equilibrium boundary layers, and 0.05 in a strong homogeneous shear flow.

12-18

c Fluent Inc. September 29, 2006


12.4 Standard, RNG, and Realizable k- Models Theory

Both the realizable and RNG k- models have shown substantial improvements over the
standard k- model where the flow features include strong streamline curvature, vortices,
and rotation. Since the model is still relatively new, it is not clear in exactly which
instances the realizable k- model consistently outperforms the RNG model. However,
initial studies have shown that the realizable model provides the best performance of all
the k- model versions for several validations of separated flows and flows with complex
secondary flow features.
One of the weaknesses of the standard k- model or other traditional k- models lies with
the modeled equation for the dissipation rate (). The well-known round-jet anomaly
(named based on the finding that the spreading rate in planar jets is predicted reasonably
well, but prediction of the spreading rate for axisymmetric jets is unexpectedly poor) is
considered to be mainly due to the modeled dissipation equation.
The realizable k- model proposed by Shih et al. [330] was intended to address these
deficiencies of traditional k- models by adopting the following:
A new eddy-viscosity formula involving a variable C originally proposed by
Reynolds [303].
A new model equation for dissipation () based on the dynamic equation of the
mean-square vorticity fluctuation.
One limitation of the realizable k- model is that it produces non-physical turbulent
viscosities in situations when the computational domain contains both rotating and stationary 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 12.4-1712.4-19). 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 12.4.3: Modeling the Turbulent Viscosity for information about how to include
or exclude this term from the model.

c Fluent Inc. September 29, 2006


12-19

Modeling Turbulence

Transport Equations for the Realizable k- Model


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

(k) +
(kuj ) =
t
xj
xj

"

t
+
k

k
+ Gk + Gb  YM + Sk
xj

(12.4-15)

and

() +
(uj ) =
t
xj
xj

"

t
+



2

+ C1 C3 Gb + S
+ C1 S C2
xj
k + 
k
(12.4-16)
#

where
"

C1 = max 0.43,
,
+5

k
=S ,


S=

2Sij Sij

In these equations, Gk represents the generation of turbulence kinetic energy due to the
mean velocity gradients, calculated as described in Section 12.4.4: Modeling Turbulent
Production in the k- Models. Gb is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 12.4.5: Effects of Buoyancy on Turbulence in the k- Models. YM represents the contribution of the fluctuating dilatation in
compressible turbulence to the overall dissipation rate, calculated as described in Section 12.4.6: Effects of Compressibility on Turbulence in the k- Models. C2 and C1 are
constants. k and  are the turbulent Prandtl numbers for k and , respectively. Sk and
S are user-defined source terms.
Note that the k equation (Equation 12.4-15) is the same as that in the standard k model (Equation 12.4-1) and the RNG k- model (Equation 12.4-4), 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 12.4-2 and 12.4-5). One of the
noteworthy features is that the production term in the  equation (the second term on
the right-hand side of Equation 12.4-16) 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 12.4-16)
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.

12-20

c Fluent Inc. September 29, 2006


12.4 Standard, RNG, and Realizable k- Models Theory

This model has been extensively validated for a wide range of flows [183, 330], 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
t = C


(12.4-17)

The difference between the realizable k- model and the standard and RNG k- models
is that C is no longer constant. It is computed from
1

A0 + As kU

(12.4-18)

ij
ij
Sij Sij +

(12.4-19)

C =
where

and

ij = ij 2ijk k

ij = ij ijk k
where ij is the mean rate-of-rotation tensor viewed in a rotating reference frame with
the angular velocity k . The model constants A0 and As are given by
A0 = 4.04, As =

6 cos

where

1
Sij Sjk Ski q
1
= cos1 ( 6W ), W =
,
S
=
S
ij Sij , Sij =
3
2
S3

c Fluent Inc. September 29, 2006


uj
ui
+
xi xj

12-21

Modeling Turbulence

It can be seen that C is a function of the mean strain and rotation rates, the angular velocity of the system rotation, and the turbulence fields (k and ). C in Equation 12.4-17
can be shown to recover the standard value of 0.09 for an inertial sublayer in an equilibrium boundary layer.

In FLUENT, the term 2ijk k is, by default, not included in the


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

Model Constants
The model constants C2 , k , and  have been established to ensure that the model
performs well for certain canonical flows. The model constants are
C1 = 1.44, C2 = 1.9, k = 1.0,  = 1.2

12.4.4

Modeling Turbulent Production in the k- Models

The term Gk , representing the production of turbulence kinetic energy, is modeled identically for the standard, RNG, and realizable k- models. From the exact equation for
the transport of k, this term may be defined as
Gk = u0i u0j

uj
xi

(12.4-20)

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


G k = t S 2

(12.4-21)

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


S

12-22

2Sij Sij

(12.4-22)

When using the high-Reynolds number k- versions, eff is used in lieu of
t in Equation 12.4-21.

c Fluent Inc. September 29, 2006


12.4 Standard, RNG, and Realizable k- Models Theory

12.4.5

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 Equations 12.4-1, 12.4-4, and 12.4-15), and the corresponding contribution to the production
of  in Equations 12.4-2, 12.4-5, and 12.4-16.
The generation of turbulence due to buoyancy is given by
Gb = gi

t T
Prt xi

(12.4-23)

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/, where
is given by Equation 12.4-9, but with 0 = 1/Pr = k/cp . The coefficient of thermal
expansion, , is defined as
1
=

(12.4-24)

For ideal gases, Equation 12.4-23 reduces to


Gb = gi

t
Prt xi

(12.4-25)

It can be seen from the transport equations for k (Equations 12.4-1, 12.4-4, and 12.4-15)
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 FLUENT, the effects of buoyancy on the generation of k are always included when you have
both a non-zero gravity field and a non-zero temperature (or density) gradient.
While the buoyancy effects on the generation of k are relatively well understood, the
effect on  is less clear. In FLUENT, by default, the buoyancy effects on  are neglected
simply by setting Gb to zero in the transport equation for  (Equation 12.4-2, 12.4-5, or
12.4-16).
However, you can include the buoyancy effects on  in the Viscous Model panel. In this
case, the value of Gb given by Equation 12.4-25 is used in the transport equation for 
(Equation 12.4-2, 12.4-5, or 12.4-16).

c Fluent Inc. September 29, 2006


12-23

Modeling Turbulence

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 [140]:
C3


v
= tanh

(12.4-26)

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.

12.4.6

Effects of Compressibility on Turbulence in the k- Models

For high-Mach-number flows, compressibility affects turbulence through so-called dilatation dissipation, which is normally neglected in the modeling of incompressible
flows [403]. 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 [315]:
YM = 2M2t

(12.4-27)

where Mt is the turbulent Mach number, defined as

Mt =
where a (

k
a2

(12.4-28)

RT ) is the speed of sound.

This compressibility modification always takes effect when the compressible form of the
ideal gas law is used.

12-24

c Fluent Inc. September 29, 2006


12.4 Standard, RNG, and Realizable k- Models Theory

12.4.7

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:

(E) +
[ui (E + p)] =
t
xi
xj

T
keff
+ ui (ij )eff + Sh
xj

(12.4-29)

where E is the total energy, keff is the effective thermal conductivity, and
(ij )eff is the deviatoric stress tensor, defined as
(ij )eff = eff

uj
ui
+
xi xj

2
uk
eff
ij
3
xk

The term involving (ij )eff represents the viscous heating, and is always computed in the
density-based solvers. It is not computed by default in the pressure-based solver, but it
can be enabled in the Viscous Model panel.
Additional terms may appear in the energy equation, depending on the physical models
you are using. See Section 13.2.1: Heat Transfer Theory for more details.
For the standard and realizable k- models, the effective thermal conductivity is given
by
keff = k +

c p t
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 = cp eff
where is calculated from Equation 12.4-9, but with 0 = 1/Pr = k/cp .
The fact that varies with mol /eff , as in Equation 12.4-9, 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 [175]. Equation 12.4-9
works well across a very broad range of molecular Prandtl numbers, from liquid metals
(Pr 102 ) to paraffin oils (Pr 103 ), which allows heat transfer to be calculated in lowReynolds-number regions. Equation 12.4-9 smoothly predicts the variation of effective

c Fluent Inc. September 29, 2006


12-25

Modeling Turbulence

Prandtl number from the molecular value ( = 1/Pr) in the viscosity-dominated region
to the fully turbulent value ( = 1.393) in the fully turbulent regions of the 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 0 in Equation 12.4-9 is 0 = 1/Sc, where Sc is the molecular
Schmidt number.

12.5

Standard and SST k- Models Theory

This section presents the standard [403] and shear-stress transport (SST) [237] k- models. Both models have similar forms, with transport equations for k and . The major
ways in which the SST model [238] differs from the standard model are as follows:
gradual change from the standard k- model in the inner region of the boundary
layer to a high-Reynolds-number version of the k- model in the outer part of the
boundary layer
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.

12.5.1

Standard k- Model

Overview
The standard k- model in FLUENT is based on the Wilcox k- model [403], 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- model called the SST k- model is also available in FLUENT, and is
described in Section 12.5.2: Shear-Stress Transport (SST) k- Model.
The standard k- model is an empirical model based on model transport equations for
the turbulence kinetic energy (k) and the specific dissipation rate (), which can also be
thought of as the ratio of  to k [403].
As the k- model has been modified over the years, production terms have been added
to both the k and equations, which have improved the accuracy of the model for
predicting free shear flows.

12-26

c Fluent Inc. September 29, 2006


12.5 Standard and SST k- Models Theory

Transport Equations for the Standard k- Model


The turbulence kinetic energy, k, and the specific dissipation rate, , are obtained from
the following transport equations:

(k) +
(kui ) =
t
xi
xj

k
k
xj

+ G k Yk + S k

(12.5-1)

() +
(ui ) =
t
xi
xj

xj

+ G Y + S

(12.5-2)

and

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

Modeling the Effective Diffusivity


The effective diffusivities for the k- model are given by
t
k
t
= +

k = +

(12.5-3)

(12.5-4)

where k and are the turbulent Prandtl numbers for k and , respectively. The
turbulent viscosity, t , is computed by combining k and as follows:
t =

c Fluent Inc. September 29, 2006


(12.5-5)

12-27

Modeling Turbulence

Low-Reynolds-Number Correction
The coefficient damps the turbulent viscosity causing a low-Reynolds-number correction. It is given by

0 + Ret /Rk
1 + Ret /Rk

(12.5-6)

where

= 6
i
=
3
= 0.072

Ret =

(12.5-7)

Rk

(12.5-8)

0
i

(12.5-9)
(12.5-10)

Note that, in the high-Reynolds-number form of the k- model, =


= 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
Gk = u0i u0j

uj
xi

(12.5-11)

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


G k = t S 2

(12.5-12)

where S is the modulus of the mean rate-of-strain tensor, defined in the same way as for
the k- model (see Equation 12.4-22).
Production of
The production of is given by

G = Gk
k

12-28

(12.5-13)

c Fluent Inc. September 29, 2006


12.5 Standard and SST k- Models Theory

where Gk is given by Equation 12.5-11.


The coefficient is given by

0 + Ret /R
1 + Ret /R

(12.5-14)

where R = 2.95. and Ret are given by Equations 12.5-6 and 12.5-7, respectively.
Note that, in the high-Reynolds-number form of the k- model, = = 1.

Modeling the Turbulence Dissipation


Dissipation of k
The dissipation of k is given by
Yk = f k

(12.5-15)

where

f =

1+6802k
1+4002k

k 0
k > 0

(12.5-16)

where
k

1 k
3 xj xj

(12.5-17)

and

= i [1 + F (Mt )]
!
4/15 + (Ret /R )4

i =
1 + (Ret /R )4
= 1.5
R = 8

= 0.09

(12.5-18)
(12.5-19)
(12.5-20)
(12.5-21)
(12.5-22)

where Ret is given by Equation 12.5-7.

c Fluent Inc. September 29, 2006


12-29

Modeling Turbulence

Dissipation of
The dissipation of is given by
Y = f 2

(12.5-23)

where

1 + 70
1 + 80


S
ij jk ki
=

( )3

!
1 ui
uj
=

2 xj
xi

f =

(12.5-24)

(12.5-25)

ij

(12.5-26)

The strain rate tensor, Sij is defined in Equation 12.3-11. Also,

= i 1 i F (Mt )
i
"

(12.5-27)

i and F (Mt ) are defined by Equations 12.5-19 and 12.5-28, respectively.


Compressibility Correction
The compressibility function, F (Mt ), is given by
F (Mt ) =

0
Mt Mt0
M2t M2t0 Mt > Mt0

(12.5-28)

where

2k
a2
= 0.25

M2t
Mt0

a =

RT

(12.5-29)
(12.5-30)
(12.5-31)

Note that, in the high-Reynolds-number form of the k- model, i =


. In the incom

pressible form, = i .

12-30

c Fluent Inc. September 29, 2006


12.5 Standard and SST k- Models Theory

Model Constants
1

= 0.09, i = 0.072, R = 8

= 1, = 0.52, 0 = ,
9
Rk = 6, R = 2.95, = 1.5, Mt0 = 0.25, k = 2.0, = 2.0

12.5.2

Shear-Stress Transport (SST) k- Model

Overview
The shear-stress transport (SST) k- model was developed by Menter [237] to effectively
blend the robust and accurate formulation of the k- model in the near-wall region with
the free-stream independence of the k- model in the far field. To achieve this, the k-
model is converted into a k- formulation. The SST k- model is similar to the standard
k- model, but includes the following refinements:
The standard k- model and the transformed k- model are both multiplied by a
blending function and both models are added together. The blending function is
designed to be one in the near-wall region, which activates the standard k- model,
and zero away from the surface, which activates the transformed k- model.
The SST model incorporates a damped cross-diffusion derivative term in the
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- 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- model. Other modifications include the addition of a cross-diffusion term
in the equation and a blending function to ensure that the model equations behave
appropriately in both the near-wall and far-field zones.

c Fluent Inc. September 29, 2006


12-31

Modeling Turbulence

Transport Equations for the SST k- Model


The SST k- model has a similar form to the standard k- model:

(k) +
(kui ) =
t
xi
xj

k
k
xj

k Yk + S k
+G

(12.5-32)

+ G Y + D + S

(12.5-33)

and

() +
(ui ) =
t
xi
xj

xj

k represents the generation of turbulence kinetic energy due to


In these equations, G
mean velocity gradients, calculated as described in Section 12.5.1: Modeling the Turbulence Production. G represents the generation of , calculated as described in Section 12.5.1: Modeling the Turbulence Production. k and represent the effective
diffusivity of k and , respectively, which are calculated as described below. Yk and
Y represent the dissipation of k and due to turbulence, calculated as described in
Section 12.5.1: Modeling the Turbulence Dissipation. D represents the cross-diffusion
term, calculated as described below. Sk and S are user-defined source terms.

Modeling the Effective Diffusivity


The effective diffusivities for the SST k- model are given by
t
k
t
= +

k = +

(12.5-34)

(12.5-35)

where k and are the turbulent Prandtl numbers for k and , respectively. The
turbulent viscosity, t , is computed as follows:
t =

k
1
h
i
max 1 , SF2

(12.5-36)

a1

where S is the strain rate magnitude and

12-32

1
F1 /k,1 + (1 F1 )/k,2
1
=
F1 /,1 + (1 F1 )/,2

k =

(12.5-37)

(12.5-38)

c Fluent Inc. September 29, 2006


12.5 Standard and SST k- Models Theory


is defined in Equation 12.5-6. The blending functions, F1 and F2 , are given by

F1 = tanh 41

"

(12.5-39)

4k
k 500
, 2
,
0.09y y
,2 D+ y 2
#
"
1 1 k
10
= max 2
, 10
,2 xj xj

1 = min max
D+

F2 = tanh 22
"

2 = max 2

k 500
,
0.09y y 2

(12.5-40)
(12.5-41)

(12.5-42)
#

(12.5-43)

where y is the distance to the next surface and D+ is the positive portion of the crossdiffusion term (see Equation 12.5-52).

Modeling the Turbulence Production


Production of k
k represents the production of turbulence kinetic energy, and is defined as:
The term G
k = min(Gk , 10 k)
G
where Gk is defined in the same manner as in the standard k- model.
tion 12.5.1: Modeling the Turbulence Production for details.

(12.5-44)
See Sec-

Production of
The term G represents the production of and is given by
G =

Gk
t

(12.5-45)

Note that this formulation differs from the standard k- model. The difference between
the two models also exists in the way the term is evaluated. In the standard k-
model, is defined as a constant (0.52). For the SST k- model, is given by
= F1 ,1 + (1 F1 ),2

c Fluent Inc. September 29, 2006


(12.5-46)

12-33

Modeling Turbulence

where

,1 =

2
i,1
q

w,1

(12.5-47)

,2 =

2
i,2
q

w,2

(12.5-48)

where is 0.41.

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- model (see Section 12.5.1: Modeling the Turbulence
Dissipation). The difference is in the way the term f is evaluated. In the standard k-
model, f is defined as a piecewise function. For the SST k- model, f is a constant
equal to 1. Thus,
Yk = k

(12.5-49)

Dissipation of
The term Y represents the dissipation of , and is defined in a similar manner as in
the standard k- model (see Section 12.5.1: Modeling the Turbulence Dissipation). The
difference is in the way the terms i and f are evaluated. In the standard k- model, i
is defined as a constant (0.072) and f is defined in Equation 12.5-24. For the SST k-
model, f is a constant equal to 1. Thus,
Yk = 2

(12.5-50)

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


i = F1 i,1 + (1 F1 )i,2

(12.5-51)

and F1 is obtained from Equation 12.5-39.

12-34

c Fluent Inc. September 29, 2006


12.5 Standard and SST k- Models Theory

Cross-Diffusion Modification
The SST k- model is based on both the standard k- model and the standard k- model.
To blend these two models together, the standard k- model has been transformed into
equations based on k and , which leads to the introduction of a cross-diffusion term
(D in Equation 12.5-33). D is defined as
D = 2 (1 F1 ) ,2

1 k
xj xj

(12.5-52)

For details about the various k- models, see Section 12.4: Standard, RNG, and Realizable
k- Models Theory.

Model Constants
k,1 = 1.176, ,1 = 2.0, k,2 = 1.0, ,2 = 1.168

a1 = 0.31, i,1 = 0.075 i,2 = 0.0828

All additional model constants (


, , 0 ,
, R , Rk , R , , and Mt0 ) have the same
values as for the standard k- model (see Section 12.5.1: Model Constants).

12.5.3

Wall Boundary Conditions

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

(u )2 +

(12.5-53)

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


+

= min

c Fluent Inc. September 29, 2006


w+ ,

6
i (y + )2

(12.5-54)

12-35

Modeling Turbulence

where

w+

 2
50

ks+

100
ks+

ks+ < 25
(12.5-55)
ks+

25

where
ks+

ks u
= max 1.0,

(12.5-56)

and ks is the roughness height.


In the logarithmic (or turbulent) region, the value of + is
1 du+
turb
+ = q
+

dy

(12.5-57)

which leads to the value of in the wall cell as


u
=q
y

(12.5-58)

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

12.6

The v 2 -f Model Theory

The v 2 -f model is similar to the standard k- model, but incorporates near-wall turbulence anisotropy and non-local pressure-strain effects. A limitation of the v 2 -f model is
that it cannot be used to solve Eulerian multiphase problems, whereas the k- model is
typically used in such applications. The v 2 -f model is a general low-Reynolds-number
turbulence model that is valid all the way up to solid walls, and therefore does not need
to make use of wall functions. Although the model was originally developed for attached
or mildly separated boundary layers [92], it also accurately simulates flows dominated by
separation [31].
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.

12-36

c Fluent Inc. September 29, 2006


12.7 Reynolds Stress Model (RSM) Theory


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).

12.7

Reynolds Stress Model (RSM) Theory

12.7.1

Overview

The Reynolds stress model (RSM) [119, 193, 194] is the most elaborate turbulence model
that FLUENT 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
transport 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.
The exact form of the Reynolds stress transport equations may be derived by taking moments of the exact momentum equation. This is a process wherein the exact momentum
equations are multiplied by a fluctuating property, the product then being Reynoldsaveraged. Unfortunately, several of the terms in the exact equation are unknown and
modeling assumptions are required in order to close the equations.

c Fluent Inc. September 29, 2006


12-37

Modeling Turbulence

12.7.2

Reynolds Stress Transport Equations

The exact transport equations for the transport of the Reynolds stresses, u0i u0j , may be
written as follows:




( u0i u0j )
+
(uk u0i u0j ) =
u0i u0j u0k + p kj u0i + ik u0j
x
xk
|t {z
}
| k {z
}
|
{z
}
Local Time Derivative Cij Convection
DT,ij Turbulent Diffusion


"

(u0 u0 )
xk
xk i j

{z

DL,ij Molecular Diffusion

u0i u0k

uj
ui
+ u0j u0k
xk
xk

{z

Pij Stress Production

u0i u0j
p
+
xj
xi

{z

ij Pressure Strain
2k u0j u0m ikm + u0i u0m jkm

{z

Fij Production by System Rotation

(gi u0j + gj u0i )


|

{z

Gij Buoyancy Production

u0i u0j
xk xk

{z

ij Dissipation

S
user
| {z
}

(12.7-1)

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 , ij , and ij need to be modeled to close the equations.
The following sections describe the modeling assumptions required to close the equation
set.

12.7.3

Modeling Turbulent Diffusive Transport

DT,ij can be modeled by the generalized gradient-diffusion model of Daly and Harlow [75]:

DT,ij

ku0 u0 u0 u0
= Cs
k ` i j
xk

x`

(12.7-2)

However, this equation can result in numerical instabilities, so it has been simplified in
FLUENT to use a scalar turbulent diffusivity as follows [208]:

DT,ij

=
xk

t u0i u0j
k xk

(12.7-3)

The turbulent viscosity, t , is computed using Equation 12.7-33.

12-38

c Fluent Inc. September 29, 2006


12.7 Reynolds Stress Model (RSM) Theory

Lien and Leschziner [208] derived a value of k = 0.82 by applying the generalized
gradient-diffusion model, Equation 12.7-2, to the case of a planar homogeneous shear
flow. Note that this value of k is different from that in the standard and realizable k-
models, in which k = 1.0.

12.7.4

Modeling the Pressure-Strain Term

Linear Pressure-Strain Model


By default in FLUENT, the pressure-strain term, ij , in Equation 12.7-1 is modeled
according to the proposals by Gibson and Launder [119], Fu et al. [114], and Launder [192,
193].
The classical approach to modeling ij uses the following decomposition:
ij = ij,1 + ij,2 + ij,w

(12.7-4)

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

 0 0 2
C1
u u ij k
k i j 3


(12.7-5)

with C1 = 1.8.
The rapid pressure-strain term, ij,2 , is modeled as
ij,2 C2

2
(Pij + Fij + Gij Cij ) ij (P + G C)
3

(12.7-6)

where C2 = 0.60, Pij , Fij , Gij , and Cij are defined as in Equation 12.7-1, P = 12 Pkk ,
G = 12 Gkk , and C = 12 Ckk .
The wall-reflection term, ij,w , is responsible for the redistribution of normal stresses near
the wall. It tends to damp the normal stress perpendicular to the wall, while enhancing
the stresses parallel to the wall. This term is modeled as

ij,w


3
3
C` k 3/2

u0k u0m nk nm ij u0i u0k nj nk u0j u0k ni nk


k
2
2
d


3
3
C` k 3/2
+ C20 km,2 nk nm ij ik,2 nj nk jk,2 ni nk
2
2
d
C10

(12.7-7)

c Fluent Inc. September 29, 2006


12-39

Modeling Turbulence
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` = C3/4 /, where C = 0.09 and is the von
Karman constant (= 0.4187).
ij,w is included by default in the Reynolds stress model.

Low-Re Modifications to the Linear Pressure-Strain Model


When the RSM is applied to near-wall flows using the enhanced wall treatment described
in Section 12.10.4: Two-Layer Model for Enhanced Wall Treatment, 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 [195]:

C1 = 1 + 2.58A A2 1 exp (0.0067Ret )2

C2 = 0.75 A
2
C10 = C1 + 1.67
3 "
#
2
1
C

2
0
6
C2 = max 3
,0
C2

io

(12.7-8)
(12.7-9)
(12.7-10)
(12.7-11)

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

9
(A2 A3 )
8

A2 aik aki
A3 aik akj aji

(12.7-12)
(12.7-13)
(12.7-14)

aij is the Reynolds-stress anisotropy tensor, defined as


u0i u0j + 23 kij
aij =
k

(12.7-15)

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

12-40

c Fluent Inc. September 29, 2006


12.7 Reynolds Stress Model (RSM) Theory

Quadratic Pressure-Strain Model


An optional pressure-strain model proposed by Speziale, Sarkar, and Gatski [352] 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:

ij = (C1  +

C1 P ) bij

q


1
+ C2  bik bkj bmn bmn ij + C3 C3 bij bij kSij
3


2
+ C4 k bik Sjk + bjk Sik bmn Smn ij + C5 k (bik jk + bjk ik )
3


(12.7-16)

where bij is the Reynolds-stress anisotropy tensor defined as


u0i u0j + 23 kij
bij =
2k

(12.7-17)

The mean strain rate, Sij , is defined as


1
Sij =
2

uj
ui
+
xi xj

(12.7-18)

(12.7-19)

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


1
ij =
2

ui
uj

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 pressurestrain model is not available when the enhanced wall treatment is selected in the Viscous
Model panel.

c Fluent Inc. September 29, 2006


12-41

Modeling Turbulence

Low-Re Stress-Omega Model


The low-Re stress-omega model is a stress-transport model that is based on the omega
equations and LRR model [403]. This model is ideal for modeling flows over curved
surfaces and swirling flows. The low-Re stress-omega model can be selected in the Viscous Model panel and requires no treatments of wall reflections. The closure coefficients
are identical to the k- model (Section 12.5.1: Model Constants), however, there are
additional closure coefficients, C1 and C2 , defined in Section 12.7.4: Model Constants.
The low-Re stress-omega model resembles the k- model due to its excellent predictions
for a wide range of turbulent flows. Furthermore, low Reynolds number modifications
and surface boundary conditions for rough surfaces are similar to the k- model.
Equation 12.7-4 can be re-written for the low-Re stress-omega model such that wall
reflections are excluded:
ij = ij,1 + ij,2

(12.7-20)

Hence,

ij = (C1  +

C1 P ) bij

q


1
+ C2  bik bkj bmn bmn ij + C3 C3 bij bij kSij
3


2
+ C4 k bik Sjk + bjk Sik bmn Smn ij + C5 k (bik jk + bjk ik )
3


(12.7-21)

where bij is the Reynolds-stress anisotropy tensor defined as


u0i u0j + 23 kij
bij =
2k

(12.7-22)

The mean strain rate, Sij , is defined in Equation 12.7-18 and the mean rate-of-rotation
tensor, ij , is defined by Equation 12.7-19.
The constants are
C1 = 3.4, C1 = 1.8, C2 = 4.2, C3 = 0.8, C3 = 1.3, C4 = 1.25, C5 = 0.4
Near-wall treatment options in the Viscous Model panel are not available with the low-Re
stress-omega model.

12-42

c Fluent Inc. September 29, 2006


12.7 Reynolds Stress Model (RSM) Theory

Model Constants
C1 = 1.8, C2 = 0.52
1

= 0.09, i = 0.072, R = 8

= 1, = 0.52, 0 = ,
9
Rk = 6, R = 2.95, = 1.5, Mt0 = 0.25, k = 2.0, = 2.0
Wall Boundary Conditions
The wall boundary conditions for the low-Re stress-omega equation in the RSM models
are treated in the same way as the k equation in the k- models.
FLUENT defines the value of at the wall as
(u )2 +

w =

(12.7-23)

where + is dimensionless and is defined as

w+ =

 2
50

ks+

500
ks+

ks+ < 25
(12.7-24)
ks+

25

where
ks+ =

ks u

(12.7-25)

and ks is the roughness height.

c Fluent Inc. September 29, 2006


12-43

Modeling Turbulence

12.7.5

Effects of Buoyancy on Turbulence

The production terms due to buoyancy are modeled as


t
T
T
Gij =
gi
+ gj
Prt
xj
xi

(12.7-26)

where Prt is the turbulent Prandtl number for energy, with a default value of 0.85.
Using the definition of the coefficient of thermal expansion, , given by Equation 12.4-24,
the following expression is obtained for Gij for ideal gases:
t

Gij =
gi
+ gj
Prt
xj
xi

12.7.6

(12.7-27)

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
2

(12.7-28)

As described in Section 12.7.9: Wall Boundary Conditions, 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:

(k) +
(kui ) =
t
xi
xj

"

t
+
k

k
1
+ (Pii + Gii ) (1 + 2M2t ) + Sk (12.7-29)
xj
2

where k = 0.82 and Sk is a user-defined source term. Equation 12.7-29 is obtainable


by contracting the modeled equation for the Reynolds stresses (Equation 12.7-1). As
one might expect, it is essentially identical to Equation 12.4-1 used in the standard k-
model.
Although Equation 12.7-29 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 12.7-28. This is a minor point, however, since the values of k obtained with
either method should be very similar.

12-44

c Fluent Inc. September 29, 2006


12.7 Reynolds Stress Model (RSM) Theory

12.7.7

Modeling the Dissipation Rate

The dissipation tensor, ij , is modeled as


2
ij = ij ( + YM )
3

(12.7-30)

where YM = 2M2t is an additional dilatation dissipation term according to the model


by Sarkar [315]. The turbulent Mach number in this term is defined as

Mt =

k
a2

(12.7-31)

where a ( RT ) 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:

() +
(ui ) =
t
xi
xj

"

t
+



1

2
C1 [Pii + C3 Gii ] C2 + S (12.7-32)
xj
2
k
k
#

where  = 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 12.4.5: Effects of
Buoyancy on Turbulence in the k- Models, and S is a user-defined source term.

12.7.8

Modeling the Turbulent Viscosity

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


t = C

k2


(12.7-33)

where C = 0.09.

c Fluent Inc. September 29, 2006


12-45

Modeling Turbulence

12.7.9

Wall Boundary Conditions

The RSM model in FLUENT requires boundary conditions for individual Reynolds stresses,
u0i u0j , and for the turbulence dissipation rate,  (or if the low-Re stress-omega model
is used). These quantities can be input directly or derived from the turbulence intensity
and characteristic length, as described in Section 12.20.3: Reynolds Stress Model.
At walls, FLUENT computes the near-wall values of the Reynolds stresses and  from
wall functions (see Section 12.10.2: Standard Wall Functions, Section 12.10.3: NonEquilibrium Wall Functions, and Section 12.10.4: Enhanced Wall Functions). 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 12.7-1). Using a local coordinate system, where
is the tangential coordinate, is the normal coordinate, and is the binormal coordinate, the Reynolds stresses at the wall-adjacent cells (assuming standard wall functions
or non-equilibrium wall functions) are computed from
0
u0 2
u0 u0
u02
u2
= 1.098, = 0.247, = 0.655, = 0.255
k
k
k
k

(12.7-34)

To obtain k, FLUENT solves the transport equation of Equation 12.7-29. 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
normal Reynolds stresses using Equation 12.7-28. By default, the values of the Reynolds
stresses near the wall are fixed using the values computed from Equation 12.7-34, and
the transport equations in Equation 12.7-1 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:
0
u02
u0 u0
u02
u2
=
5.1,
=
1.0,
=
2.3,

= 1.0
u2
u2
u2
u2

(12.7-35)

where u is the friction velocity defined by u w /, where w is the wall-shear stress.


When this option is chosen, the k transport equation is not solved.
When using enhanced wall treatments as the near-wall treatment, FLUENT applies zero
flux wall boundary conditions to the Reynolds stress equations.

12-46

c Fluent Inc. September 29, 2006


12.8 Detached Eddy Simulation (DES) Model Theory

12.7.10

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:

(E) +
[ui (E + p)] =
t
xi
xj

"

c p t
k+
Prt

T
+ ui (ij )eff + Sh
xj

(12.7-36)

where E is the total energy and (ij )eff is the deviatoric stress tensor, defined as
(ij )eff = eff

uj
ui
+
xi xj

2
uk
eff
ij
3
xk

The term involving (ij )eff represents the viscous heating, and is always computed in the
density-based solvers. It is not computed by default in the pressure-based solver, but it
can be enabled in the Viscous Model 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.

12.8

Detached Eddy Simulation (DES) Model Theory


Overview

FLUENT offers three different models for the detached eddy simulation: the SpalartAllmaras model, the realizable k- model, and the SST k- model.
In the DES approach, the unsteady RANS models are employed in the near-wall regions,
while the filtered versions of the same models are used in the regions away from the
near-wall. The LES region is normally associated with the core turbulent region where
large turbulence scales play a dominant role. In this region, the DES models recover
the respective subgrid models. In the near-wall region, the respective RANS models are
recovered.
The application of DES, however, may still require significant CPU resources and therefore, as a general guideline, it is recommended that the conventional turbulence models
employing the Reynolds-averaged approach be used for practical calculations.
The DES models, often referred to as the hybrid LES/RANS models combine RANS
modeling with LES for applications such as high-Re external aerodynamics simulations.
In FLUENT, the DES model is based on the one-equation Spalart-Allmaras model, the
realizable k- model, and the SST k- model. The computational costs, when using the
DES models, is less than LES computational costs, but greater than RANS.

c Fluent Inc. September 29, 2006


12-47

Modeling Turbulence

12.8.1

Spalart-Allmaras RANS Model

The standard Spalart-Allmaras model uses the distance to the closest wall as the definition for the length scale d, which plays a major role in determining the level of production
and destruction of turbulent viscosity (Equations 12.3-6, 12.3-12, and 12.3-15). The DES

model, as proposed by Shur et al. [331] replaces d everywhere with a new length scale d,
defined as
d = min(d, Cdes )

(12.8-1)

where the grid spacing, , is based on the largest grid space in the x, y, or z directions
forming the computational cell. The empirical constant Cdes has a value of 0.65.

12.8.2

Realizable k- RANS Model

This RANS model is similar to the Realizable k- model discussed in Section 12.4.3: Realizable k- Model, with the exception of the dissipation term in the k equation. In the
DES model, the Realizable k- RANS dissipation term is modified such that:
3

k 2
Yk =
ldes

(12.8-2)

where
ldes = min(lrke , lles )
k

= Cdes

lrke =
lles

3
2

(12.8-3)
(12.8-4)
(12.8-5)

where Cdes is a calibration constant used in the DES model and has a value of 0.61 and
is the maximum local grid spacing (x, y, z).
For the case where ldes = lrke , you will obtain an expression for the dissipation of the k
formulation for the Realizable k- model (Section 12.4.3: Realizable k- Model):
Yk = 

12-48

c Fluent Inc. September 29, 2006


12.9 Large Eddy Simulation (LES) Model Theory

12.8.3

SST k- RANS Model

The dissipation term of the turbulent kinetic energy (see Section 12.5.1: Modeling the Turbulence Dissipation) is modified for the DES turbulence model as described in Menters
work [238] such that
Yk = kf

(12.8-6)

where f is no longer a constant equal to 1 as in the SST k- model (see Section 12.5.1: Modeling the Turbulence Dissipation), but is now expressed as
f = max

Lt
Cdes

,1

(12.8-7)

where Cdes is a calibration constant used in the DES model and has a value of 0.61, is
the maximum local grid spacing (x, y, z) and f is defined in Equation 12.5-16.
The turbulent length scale is the parameter that defines this RANS model:

Lt =

12.9

(12.8-8)

Large Eddy Simulation (LES) Model Theory

12.9.1

Overview

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 possible, in theory, to directly resolve the whole spectrum of turbulent scales using
an approach known as direct numerical simulation (DNS). No modeling is required in
DNS. However, DNS is not feasible for practical engineering problems involving high
Reynolds number flows. The cost required for DNS to resolve the entire range of scales
is proportional to Re3t , where Ret is the turbulent Reynolds number. Clearly, for high
Reynolds numbers, the cost becomes prohibitive.

c Fluent Inc. September 29, 2006


12-49

Modeling Turbulence

In LES, large eddies are resolved directly, while small eddies are modeled. Large eddy
simulation (LES) thus falls between DNS and RANS in terms of the fraction of the
resolved scales. The rationale behind LES can be summarized as follows:
Momentum, mass, energy, and other passive scalars are transported mostly by large
eddies.
Large eddies are more problem-dependent. They are dictated by the geometries
and boundary conditions of the 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 turbulence model is much higher for small eddies.
Resolving only the large eddies allows one to use much coarser mesh and larger timesstep sizes in LES than in DNS. However, LES still requires substantially finer meshes
than those typically used for RANS calculations. In addition, LES has to be run for
a sufficiently long flow-time to obtain stable statistics of the flow being modeled. As
a result, the computational cost involved with LES is normally orders of magnitudes
higher than that for steady RANS calculations in terms of memory (RAM) and CPU
time. Therefore, high-performance computing (e.g., parallel computing) is a necessity for
LES, especially for industrial applications.
The following sections give details of the governing equations for LES, the subgrid-scale
turbulence models, and the boundary conditions.

12.9.2

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
(x) =

(x0 )G(x, x0 )dx0

(12.9-1)

where D is the fluid domain, and G is the filter function that determines the scale of the
resolved eddies.

12-50

c Fluent Inc. September 29, 2006


12.9 Large Eddy Simulation (LES) Model Theory

In FLUENT, the finite-volume discretization itself implicitly provides the filtering operation:
1 Z
(x) =
(x0 ) dx0 , x0 V
V V

(12.9-2)

where V is the volume of a computational cell. The filter function, G(x, x0 ), implied here
is then
0

G(x, x )

1/V, x0 V
0,
x0 otherwise

(12.9-3)

The LES capability in FLUENT is applicable to compressible flows. For the sake of concise
notation, however, the theory is presented here for incompressible flows.
Filtering the Navier-Stokes equations, one obtains

+
(ui ) = 0
t xi

(12.9-4)

and

(ui ) +
(ui uj ) =
t
xj
xj

ij

xj

p
ij

xi
xj

(12.9-5)

where ij is the stress tensor due to molecular viscosity defined by


"

ui uj
ij
+
xj
xi

!#

2 ul

ij
3 xl

(12.9-6)

and ij is the subgrid-scale stress defined by


ij ui uj ui uj

c Fluent Inc. September 29, 2006


(12.9-7)

12-51

Modeling Turbulence

12.9.3

Subgrid-Scale Models

The subgrid-scale stresses resulting from the filtering operation are unknown, and require modeling. The subgrid-scale turbulence models in FLUENT employ the Boussinesq
hypothesis [142] as in the RANS models, computing subgrid-scale turbulent stresses from
1
ij kk ij = 2t S ij
3

(12.9-8)

where t is the subgrid-scale turbulent viscosity. The isotropic part of the subgrid-scale
stresses kk is not modeled, but added to the filtered static pressure term. S ij is the
rate-of-strain tensor for the resolved scale defined by
1
S ij
2

ui uj
+
xj
xi

(12.9-9)

For compressible flows, it is convenient to introduce the density-weighted (or Favre)


filtering operator:

(12.9-10)

The Favre Filtered Navier-Stokes equation takes the same form as Equation 12.9-5. The
compressible form of the subgrid stress tensor is defined as:
Tij = ui uj + ui uj

(12.9-11)

This term is split into its isotropic and deviatoric parts


1
1
Tij = Tij Tll ij + Tll ij
3
|
{z
}
|3 {z }
isotropic
deviatoric

(12.9-12)

The deviatoric part of the subgrid-scale stress tensor is modeled using the compressible
form of the Smagorinsky model:
1
1
Tij Tll ij = 2t (ij ii ij )
3
3

(12.9-13)

As for incompressible flows, the term involving Tll can be added to the filtered pressure
or simply neglected [99]. Indeed, this term can be re-written as Tll = M 2 sgs p where
Msgs is the subgrid Mach number. This subgrid Mach number can be expected to be
small when the turbulent Mach number of the flow is small.

12-52

c Fluent Inc. September 29, 2006


12.9 Large Eddy Simulation (LES) Model Theory

FLUENT offers four models for t : the Smagorinsky-Lilly model, the dynamic SmagorinskyLilly model, the WALE model, and the dynamic kinetic energy subgrid-scale model.
Subgrid-scale turbulent flux of a scalar, , is modeled using s subgrid-scale turbulent
Prandtl number by
qj =

t
t xj

(12.9-14)

where qj is the subgrid-scale flux.


In the dynamic models, the subgrid-scale turbulent Prandtl number or Schmidt number
is obtained by applying the dynamic procedure originally proposed by Germano [116] to
the subgrid-scale flux.

Smagorinsky-Lilly Model
This simple model was first proposed by Smagorinsky [337]. In the Smagorinsky-Lilly
model, the eddy-viscosity is modeled by

t = L2s S

(12.9-15)

where Ls is the mixing length for subgrid scales and S


computed using


Ls = min d, Cs V 1/3

2S ij S ij . In FLUENT, Ls is

(12.9-16)

where is the von Karman constant, d is the distance to the closest wall, Cs is the
Smagorinsky constant, and V is the volume of the computational cell.
Lilly derived a value of 0.17 for Cs for homogeneous isotropic turbulence in the inertial
subrange. However, this value was found to cause excessive damping of large-scale fluctuations in the presence of mean shear and in transitional flows as near solid boundary,
and has to be reduced in such regions. In short, Cs is not an universal constant, which
is the most serious shortcoming of this simple model. Nonetheless, Cs value of around
0.1 has been found to yield the best results for a wide range of flows, and is the default
value in FLUENT.

c Fluent Inc. September 29, 2006


12-53

Modeling Turbulence

Dynamic Smagorinsky-Lilly Model


Germano et al. [116] and subsequently Lilly [211] conceived a procedure in which the
Smagorinsky model constant, Cs , is dynamically computed based on the information
provided by the resolved scales of motion. The dynamic procedure thus obviates the
need for users to specify the model constant Cs in advance. The details of the model
implementation in FLUENT and its validation can be found in [181].
The Cs obtained using the dynamic Smagorinsky-Lilly model varies in time and space
over a fairly wide range. To avoid numerical instability, in FLUENT, Cs is clipped at zero
and 0.23 by default.

Wall-Adapting Local Eddy-Viscosity (WALE) Model


In the WALE model [262], the eddy viscosity is modeled by
t = L2s

(Sijd Sijd )3/2


(S ij S ij )5/2 + (Sijd Sijd )5/4

(12.9-17)

where Ls and Sijd in the WALE model are defined, respectively, as




Ls = min d, Cw V 1/3

Sijd =


1 2
1
ui
g ij + g 2ji ij g 2kk , g ij =
xj
2
3

(12.9-18)

(12.9-19)

In FLUENT, the default value of the WALE constant, Cw , is 0.325 and has been found
to yield satisfactory results for a wide range of flow. The rest of the notation is the
same as for the Smagorinsky-Lilly model. With this spatial operator, the WALE model
is designed to return the correct wall asymptotic (y 3 ) behavior for wall bounded flows.

12-54

c Fluent Inc. September 29, 2006


12.9 Large Eddy Simulation (LES) Model Theory

Dynamic Kinetic Energy Subgrid-Scale Model


The original and dynamic Smagorinsky-Lilly models, discussed previously, are essentially
algebraic models in which subgrid-scale stresses are parameterized using the resolved velocity scales. The underlying assumption is the local equilibrium between the transferred
energy through the grid-filter scale and the dissipation of kinetic energy at small subgrid scales. The subgrid-scale turbulence can be better modeled by accounting for the
transport of the subgrid-scale turbulence kinetic energy.
The dynamic subgrid-scale kinetic energy model in FLUENT replicates the model proposed by Kim and Menon [184].
The subgrid-scale kinetic energy is defined as
ksgs =


1 2
uk u2k
2

(12.9-20)

which is obtained by contracting the subgrid-scale stress in Equation 12.9-7.


The subgrid-scale eddy viscosity, t , is computed using ksgs as
1/2
t = Ck ksgs
f

(12.9-21)

where f is the filter-size computed from f V 1/3 .


The subgrid-scale stress can then be written as
2
1/2
ij ksgs ij = 2Ck ksgs
f S ij
3

(12.9-22)

ksgs is obtained by solving its transport equation


3/2
ksgs
k sgs uj k sgs
ui

+
= ij
C
+
t
xj
xj
f
xj

t ksgs
k xj

(12.9-23)

In the above equations, the model constants, Ck and C , are determined dynamically [184].
k is hardwired to 1.0. The details of the implementation of this model in FLUENT and
its validation is given by Kim [181].

c Fluent Inc. September 29, 2006


12-55

Modeling Turbulence

12.9.4

Inlet Boundary Conditions for the LES Model

This section describes the three algorithms available in FLUENT to model the fluctuating
velocity at velocity inlet boundaries.

No Perturbations
The stochastic components of the flow at the velocity-specified inlet boundaries are neglected if the No Perturbations option is used. In such cases, individual instantaneous
velocity components are simply set equal to their mean velocity counterparts. This option is suitable only when the level of turbulence at the inflow boundaries is negligible or
does not play a major role in the accuracy of the overall solution.

Vortex Method
To generate a time-dependent inlet condition, a random 2D vortex method is considered.
With this approach, a perturbation is added on a specified mean velocity profile via a
fluctuating vorticity field (i.e. two-dimensional in the plane normal to the streamwise
direction). The vortex method is based on the Lagrangian form of the 2D evolution
equation of the vorticity and the Biot-Savart law. A particle discretization is used to
solve this equation. These particles, or vortex points are convected randomly and
carry information about the vorticity field. If N is the number of vortex points and A
is the area of the inlet section, the amount of vorticity carried by a given particle i is
represented by the circulation i and an assumed spatial distribution :

v
u
u
i (x, y) = 4t

(~x) =

Ak(x, y)
3N [2 ln(3) 3 ln(2)]

(12.9-24)


1  |x|2 /22
2
2
2e

1
2e|x| /2
2
2

(12.9-25)

where k is the turbulence kinetic energy. The parameter provides control over the size
of a vortex particle. The resulting discretization for the velocity field is given by
0 2
2
N
1 X
((~xi ~x) ~z)(1 e|~x~x | /2 )
~u(~x) =
i
2 i=1
|~x ~x0i |2

12-56

(12.9-26)

c Fluent Inc. September 29, 2006


12.9 Large Eddy Simulation (LES) Model Theory

Where ~z is the unit vector in the streamwise direction. Originally [327], the size of
the vortex was fixed by an ad hoc value of . To make the vortex method generally
applicable, a local vortex size is specified through a turbulent mixing length hypothesis.
is calculated from a known profile of mean turbulence kinetic energy and mean dissipation
rate at the inlet according to the following:

ck 3/2
2

(12.9-27)

where c = 0.16. To ensure that the vortex will always belong to resolved scales, the
minimum value of in Equation 12.9-27 is bounded by the local grid size. The sign
of the circulation of each vortex is changed randomly each characteristic time scale .
In the general implementation of the vortex method, this time scale represents the time
necessary for a 2D vortex convected by the bulk velocity in the boundary normal direction
to travel along n times its mean characteristic 2D size (m ), where n is fixed equal to
100 from numerical testing. The vortex method considers only velocity fluctuations in
the plane normal to the streamwise direction.
In FLUENT however, a simplified linear kinematic model (LKM) for the streamwise
velocity fluctuations is used [231]. It is derived from a linear model that mimics the
influence of the two-dimensional vortex in the streamwise mean velocity field. If the
mean streamwise velocity U is considered as a passive scalar, the fluctuation u0 resulting
from the transport of U by the planar fluctuating velocity field v 0 is modeled by
u0 = ~v 0 ~g

(12.9-28)

~ . When this mean


where ~g is the unit vector aligned with the mean velocity gradient U
velocity gradient is equal to zero, a random perturbation can be considered instead.

Since the vortex method theory is based on the modification of the velocity
field normal to the streamwise direction, it is imperative that the user
creates an inlet plane normal (or as close as possible) to the streamwise
velocity direction.

Spectral Synthesizer
The spectral synthesizer provides an alternative method of generating fluctuating velocity
components. It is based on the random flow generation technique originally proposed by
Kraichnan [186] and modified by Smirnov et al. [338]. In this method, fluctuating velocity
components are computed by synthesizing a divergence-free velocity-vector field from
the summation of Fourier harmonics. In the implementation in FLUENT, the number of
Fourier harmonics is fixed to 100.

c Fluent Inc. September 29, 2006


12-57

Modeling Turbulence

12.10 Near-Wall Treatments for Wall-Bounded Turbulent Flows


12.10.1 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 nearwall 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 nearwall 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
needs to be given as to how to make these models suitable for wall-bounded flows. The
Spalart-Allmaras and k- 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 12.10.1 illustrates these subdivisions of the near-wall region, plotted
in semi-log coordinates.

12-58

c Fluent Inc. September 29, 2006


12.10 Near-Wall Treatments for Wall-Bounded Turbulent Flows

Figure 12.10.1: Subdivisions of the Near-Wall Region

In Figure 12.10.1, y + u y/, where u is the friction velocity, defined as

w
.

Wall Functions vs. Near-Wall Model


Traditionally, there are two approaches to modeling the near-wall region. In one approach, the viscosity-affected inner region (viscous sublayer and buffer layer) is not resolved. 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 12.10.2.
In most high-Reynolds-number flows, the wall function approach substantially saves computational 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-Reynoldsnumber effects are pervasive in the flow domain in question, and the hypotheses underlying the wall functions cease to be valid. Such situations require near-wall models that

c Fluent Inc. September 29, 2006


12-59

Modeling Turbulence

Figure 12.10.2: Near-Wall Treatments in FLUENT

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.

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
laws-of-the-wall for mean velocity and temperature (or other scalars)
formulas for near-wall turbulent quantities
Depending on the turbulent model you choose, FLUENT offers three to four choices of
wall function approaches:
Standard Wall Functions
Non-Equilibrium Wall Functions
Enhanced Wall Treatment
User-Defined Wall Functions

12-60

c Fluent Inc. September 29, 2006


12.10 Near-Wall Treatments for Wall-Bounded Turbulent Flows

12.10.2

Standard Wall Functions

The standard wall functions in FLUENT are based on the proposal of Launder and Spalding [197], 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
U =

1
ln(Ey )

(12.10-1)

where
1/2

UP C1/4 kP
U
w /

(12.10-2)

1/2

C1/4 kP yP
y

and
E
UP
kP
yP

=
=
=
=
=
=

(12.10-3)

von Karman constant (= 0.4187)


empirical constant (= 9.793)
mean velocity of the fluid at point P
turbulence kinetic energy at point P
distance from point P to the wall
dynamic viscosity of the fluid

The logarithmic law for mean velocity is known to be valid for 30 < y < 300. In
FLUENT, the log-law is employed when y > 11.225.
When the mesh is such that y < 11.225 at the wall-adjacent cells, FLUENT applies the
laminar stress-strain relationship that can be written as
U = y

(12.10-4)

It should be noted that, in FLUENT, the laws-of-the-wall for mean velocity and temperature are based on the wall unit, y , rather than y + ( u y/). These quantities are
approximately equal in equilibrium turbulent boundary layers.

c Fluent Inc. September 29, 2006


12-61

Modeling Turbulence

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 conduction
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 [382].
The law-of-the-wall implemented in FLUENT has the following composite form:

1/2

(Tw TP ) cp C1/4 kP
T
q

1/4 1/2
k
C

Pr y + 12 Pr q P UP2

h
i

Pr 1 ln(Ey ) + P +
t
=
1/4 1/2

1 C kP

{Prt UP2 + (Pr Prt )Uc2 }

2
q

(y < yT )
(y > yT )
(12.10-5)

where P is computed by using the formula given by Jayatilleke [165]:

P = 9.24

12-62

"

Pr
Prt

3/4

1 + 0.28e0.007Pr/Prt

(12.10-6)

c Fluent Inc. September 29, 2006


12.10 Near-Wall Treatments for Wall-Bounded Turbulent Flows

and
kP

cp
q
TP
Tw
Pr
Prt
A
Uc

=
=
=
=
=
=
=
=
=
=

turbulent kinetic energy at point P


density of fluid
specific heat of fluid
wall heat flux
temperature at the cell adjacent to wall
temperature at the wall
molecular Prandtl number (cp /kf )
turbulent Prandtl number (0.85 at the wall)
Van Driest constant (= 26)
mean velocity magnitude at y = yT

Note that, for the pressure-based solver, the terms


1/2

C1/4 kP
1
Pr
UP2
2
q
and

1/2
o
1 C1/4 kP n

Prt UP2 + (Pr Prt )Uc2


2
q

will be included in Equation 12.10-5 only for compressible flow calculations.


The non-dimensional thermal sublayer thickness, yT , in Equation 12.10-5 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 12.10-5 is applied to compute the wall temperature
Tw or heat flux q (depending on the type of the thermal boundary conditions).
The function for P given by equation Equation 12.10-6 is relevant for the smooth walls.
For the rough walls, however, this function is modified as follows:
0.695

Prough = 3.15Pr

1
1

E0 E

0.359

E0
+
E

!0.6

(12.10-7)

where E 0 is the wall function constant modified for the rough walls, defined by E 0 = E/fr .
To find a description of the roughness function fr , you may refer to Equation 7.13-3 in
Section 7.13.1: Wall Roughness Effects in Turbulent Wall-Bounded Flows.

c Fluent Inc. September 29, 2006


12-63

Modeling Turbulence

Species
When using wall functions for species transport, FLUENT assumes that species transport
behaves analogously to heat transfer. Similarly to Equation 12.10-5, the law-of-the-wall
for species can be expressed for constant property flow with no viscous dissipation as
1/2

(Yi,w Yi ) C1/4 kP

Y
Ji,w

Sc yh
i
Sct 1 ln(Ey ) + Pc

(y < yc )
(y > yc )

(12.10-8)

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
n

(12.10-9)

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
G k w

U
w
= w
1/4
1/2
y
C kP yP

(12.10-10)

and  is computed from


3/2

P =

C3/4 kP
yP

(12.10-11)

The  equation is not solved at the wall-adjacent cells, but instead is computed using Equation 12.10-11. and Reynolds stress equations are solved as detailed in Sections 12.5.3 and 12.7.9, respectively.

12-64

c Fluent Inc. September 29, 2006


12.10 Near-Wall Treatments for Wall-Bounded Turbulent Flows

Note that, as shown here, the wall boundary conditions for the solution variables, including 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 constantshear 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
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.

Standard wall functions are available with the following viscous models:
K-epsilon
Reynolds Stress

12.10.3

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 [182] is also available. The
key elements in the non-equilibrium wall functions are as follows:
Launder and Spaldings log-law for mean velocity is sensitized to pressure-gradient
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
!
U C1/4 k 1/2
C1/4 k 1/2 y
1
= ln E
w /

c Fluent Inc. September 29, 2006


(12.10-12)

12-65

Modeling Turbulence

where
"

y
1 dp
yv
ln
U = U
2 dx k
yv

y yv y 2
+ + v

(12.10-13)

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


yv

yv

(12.10-14)

1/4 1/2

C kP

where yv = 11.225.
The non-equilibrium wall function employs the two-layer concept in computing the budget 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:

t =

y
0, y < yv
k = yv
w , y > yv
kP ,

2

kP , y < yv
=
y > yv

2k
,
y2
k3/2
,
C` y

y < yv
y > yv

(12.10-15)

where C` = C3/4 , and yv is the dimensional thickness of the viscous sublayer, defined
in Equation 12.10-14.
Using these profiles, the cell-averaged production of k, Gk , and the cell-averaged dissipation 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 approximated with a depth-average,
!
1 Z yn U
1
w2
yn
Gk
t
dy =
ln
yn 0
y
yn C1/4 kP1/2
yv

(12.10-16)

and

!
1/2
1 Z yn
1 2 kP
yn

 dy =
+ ln
kP
yn 0
yn yv
C`
yv

(12.10-17)

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 12.10-16 and 12.10-17, the turbulence kinetic energy budget for the wallneighboring 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

12-66

c Fluent Inc. September 29, 2006


12.10 Near-Wall Treatments for Wall-Bounded Turbulent Flows

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.

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
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, the k- models,
and the RSM.

c Fluent Inc. September 29, 2006


12-67

Modeling Turbulence

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).

Non-equilibrium wall functions are available with the following viscous


models:
K-epsilon
Reynolds Stress

12.10.4

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.

12-68

c Fluent Inc. September 29, 2006


12.10 Near-Wall Treatments for Wall-Bounded Turbulent Flows

Two-Layer Model for Enhanced Wall Treatment


In FLUENTs near-wall model, the viscosity-affected near-wall region is completely resolved 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 viscosity 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 regions is determined by a wall-distance-based, turbulent Reynolds number, Rey , defined
as

y k
Rey

(12.10-18)

where y is the normal distance from the wall at the cell centers. In FLUENT, y is
interpreted as the distance to the nearest wall:
y min k~r ~rw k

(12.10-19)

~
rw w

where ~r is the position vector at the field point, and ~rw is the position vector on the
wall boundary. w is the union of all the wall boundaries involved. This interpretation
allows y to be uniquely 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 > Rey ; Rey = 200), the k- models or the RSM
(described in Sections 12.4 and 12.7) are employed.
In the viscosity-affected near-wall region (Rey < Rey ), the one-equation model of Wolfstein [406] is employed. In the one-equation model, the momentum equations and the
k equation are retained as described in Sections 12.4 and 12.7. However, the turbulent
viscosity, t , is computed from

t,2layer = C ` k

(12.10-20)

where the length scale that appears in Equation 12.10-20 is computed from [54]


` = yC` 1 eRey /A

(12.10-21)

The two-layer formulation for turbulent viscosity described above is used as a part of the
enhanced wall treatment, in which the two-layer definition is smoothly blended with the
high-Reynolds-number t definition from the outer region, as proposed by Jongen [168]:
t,enh =  t + (1  )t,2layer

c Fluent Inc. September 29, 2006


(12.10-22)

12-69

Modeling Turbulence

where t is the high-Reynolds-number definition as described in Section 12.4: Standard,


RNG, and Realizable k- Models Theory or 12.7 for the k- models or the RSM. A
blending function,  , 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 Rey
1
 =
1 + tanh
2
A
"

!#

(12.10-23)

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

|Rey |
tanh(0.98)

(12.10-24)

Typically, Rey would be assigned a value that is between 5% and 20% of Rey . The
main purpose of the blending function  is to prevent solution convergence from being
impeded when the k- solution in the outer layer does not match with the two-layer
formulation.
The  field is computed from

=

k 3/2
`

(12.10-25)

The length scales that appear in Equation 12.10-25 are again computed from Chen and
Patel [54]:


` = yC` 1 eRey /A

(12.10-26)

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 12.10-25. FLUENT uses a procedure for the  specification that is similar to
the t 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 12.10-21 and 12.10-26, are taken
from [54]:
C` = C3/4 ,

12-70

A = 70,

A = 2C`

(12.10-27)

c Fluent Inc. September 29, 2006


12.10 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 [170]:
1

u+ = e u+
lam + e uturb

(12.10-28)

where the blending function is given by:


=

a(y + )4
1 + by +

(12.10-29)

where a = 0.01 and b = 5.


Similarly, the general equation for the derivative

du+
dy +

is

+
+
1 du
du+
turb
dulam

=
e
+
e
dy +
dy +
dy +

(12.10-30)

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 compressible flow with heat transfer and pressure gradients has been derived by combining the
approaches of White and Cristoph [402] and Huang et al. [149]:
i
du+
1 h 0
turb
+
+ 2 1/2
=
S
(1

(u
)
)
dy +
y +

(12.10-31)

where
S0 =

c Fluent Inc. September 29, 2006


1 + y + for y + < ys+


1 + ys+ for y + ys+

(12.10-32)

12-71

Modeling Turbulence

and

w dp
dp
= 2 3

w u dx
(u ) dx

t qw
t qw u

=
cp w Tw
cp u Tw
t (u )2

2cp Tw

(12.10-33)
(12.10-34)
(12.10-35)

where ys+ is the location at which the log-law slope will remain fixed. By default, ys+ = 60.
The coefficient in Equation 12.10-31 represents the influences of pressure gradients
while the coefficients and represent thermal effects. Equation 12.10-31 is an ordinary
differential equation and FLUENT will provide an appropriate analytical solution. If , ,
and all equal 0, an analytical solution would lead to the classical turbulent logarithmic
law-of-the-wall.
The laminar law-of-the-wall is determined from the following expression:
du+
lam
= 1 + y +
+
dy

(12.10-36)

Note that the above expression only includes effects of pressure gradients through ,
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 12.10-36
results in
u+
lam

=y

1 + y+
2

(12.10-37)

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 [170]:

1
(Tw TP ) cp u
+
+

= e Tlam
+ e Tturb
q

(12.10-38)

where the notation for TP and q is the same as for standard thermal wall functions (see
Equation 12.10-5). Furthermore, the blending factor is defined as
=

12-72

a(Pr y + )4
1 + bPr3 y +

(12.10-39)

c Fluent Inc. September 29, 2006


12.10 Near-Wall Treatments for Wall-Bounded Turbulent Flows

where Pr is the molecular Prandtl number, and the coefficients a and b are defined as in
Equation 12.10-29.
Apart from the formulation for T + in Equation 12.10-38, enhanced thermal wall functions
follow the same logic as for standard thermal wall functions (see Section 12.10.2: Energy),
resulting in the following definition for turbulent and laminar thermal wall functions:
+
Tlam

+
Tturb

= Prt

u+
turb

= Pr

u+
lam

u 2
+
u
2q

Pr
u 2
2
2
+P +
u
1 (u+
c ) (u)
2q
Prt


(12.10-40)
)

(12.10-41)

+
where the quantity u+
c is the value of u at the fictitious crossover between the laminar
and turbulent region. The function P is defined in the same way as for standard wall
functions.

A similar procedure is also used for species wall functions when the enhanced wall treatment is used. In this case, the Prandtl numbers in Equations 12.10-40 and 12.10-41 are
replaced by adequate Schmidt numbers. See Section 12.10.2: Species for details about
species wall functions.
The boundary condition for turbulence kinetic energy is the same as for standard wall
functions (Equation 12.10-9). However, the production of turbulence kinetic energy Gk is
computed using the velocity gradients that are consistent with the enhanced law-of-thewall (Equations 12.10-28 and 12.10-30), ensuring a formulation that is valid throughout
the near-wall region.

The enhanced wall treatment is available with the following viscous models:
K-epsilon
Reynolds Stress
Enhanced wall functions are available with the following viscous models:
Spalart-Allmaras
K-omega
Large Eddy Simulation

c Fluent Inc. September 29, 2006


12-73

Modeling Turbulence

12.10.5

User-Defined Wall Functions

This option is only available when the k- model is enabled. Selecting the User-Defined
Wall Functions under Near-wall Treatment allows you to hook a Law-of-the-Wall UDF.
See Section 2.3.23: DEFINE WALL FUNCTIONS of the separate UDF Manual for details on
user-defined wall functions.

User-defined wall functions are available with the following viscous model:
K-epsilon

12.10.6

LES Near-Wall Treatment

When the mesh is fine enough to resolve the laminar sublayer, the wall shear stress is
obtained from the laminar stress-strain relationship:
u
u y
=
u

(12.10-42)

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:
1
u
u y
= ln E
u

(12.10-43)

where is the von Karman constant and E = 9.793. If the mesh is a such that the first
near wall point is within the buffer region, then two above laws are blended in accordance
with equation Equation 12.10-28.
For the LES simulations in FLUENT, there is an alternative near wall approach based on
the work of Werner and Wengle [398], who proposed analytical integration of power-law
near-wall velocity distribution resulting in the following expressions for the wall shear
stress:

|w | =

2|up |
z

1+B
1B 1B
A
2

1+B

1+B
A

B

|up |

2
1+B

for |up |

A 1B
2z

for |up | >

A 1B
2z

(12.10-44)

where up is velocity parallel to the wall, A = 8.3, B = 1/7 are the constants, and z is
the near-wall control volume length scale.

12-74

c Fluent Inc. September 29, 2006


12.11 Grid Considerations for Turbulent Flow Simulations

The Werner-Wengle wall functions can be enabled using the define/models/viscous/


near-wall-treatment/werner-wengle-wall-fn? text command.

12.11

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 parameters, you must
ascertain that turbulence quantities in complex turbulent flows 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 12.11.1 and 12.11.1 are general guidelines for
the near-wall mesh.

12.11.1

Near-Wall Mesh Guidelines

Wall Functions
The log-law, which is valid for equilibrium boundary layers and fully developed flows,
provides upper and lower bounds on the acceptable distance between the cell centroid
and the wall for wall-adjacent cells. The distance is usually measured in the wall unit,
y + ( u y/), or y . Note that y + and y have comparable values when the first cell is
placed in the log-layer.
For standard or non-equilibrium wall functions, each wall-adjacent cells centroid
should be located within the log-law layer, 30 < y + < 300. A y + value close to the
lower bound (y + 30) is most desirable.
Although FLUENT employs the linear (laminar) law when y < 11.225, using an
excessively fine mesh near the walls should be avoided, because the wall functions
cease to be valid in the viscous sublayer.

c Fluent Inc. September 29, 2006


12-75

Modeling Turbulence

As much as possible, the mesh should be made either coarse or fine enough to
prevent the wall-adjacent cells from being placed in the buffer layer (y + = 5 30).
The upper bound of the log-layer depends on, among others, pressure gradients
and Reynolds number. As the Reynolds number increases, the upper bound tends
to also increase. y + values that are too large are not desirable, because the wake
component becomes substantially large above the log-layer.
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.

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.

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
the full benefit of the Spalart-Allmaras model, the near-wall mesh spacing should be as
described in Section 12.11.1: Enhanced Wall Treatment for the enhanced wall treatment.
However, as discussed in Section 12.3.7: Wall Boundary Conditions, 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 12.11.1: Wall
Functions.

12-76

c Fluent Inc. September 29, 2006


12.12 Steps in Using a Turbulence Model

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.

k- Models
Both k- models available in FLUENT are available as low-Reynolds-number models as
well as high-Reynolds-number models. If the Transitional Flows option is enabled in the
Viscous Model panel, low-Reynolds-number variants will be used, and, in that case, mesh
guidelines should be the same as for the enhanced wall treatment. However, if this option
is not active, then the mesh guidelines should be the same as for the wall functions.

Large Eddy Simulation


For the LES implementation in FLUENT, the wall boundary conditions have been implemented using a law-of-the-wall approach as described in Section 12.9.4: Inlet Boundary
Conditions for the LES Model. This means that there are no computational restrictions
on the near-wall mesh spacing. However, for best results, it might be necessary to use a
very fine near-wall mesh spacing (on the order of y + = 1).

12.12

Steps in Using a Turbulence Model

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.)

c Fluent Inc. September 29, 2006


12-77

Modeling Turbulence

1. To activate the turbulence model, select Spalart-Allmaras, k-epsilon, k-omega, Reynolds


Stress, Detached Eddy Simulation (3D), and Large Eddy Simulation (LES) (3D) under
Model in the Viscous Model panel (Figure 12.12.1).
Define Models Viscous...

Figure 12.12.1: The Viscous Model Panel

12-78

c Fluent Inc. September 29, 2006


12.12 Steps in Using a Turbulence Model

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 Detached Eddy Simulation and the Large Eddy Simulation (LES) models
are 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
User-Defined Wall Functions
These near-wall options are described in detail in Section 12.10: Near-Wall Treatments for Wall-Bounded Turbulent Flows. By default, the standard wall function
is enabled.
The near-wall treatment for the Spalart-Allmaras, k-, and LES models is defined
automatically, as described in Sections 12.3.7, 12.5.3, and 12.9.4, respectively.
3. Enable the appropriate turbulence modeling options in the Viscous Model panel.
See Section 12.19: Setup Options for all Turbulence Modeling for details.
4. Specify the boundary conditions for the solution variables.
Define Boundary Conditions...
See Section 12.20: Defining Turbulence Boundary Conditions for details.
5. Specify the initial guess for the solution variables.
Solve Initialize Initialize...
See Section 12.21: Providing an Initial Guess for k and  (or k and ) for details.
Note that Reynolds stresses are automatically initialized using k, and therefore
need not be initialized.

c Fluent Inc. September 29, 2006


12-79

Modeling Turbulence

12.13 Setting Up the Spalart-Allmaras Model


If you choose the Spalart-Allmaras model, the following options are available:
vorticity-based production (Section 12.19.3: Vorticity- and Strain/Vorticity-Based
Production)
strain/vorticity-based production (Section 12.19.3: Vorticity- and Strain/VorticityBased Production)
viscous heating (always activated for the density-based solvers) (Section 12.19.1: Including the Viscous Heating Effects)

Figure 12.13.1: The Viscous Model Panel Displaying the Spalart-Allmaras


Options

12-80

c Fluent Inc. September 29, 2006


12.14 Setting Up the k- Model

12.14 Setting Up the k- Model


12.14.1

Setting Up the Standard or Realizable k- Model

If you choose the standard k- model or the realizable k- model, the following options
are available:
viscous heating (always activated for the density-based solvers) (Section 12.19.1: Including the Viscous Heating Effects)
inclusion of buoyancy effects on  (Section 12.4.5: Effects of Buoyancy on Turbulence
in the k- Models)

Figure 12.14.1: The Viscous Model Panel Displaying the Standard k- Model

c Fluent Inc. September 29, 2006


12-81

Modeling Turbulence

12.14.2

Setting Up the RNG k- Model

If you choose the RNG k- model, the following options are available:
differential viscosity model (Section 12.19.5: Differential Viscosity Modification)
swirl modification (Section 12.19.6: Swirl Modification)
viscous heating (always activated for the density-based solvers) (Section 12.19.1: Including the Viscous Heating Effects)
inclusion of buoyancy effects on  (Section 12.4.5: Effects of Buoyancy on Turbulence
in the k- Models)
For all k- models, one the following near-wall treatments must be selected (Section 12.10: NearWall Treatments for Wall-Bounded Turbulent Flows):
standard wall functions
non-equilibrium wall functions
enhanced wall treatment
user-defined wall functions
If you choose the enhanced wall treatment, the following options are available:
pressure gradient effects (Section 12.19.9: Including Pressure Gradient Effects)
thermal effects (Section 12.19.10: Including Thermal Effects)
If you choose the user-defined wall functions near-wall treatment, hook your UDF under
Law of the Wall, as shown in Figure 12.14.1.

12-82

c Fluent Inc. September 29, 2006


12.14 Setting Up the k- Model

Figure 12.14.2: The Viscous Model Panel Displaying the RNG k- Model

c Fluent Inc. September 29, 2006


12-83

Modeling Turbulence

12.15 Setting Up the k- Model


12.15.1

Setting Up the Standard k- Model

If you choose the standard k- model, the following options are available:
transitional flows (Section 12.19.7: Transitional Flows)
shear flow corrections (Section 12.19.8: Shear Flow Corrections)
viscous heating (always activated for the density-based solvers) (Section 12.19.1: Including the Viscous Heating Effects)

Figure 12.15.1: The Viscous Model Panel Displaying the Standard k- Model

The k- models use enhanced wall functions, described in Section 12.10.4: Enhanced
Wall Functions, as the near-wall treatment.

12-84

c Fluent Inc. September 29, 2006


12.15 Setting Up the k- Model

12.15.2

Setting Up the Shear-Stress Transport k- Model

If you choose the shear-stress transport k- model, the following options are available:
transitional flows (Section 12.19.7: Transitional Flows)
viscous heating (always activated for the density-based solvers) (Section 12.19.1: Including the Viscous Heating Effects)

Figure 12.15.2: The Viscous Model Panel Displaying the SST k- Model

c Fluent Inc. September 29, 2006


12-85

Modeling Turbulence

12.16 Setting Up the Reynolds Stress Model


If you choose the RSM, the following submodels are available:
Linear pressure-strain model (Section 12.7.4: Linear Pressure-Strain Model)
Quadratic pressure-strain model (Section 12.19.13: Quadratic Pressure-Strain Model)
Low-Re Stress-Omega (Section 12.19.14: Low-Re Stress-Omega Pressure-Strain)

Figure 12.16.1: The Viscous Model Panel Displaying the Reynolds Stress
Model Options

12-86

c Fluent Inc. September 29, 2006


12.16 Setting Up the Reynolds Stress Model

The following Reynolds-stress options are available:


wall boundary conditions for the Reynolds stresses from the k equation (Section 12.19.12: Solving the k Equation to Obtain Wall Boundary Conditions) for
the linear and quadratic pressure-strain models
wall reflection effects on Reynolds stresses (Section 12.19.11: Including the Wall
Reflection Term) for the linear pressure-strain model
Other options that are available based on your case setup include:
viscous heating (always activated for the density-based solvers) (Section 12.19.1: Including the Viscous Heating Effects)
inclusion of buoyancy effects on  (Section 12.4.5: Effects of Buoyancy on Turbulence
in the k- Models)
For the Reynolds stress model, the following near-wall treatments are available (Section 12.10: Near-Wall Treatments for Wall-Bounded Turbulent Flows):
standard wall functions
non-equilibrium wall functions
enhanced wall treatment
If wall boundary conditions for the Reynolds stresses from the k equation and/or wall
reflection effects on Reynolds stresses are/is selected, then all the above near-wall treatments are available for selection.
If you choose the enhanced wall treatment, the following options are available:
pressure gradient effects (Section 12.19.9: Including Pressure Gradient Effects)
thermal effects (Section 12.19.10: Including Thermal Effects)
If the quadratic pressure-strain model is selected, then you can set either the standard
wall functions or the non-equilibrium wall functions.
If Low-Re Stress-Omega is selected, you cannot select any near-wall treatments. You do
have the option of selecting any or all of the following k options:
transitional flows (Section 12.19.7: Transitional Flows)
shear flow corrections (Section 12.19.8: Shear Flow Corrections)

c Fluent Inc. September 29, 2006


12-87

Modeling Turbulence

Figure 12.16.2: The Viscous Model Panel Displaying the Low-Re StressOmega Model Options

12-88

c Fluent Inc. September 29, 2006


12.17 Setting Up the Detached Eddy Simulation Model

12.17 Setting Up the Detached Eddy Simulation Model


The following submodels are available when selecting the DES model (Section 12.19.4: Detached Eddy Simulation (DES) Modeling):
Spalart-Allmaras
Realizable k-
SST k-

Figure 12.17.1: The Viscous Model Panel Displaying the Detached Eddy Simulation Model Options

For the Realizable k- submodel, there are no other model-specific options to set. As for
the SST k- sub-model, the model-specific option that you can select is the Transitional
Flows k-omega Option (Section 12.19.7: Transitional Flows).

c Fluent Inc. September 29, 2006


12-89

Modeling Turbulence

Additionally, you can perform the following DES-specific functions by using the
/define/models/viscous/detached-eddy-simulation? text command:
Use cell volume-based LES length scale (default is to use maximum cell edge)
Modify only the length scales that appear in the destruction term in t equation
(the default is to modify all length scales within the t equation)

12.18

Setting Up the Large Eddy Simulation Model

If you choose the LES model, the following subgrid-scale submodels are available (Section 12.19.15: Subgrid-Scale Model):
Smagorinsky-Lilly
WALE
Kinetic-Energy Transport
The LES options that are available for the Smagorinsky-Lilly are
Dynamic Stress
Dynamic Energy Flux (available only when the Dynamic Stress Model is enabled)
Dynamic Scalar Flux
The LES option that is available when the Kinetic-Energy Transport submodel is selected
is the Dynamic Energy Flux Model.
It is also possible to modify the Model Constants, but this is not necessary for most
applications. See Sections 12.3 through 12.9 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 12.7-5 and 12.7-6, and C1-PS and C2-PS are the constants C10 and C20 in
Equation 12.7-7. C1-SSG-PS, C1-SSG-PS, C2-SSG-PS, C3-SSG-PS, C3-SSG-PS, C4-SSGPS, and C5-SSG-PS are the constants C1 , C1 , C2 , C3 , C3 , C4 , and C5 in the quadratic
pressure-strain approximation of Equation 12.7-16.

12-90

c Fluent Inc. September 29, 2006


12.18 Setting Up the Large Eddy Simulation Model

Figure 12.18.1: The Viscous Model Panel Displaying the Large Eddy Simulation Model Options

c Fluent Inc. September 29, 2006


12-91

Modeling Turbulence

12.19 Setup Options for all Turbulence Modeling


The various options available for the turbulence models are described in detail in Sections 12.3 through 12.9. Instructions for activating these options are provided here.

12.19.1

Including the Viscous Heating Effects

See Sections 13.2.1 and 13.2.2 for information on including viscous heating effects in your
model.

12.19.2

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 buoyancy (Gb in Equation 12.4-1) 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.

12.19.3

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 12.3-8 to
compute the value of the deformation tensor S; if you choose Strain/Vorticity-Based Production, it will use Equation 12.3-10.
(These options will not appear unless you have activated the Spalart-Allmaras model.)

12.19.4

Detached Eddy Simulation (DES) Modeling

If you enabled DES for the Spalart-Allmaras model as described at the beginning of this
By
section, FLUENT will use Equation 12.8-1 to compute the value of the length scale d.
default, the empirical constant Cdes is set to 0.65. You can change its value in the Cdes
field under Model Constants. Cdes is set to 0.61 for the Realizable k- and SST k- RANS
models (Sections 12.8.2 and 12.8.3).

12-92

c Fluent Inc. September 29, 2006


12.19 Setup Options for all Turbulence Modeling

12.19.5

Differential Viscosity Modification

In the RNG turbulence model in FLUENT, you have an option to use a differential formula
for effective viscosity eff (Equation 12.4-6) 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.)

12.19.6

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
(s in Equation 12.4-8) is set to 0.07, 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.)

12.19.7

Transitional Flows

If either of the k- models are used, you may enable a low-Reynolds-number correction to
the turbulent viscosity by enabling the Transitional Flows option under k-omega Options
in the Viscous Model panel. By default, this option is not enabled, and the damping
coefficient ( in Equation 12.5-6) is equal to 1.

12.19.8

Shear Flow Corrections

In the standard k- 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- model [403]. When this option is enabled, FLUENT will
calculate f and f using Equations 12.5-16 and 12.5-24, respectively. If this option is
disabled, f and f will be set equal to 1.

12.19.9

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 in Equation 12.10-31.

c Fluent Inc. September 29, 2006


12-93

Modeling Turbulence

12.19.10

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 in Equation 12.10-31. will also be
included in Equation 12.10-31 when the Thermal Effects option is enabled if the ideal gas
law is selected for the fluid density in the Materials panel.

12.19.11

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 ij,w using Equation 12.7-7 and include it in Equation 12.7-4. Note that wallreflection 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 ij,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 12.7-7
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.

12.19.12

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 12.7-34.
k is calculated by solving the k equation obtained by summing Equation 12.7-1 for
normal stresses. To disable this option and use the wall boundary conditions given in
Equation 12.7-35, 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.)

12.19.13

Quadratic Pressure-Strain Model

To use the quadratic pressure-strain model described in Section 12.7.4: Quadratic PressureStrain Model, turn on the Quadratic Pressure-Strain Model option under Reynolds-Stress
Options in the Viscous Model panel. (This option will not appear unless you have activated
the RSM.) The following options are not available when the Quadratic Pressure-Strain
Model is enabled:
Wall Reflection Effects under Reynolds-Stress Options
Enhanced Wall Treatment under Near-Wall Treatment

12-94

c Fluent Inc. September 29, 2006


12.19 Setup Options for all Turbulence Modeling

12.19.14

Low-Re Stress-Omega Pressure-Strain

To use the Low-Re Stress-Omega option described in Section 12.7.4: Low-Re StressOmega Model, turn on the Low-Re Stress-Omega 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 Low-Re Stress-Omega is enabled:
Wall BC from k Equation under Reynolds-Stress Options
Quadratic Pressure-Strain Model under Reynolds-Stress Options
Wall Reflection Effects under Reynolds-Stress Options
Standard Wall Functions under Near-Wall Treatment
Non-Equilibrium Wall Functions under Near-Wall Treatment
Enhanced Wall Treatment under Near-Wall Treatment
Instead, the following options have to be set:
Transitional Flows under k-omega Options
Shear Flow Corrections under k-omega Options

12.19.15

Subgrid-Scale Model

If you have selected the Large Eddy Simulation model, you will be able to choose one of the
subgrid-scale models described in Section 12.9.3: Subgrid-Scale Models. You can choose
from the Smagorinsky-Lilly, WALE, or Kinetic-Energy Transport subgrid-scale models. Note
that the Dynamic Model is an option with the Smagorinsky-Lilly model, while the KineticEnergy Transport model is always run as a dynamic model.
(These options will not appear unless you have activated the LES model.)

12.19.16

Customizing the Turbulent Viscosity

If you are using the Spalart-Allmaras, k-, k-, or LES model, a user-defined function
can be used to customize the turbulent viscosity. This option will enable you to modify
t in the case of the Spalart-Allmaras, k-, and k- models, and incorporate completely
new subgrid models in the case of the LES model. See the separate UDF Manual for
information about user-defined functions.
In the Viscous Model panel, under User-Defined Functions, select the appropriate userdefined 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.

c Fluent Inc. September 29, 2006


12-95

Modeling Turbulence

12.19.17

Customizing the Turbulent Prandtl Numbers

If you are using the standard or realizable k- model or the standard k- model, a userdefined function can be used to customize the turbulent Prandtl numbers. This option
will allow you to calculate k and either  or (depending on if you have enabled the
appropriate k- or k- model) by using a UDF. You will also be able to calculate the
value of the energy Prandtl number (Prt in Equation 12.4-23) and the Prandtl number
at the wall (Prt in Equation 12.10-5) in this way. See the separate UDF Manual for
information about user-defined functions.
In the Viscous Model panel, under User-Defined Functions, select the appropriate userdefined function from the drop-down lists under Prandtl Numbers. Options include: TKE
Prandtl Number, TDR Prandtl Number (k- models only), SDR Prandtl Number (k- model
only), Energy Prandtl Number, and Wall Prandtl Number.

12.19.18

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 selection of non-Newtonian options for the material viscosity. See Section 8.4.5: Viscosity
for Non-Newtonian Fluids for details about these options.

12.20

Defining Turbulence Boundary Conditions

12.20.1

The Spalart-Allmaras Model

When you are modeling turbulent flows in FLUENT using the Spalart-Allmaras model,
you must provide the boundary conditions for in addition to other mean solution variables. The boundary conditions for at the walls are internally taken care of by FLUENT,
which obviates the need for your inputs. The boundary condition input for you must
supply to FLUENT is the one at inlet boundaries (velocity inlet, pressure inlet, etc.). In
many situations, it is important to specify correct or realistic boundary conditions at the
inlets, because the inlet turbulence can significantly affect the downstream flow.
See Section 7.2.2: Determining Turbulence Parameters for details about specifying the
boundary condition for 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 7.13.1: Setting
the Roughness Parameters).

12-96

c Fluent Inc. September 29, 2006


12.20 Defining Turbulence Boundary Conditions

12.20.2

k- Models and k- Models

When you are modeling turbulent flows in FLUENT using one of the k- models or one
of the k- models, you must provide the boundary conditions for k and  (or k and )
in addition to other mean solution variables. The boundary conditions for k and  (or k
and ) at the walls are internally taken care of by FLUENT, which obviates the need for
your inputs. The boundary condition inputs for k and  (or k and ) you must supply
to FLUENT are the ones at inlet boundaries (velocity inlet, pressure inlet, etc.). In many
situations, it is important to specify correct or realistic boundary conditions at the inlets,
because the inlet turbulence can significantly affect the downstream flow.
See Section 7.2.2: Determining Turbulence Parameters for details about specifying the
boundary conditions for k and  (or k and ) at the inlets.
You may want to include the 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 7.13.1: Setting
the Roughness Parameters).
Additionally, you can control whether or not to set the turbulent viscosity to zero within
a laminar zone. If the 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 yes, FLUENT will set both
the production term in the turbulence transport equation and t to zero. In contrast,
when the Laminar Zone option is turned on in a Fluid boundary condition panel, only the
production term is set to zero. See Section 7.17.1: Specifying a Laminar Zone for details
about laminar zones.

Note that the laminar zone feature is also available for the Spalart-Allmaras
and RSM models.

c Fluent Inc. September 29, 2006


12-97

Modeling Turbulence

12.20.3

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 7.2.2: Determining Turbulence Parameters. FLUENT then uses the specified turbulence quantities to derive the Reynolds stresses at the inlet from the assumption of
isotropy of turbulence:

2
k
3
= 0.0

ui2 =
u0i u0j

(i = 1, 2, 3)

(12.20-1)
(12.20-2)

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 7.2.2: Determining Turbulence Parameters). 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.
You can set the Reynolds stresses by using constant values, profile functions of coordinates
(see Section 7.26: Boundary Profiles), or user-defined functions (see the separate UDF
Manual).

12-98

c Fluent Inc. September 29, 2006


12.20 Defining Turbulence Boundary Conditions

Figure 12.20.1: Specifying Inlet Boundary Conditions for the Reynolds


Stresses

c Fluent Inc. September 29, 2006


12-99

Modeling Turbulence

12.20.4

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 12.9.4: Inlet Boundary
Conditions for the LES Model. For all boundary types other than velocity inlets, the
boundary conditions for LES remain the same as for laminar flows.

12.21

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

For flows using one of the k- models, one of the k- models, or the RSM, the converged
solutions or (for unsteady calculations) the solutions after a sufficiently long time has
elapsed should be independent of the initial values for k and  (or k and ). For better
convergence, however, it is beneficial to use a reasonable initial guess for k and  (or k
and ).
In general, it is recommended that you start from a fully-developed state of turbulence.
When you use the enhanced wall treatment for the k- models or the RSM, it is critically
important to specify fully-developed turbulence 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 ) in the whole
domain from these boundary values. (See Section 25.14: Initializing the Solution
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. 510%
is enough to represent fully-developed turbulence. k can then be computed from
the turbulence intensity and the characteristic mean velocity magnitude of your
problem (k = 1.5(Iuavg )2 ).
2

You should specify an initial guess for  so that the resulting eddy viscosity (C 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 Equations 12.20-1 and 12.20-2.

12-100

c Fluent Inc. September 29, 2006


12.22 Solution Strategies for Turbulent Flow Simulations

12.22 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 turbulence quantities. Since the equations for mean quantities and the turbulent quantities
(t , k, , , or the Reynolds stresses) are strongly coupled in a highly non-linear fashion,
it takes more computational effort to obtain a converged turbulent solution than to obtain 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.

12.22.1

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 12.11: Grid Considerations for Turbulent
Flow Simulations for details.

12.22.2

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 12.2: Choosing a Turbulence Model).
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. September 29, 2006


12-101

Modeling Turbulence

12.22.3

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 density-based
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 ) fields. Particularly when the enhanced wall treatment
is used, it is important to start with a sufficiently developed turbulence field, as
recommended in Section 12.21: Providing an Initial Guess for k and  (or k and ),
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 switching to the RNG model. Due to the additional non-linearities in the RNG model,
lower under-relaxation factors and (for the density-based 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 12.10-25) instead of from its transport equation.

12.22.4

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 density-based 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.

12-102

c Fluent Inc. September 29, 2006


12.22 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 9.5: Swirling and Rotating Flows as well.

Under-Relaxation of the Reynolds Stresses


FLUENT applies under-relaxation to the Reynolds stresses. You can set under-relaxation
factors using the Solution Controls panel.
Solve Controls Solution...
The default settings of 0.5 are recommended for most cases. You may be able to increase
these settings and speed up the convergence when the RSM solution begins to converge.

Disabling Calculation Updates of the Reynolds Stresses


In some instances, you may wish to let the current Reynolds stress field remain fixed,
skipping the solution of the Reynolds transport equations while solving the other transport 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 103 usually indicate a practically-converged solution. However, you may need to apply tighter
convergence criteria (below 104 ) to ensure full convergence.

12.22.5

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.

c Fluent Inc. September 29, 2006


12-103

Modeling Turbulence

The following are suggestions to follow when running a large eddy simulation:
1. Start by running a steady state flow simulation using a Reynolds-averaged turbulence model such as standard k-, k-, Spalart-Allmaras, or even RSM. Run
until the flow field is reasonably converged and then use the solve/initialize/
init-instantaneous-vel text command to generate the instantaneous velocity
field out of the steady-state RANS results. This command must be executed before LES is enabled. This option is available for all RANS-based models and it will
create a much more realistic initial field for the LES run. Additionally, it will help
in reducing the time needed for the LES simulation to reach a statistically stable
mode. This step is optional.
2. When you enable LES, FLUENT will automatically turn on the unsteady solver
option and choose the second-order implicit formulation. You will need to set
the appropriate time step size and all the needed solution parameters. (See Section 25.17.1: User Inputs for Time-Dependent Problems for guidelines on setting
solution parameters for transient calculations in general.) The bounded centraldifferencing spatial discretization scheme will be automatically enabled for momentum equations. Both the bounded central-differencing and pure central-differencing
schemes are available for all equations when running LES simulations.
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 25.17.1: User Inputs for TimeDependent Problems. With this option enabled, FLUENT will gather data for
time statistics while performing a large eddy simulation. You can set the Sampling
Interval such that Data Sampling for Time Statistics can be performed at the specified
frequency. When Data Sampling for Time Statistics is enabled, the statistics collected
at each sampling interval can be postprocessed and you can then view both the
mean and the root-mean-square (RMS) values in FLUENT.
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.

12-104

c Fluent Inc. September 29, 2006


12.23 Postprocessing for Turbulent Flows

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 based schemes are recommended for all equations when you use the
LES model. FLUENT provides two central-differencing based schemes: pure centraldifferencing and bounded central-differencing. The bounded scheme is the default option
when you select LES or DES.
Solve Controls Solution...

12.23

Postprocessing for Turbulent Flows

FLUENT provides postprocessing options for displaying, plotting, and reporting various 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

c Fluent Inc. September 29, 2006


12-105

Modeling Turbulence

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- models are as follows:
Turbulent Kinetic Energy (k)
Turbulence Intensity
Specific Dissipation Rate (Omega)
Production of k
Turbulent Viscosity
Effective Viscosity
Turbulent Viscosity Ratio
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

12-106

c Fluent Inc. September 29, 2006


12.23 Postprocessing for Turbulent Flows

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
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 DES model are as follows:
Modified Turbulent Viscosity
Turbulent Viscosity
Effective Viscosity
Turbulent Viscosity Ratio
Effective Thermal Conductivity

c Fluent Inc. September 29, 2006


12-107

Modeling Turbulence

Effective Prandtl Number


Wall Yplus
Relative Length Scale (DES)
Turbulence quantities that can be reported for the LES model are as follows:
Turbulence Kinetic Energy
Turbulence Intensity
Subgrid Kinetic Energy
Production of k
Subgrid Turbulent Viscosity
Subgrid Effective Viscosity
Subgrid Turbulent Viscosity Ratio
Subgrid Filter Length
Subgrid Test-Filter Length
Subgrid Dissipation Rate
Subgrid Dynamic Viscosity Constant
Subgrid Dynamic Prandtl Number
Subgrid Dynamic Sc of Species
Subtest Kinetic Energy
Effective Thermal Conductivity
Effective Prandtl Number
Wall Ystar
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 30: Field Function
Definitions for their definitions.

12-108

c Fluent Inc. September 29, 2006


12.23 Postprocessing for Turbulent Flows

12.23.1

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:
the ratio of production of k to its dissipation (Gk /)
the ratio of the mean flow to turbulent time scale, ( Sk/)
the Reynolds stresses derived from the Boussinesq formula (e.g., uv = t u
)
y

12.23.2

Postprocessing Turbulent Flow Statistics

As described in Section 12.9: Large Eddy Simulation (LES) Model Theory, LES involves
the solution of a transient flow field, but it is the mean flow quantities that are of interest
from an engineering standpoint.
For all other turbulent flow, if Data Sampling for Time Statistics is enabled in the Iterate
panel, FLUENT gathers data for time statistics while performing the simulation. The
statistics that FLUENT collects at each sampling interval (which consists of the mean
and the root-mean-square (RMS) values) can be postprocessed by selecting Unsteady
Statistics... in any of the postprocessing panels. You can view several variables that
include, but are not limited to, shear stresses (Resolved UV/UW/VW Reynolds Stress), flow
heat fluxes (Resolved UT/VT/WT Heat Flux), and species statistics (RMS Mass Fraction
of species and Mean Mass Fraction of species). If you select Unsteady Wall Statistics...
in any of the postprocessing panels, you can view wall statistics such as Mean Pressure
Coefficient, Mean Wall Shear Stress, Mean X-Wall Shear Stress, Mean Y-Wall Shear Stress,
Mean Z-Wall Shear Stress, Mean Skin Friction Coefficient, Mean Surface Heat Flux, Mean
Surface Heat Transfer Coef., Mean Surface Nusselt Number, Mean Surface Stanton Number.
See Section 25.17.4: Postprocessing for Time-Dependent Problems for details.

Note that mean statistics are collected only in interior cells and not on
wall surfaces. Therefore, when node or cell values of mean quantities are
plotted on the wall surface, you are actually plotting values in nearby cells
attached to the wall.

c Fluent Inc. September 29, 2006


12-109

Modeling Turbulence

Figure 12.23.1: Postprocessing Choices for Unsteady Wall Statistics...

There may be cases when you want to control what set of variables are available for
postprocessing. To enable or disable certain variables, use the following text command:
solve set data-sampling
The text command prompts you with a set of questions that you will answer yes or no
to depending on whether or not you want statistics collected on certain variables. The
following example demonstrates using the text command when do not want to collect
statistics on species:

Data Sampling for Time Statistics? [yes]


Sampling interval [2]
Collect statistics for flow shear stresses? [yes]
Collect statistics for flow heat fluxes? [yes]
Collect wall statistics? [yes]
Collect statistics for ch3oh? [yes] no

12-110

When including or excluding statistics on variables, it is recommended that


you re-initialize the flow statistics.

c Fluent Inc. September 29, 2006


12.23 Postprocessing for Turbulent Flows

12.23.3

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 fullydeveloped 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.

c Fluent Inc. September 29, 2006


12-111

Modeling Turbulence

12-112

c Fluent Inc. September 29, 2006

Das könnte Ihnen auch gefallen