You are on page 1of 102

Chapter 10.

Modeling Turbulence

This chapter provides details about the turbulence models available in


FLUENT.
Information is presented in the following sections:
Section 10.1: Introduction
Section 10.2: Choosing a Turbulence Model
Section 10.3: The Spalart-Allmaras Model
Section 10.4: The Standard, RNG, and Realizable k- Models
Section 10.5: The Standard and Shear-Stress Transport (SST) k-
Models
Section 10.6: The Reynolds Stress Model (RSM)
Section 10.7: The Large Eddy Simulation (LES) Model
Section 10.8: Near-Wall Treatments for Wall-Bounded Turbulent
Flows
Section 10.9: Grid Considerations for Turbulent Flow Simulations
Section 10.10: Problem Setup for Turbulent Flows
Section 10.11: Solution Strategies for Turbulent Flow Simulations
Section 10.12: Postprocessing for Turbulent Flows

c Fluent Inc. November 28, 2001


10-1

Modeling Turbulence

10.1

Introduction

Turbulent flows are characterized by fluctuating velocity fields. These


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

10-2

c Fluent Inc. November 28, 2001


10.2 Choosing a Turbulence Model

10.2

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.
10.2.1

Reynolds-Averaged Approach vs. LES

A complete time-dependent solution of the exact Navier-Stokes equations for high-Reynolds-number turbulent flows in complex geometries
is unlikely to be attainable for some time to come. Two alternative
methods can be employed to transform the Navier-Stokes equations in
such a way that the small-scale turbulent fluctuations do not have to
be directly simulated: Reynolds averaging and filtering. Both methods
introduce additional terms in the governing equations that need to be
modeled in order to achieve closure. (Closure implies that there are a
sufficient number of equations for all the unknowns.)
The Reynolds-averaged Navier-Stokes (RANS) equations represent transport equations for the mean flow quantities only, with all the scales of
the turbulence being modeled. The approach of permitting a solution
for the mean flow variables greatly reduces the computational effort. If
the mean flow is steady, the governing equations will not contain time
derivatives and a steady-state solution can be obtained economically. A
computational advantage is seen even in transient situations, since the
time step will be determined by the global unsteadiness in the mean

c Fluent Inc. November 28, 2001


10-3

Modeling Turbulence

flow rather than by the turbulence. The Reynolds-averaged approach is


generally adopted for practical engineering calculations, and uses models
such as Spalart-Allmaras, k- and its variants, k- and its variants, and
the RSM.
LES provides an alternative approach in which the large eddies are computed in a time-dependent simulation that uses a set of filtered equations. Filtering is essentially a manipulation of the exact Navier-Stokes
equations to remove only the eddies that are smaller than the size of the
filter, which is usually taken as the mesh size. Like Reynolds averaging,
the filtering process creates additional unknown terms that must be modeled in order to achieve closure. Statistics of the mean flow quantities,
which are generally of most engineering interest, are gathered during the
time-dependent simulation. The attraction of LES is that, by modeling
less of the turbulence (and solving more), the error induced by the turbulence model will be reduced. One might also argue that it ought to be
easier to find a universal model for the small scales, which tend to be
more isotropic and less affected by the macroscopic flow features than
the large eddies.
It should, however, be stressed that the application of LES to industrial
fluid simulations is in its infancy. As highlighted in a recent review publication [72], typical applications to date have been for simple geometries.
This is mainly because of the large computer resources required to resolve the energy-containing turbulent eddies. Most successful LES has
been done using high-order spatial discretization, with great care being
taken to resolve all scales larger than the inertial subrange. The degradation of accuracy in the mean flow quantities with poorly resolved LES
is not well documented. In addition, the use of wall functions with LES
is an approximation that requires further validation.
As a general guideline, therefore, it is recommended that the conventional turbulence models employing the Reynolds-averaged approach be
used for practical calculations. The LES approach, described further in
Section 10.7, has been made available for you to try if you have the computational resources and are willing to invest the effort. The rest of this
section will deal with the choice of models using the Reynolds-averaged
approach.

10-4

c Fluent Inc. November 28, 2001


10.2 Choosing a Turbulence Model

10.2.2

Reynolds (Ensemble) Averaging

In Reynolds averaging, the solution variables in the instantaneous (exact) Navier-Stokes equations are decomposed into the mean (ensembleaveraged or time-averaged) and fluctuating components. For the velocity
components:
ui = u
i + u0i

(10.2-1)

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

(10.2-2)

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 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

"

xi xj

(10.2-3)

(ui uj ) =
(ui ) +
t
xj
ui
uj
2 ul
+
ij
xj
xi
3 xl

!#

(u0i u0j ) (10.2-4)


xj

Equations 10.2-3 and 10.2-4 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.

c Fluent Inc. November 28, 2001


10-5

Modeling Turbulence

Additional terms now appear that represent the effects of turbulence.


These Reynolds stresses, u0i u0j , must be modeled in order to close
Equation 10.2-4.
For variable-density flows, Equations 10.2-3 and 10.2-4 can be interpreted
as Favre-averaged Navier-Stokes equations [91], with the velocities representing mass-averaged values. As such, Equations 10.2-3 and 10.2-4
can be applied to density-varying flows.
10.2.3

Boussinesq Approach vs. Reynolds Stress Transport


Models

The Reynolds-averaged approach to turbulence modeling requires that


the Reynolds stresses in Equation 10.2-4 be appropriately modeled. A
common method employs the Boussinesq hypothesis [91] to relate the
Reynolds stresses to the mean velocity gradients:
u0i u0j

= t

ui
uj
+
xj
xi

ui
2

k + t
3
xi

ij

(10.2-5)

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.
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.

10-6

c Fluent Inc. November 28, 2001


10.2 Choosing a Turbulence Model

In many cases, models based on the Boussinesq hypothesis perform very


well, and the additional computational expense of the Reynolds stress
model is not justified. However, the RSM is clearly superior for situations
in which the anisotropy of turbulence has a dominant effect on the mean
flow. Such cases include highly swirling flows and stress-driven secondary
flows.
10.2.4

The Spalart-Allmaras Model

The Spalart-Allmaras model is a relatively simple one-equation model


that solves a 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 lowReynolds-number model, requiring the viscous-affected region of the
boundary layer to be properly resolved. In FLUENT, however, the SpalartAllmaras 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 5.1.2 for further discussion of
numerical error.
On a cautionary note, however, the Spalart-Allmaras model is still relatively new, and no claim is made regarding its suitability to all types
of complex engineering flows. For instance, it cannot be relied on to
predict the decay of homogeneous, isotropic turbulence. Furthermore,
one-equation models are often criticized for their inability to rapidly accommodate changes in length scale, such as might be necessary when

c Fluent Inc. November 28, 2001


10-7

Modeling Turbulence

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


10.2.5

The Standard k- Model

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 [128]. 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 [272] and the realizable k- model [209].
10.2.6

The RNG k- Model

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

10-8

c Fluent Inc. November 28, 2001


10.2 Choosing a Turbulence Model

effects. Effective use of this feature does, however, depend on an


appropriate treatment of the near-wall region.
These features make the RNG k- model more accurate and reliable for
a wider class of flows than the standard k- model.
10.2.7

The Realizable k- Model

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


from the standard k- model in two important ways:
The realizable k- model contains a new formulation for the turbulent viscosity.
A new transport equation for the dissipation rate, , has been derived from an exact equation for the transport of the mean-square
vorticity fluctuation.
The term realizable means that the model satisfies certain mathematical constraints on the Reynolds stresses, consistent with the physics of
turbulent flows. Neither the standard k- model nor the RNG k- model
is realizable.
An immediate benefit of the realizable k- model is that it more accurately predicts the spreading rate of both planar and round jets. It is
also likely to provide superior performance for flows involving rotation,
boundary layers under strong adverse pressure gradients, separation, and
recirculation.
Both the realizable and RNG k- models have shown substantial improvements over the standard k- model where the flow features include
strong streamline curvature, vortices, and rotation. Since the model is
still relatively new, it is not clear in exactly which instances the realizable k- model consistently outperforms the RNG model. However,
initial studies have shown that the realizable model provides the best
performance of all the k- model versions for several validations of separated flows and flows with complex secondary flow features.

c Fluent Inc. November 28, 2001


10-9

Modeling Turbulence

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 10.4-1710.4-19). 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.
10.2.8

The Standard k- Model

The standard k- model in FLUENT is based on the Wilcox k- model [267],


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 10.2.9.
10.2.9

The Shear-Stress Transport (SST) k- Model

The shear-stress transport (SST) k- model was developed by Menter [153]


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 nearwall region, which activates the standard k- model, and zero away
from the surface, which activates the transformed k- model.

10-10

c Fluent Inc. November 28, 2001


10.2 Choosing a Turbulence 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.
10.2.10

The Reynolds Stress Model (RSM)

The Reynolds stress model (RSM) 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 four 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.

c Fluent Inc. November 28, 2001


10-11

Modeling Turbulence

10.2.11

Computational Effort: CPU Time and Solution


Behavior

In terms of computation, the Spalart-Allmaras model is the least expensive turbulence model of the options provided in FLUENT, since only one
turbulence transport equation is solved.
The standard k- model clearly requires more computational effort than
the Spalart-Allmaras model since an additional transport equation is
solved. The realizable k- model requires only slightly more computational effort than the standard k- model. However, due to the extra
terms and functions in the governing equations and a greater degree of
non-linearity, computations with the RNG k- model tend to take 10
15% more CPU time than with the standard k- model. Like the k-
models, the k- 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.

10-12

c Fluent Inc. November 28, 2001


10.3 The Spalart-Allmaras Model

10.3

The Spalart-Allmaras Model

In turbulence models that employ the Boussinesq approach, the central


issue is how the eddy viscosity is computed. The model proposed by
Spalart and Allmaras [226] solves a transport equation for a quantity
that is a modified form of the turbulent kinematic viscosity.
10.3.1

Transport Equation for the Spalart-Allmaras Model

The transported variable in the Spalart-Allmaras model, , is identical


to the turbulent kinematic viscosity except in the near-wall (viscousaffected) region. The transport equation for is

1
G +
xj

(
ui ) =
(
) +
t
xi


( +
)
xj

+ Cb2


xj

!2
Y + S

(10.3-1)

where G is the production of turbulent viscosity and Y is the destruction of turbulent viscosity that occurs in the near-wall region due to
wall blocking and viscous damping. and 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, it is not taken into account when estimating
the Reynolds stresses in Equation 10.2-5.
10.3.2

Modeling the Turbulent Viscosity

The turbulent viscosity, t , is computed from


t =
fv1

(10.3-2)

where the viscous damping function, fv1 , is given by


fv1 =

3
3
3 + Cv1

(10.3-3)

and

c Fluent Inc. November 28, 2001


10-13

Modeling Turbulence

10.3.3

(10.3-4)

Modeling the Turbulent Production

The production term, G , is modeled as


G = Cb1 S

(10.3-5)

where

S S +

fv2
2 d2

(10.3-6)

fv2 = 1

1 + fv1

(10.3-7)

and

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

(10.3-8)

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


1
ij =
2

ui
uj

xj
xi

(10.3-9)

The justification for the default expression for S is that, for the wallbounded 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

10-14

c Fluent Inc. November 28, 2001


10.3 The Spalart-Allmaras Model

modification to the model has been proposed [46] 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 |)

(10.3-10)

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

(10.3-11)

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.
10.3.4

Modeling the Turbulent Destruction

The destruction term is modeled as


 2

Y = Cw1 fw

c Fluent Inc. November 28, 2001


(10.3-12)

10-15

Modeling Turbulence

where
"

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

#1/6

g = r + Cw2 r 6 r

(10.3-13)


(10.3-14)

S2 d2

(10.3-15)

Cw1 , Cw2 , and Cw3 are constants, and S is given by Equation 10.3-6.
Note that the modification described above to include the effects of mean
strain on S will also affect the value of S used to compute r.
10.3.5

Model Constants

The model constants Cb1 , Cb2 , , Cv1 , Cw1 , Cw2 , Cw3 , and have the following default values [226]:
Cb1 = 0.1335, Cb2 = 0.622, =

Cw1 =
10.3.6

2
, Cv1 = 7.1
3

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

10-16

(10.3-16)

c Fluent Inc. November 28, 2001


10.3 The Spalart-Allmaras Model

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
= ln E
u

u y

(10.3-17)

where u is the velocity parallel to the wall, u is the shear velocity, y is


the distance from the wall, is the von Karm
an constant (0.4187), and
E = 9.793.
10.3.7

Convective Heat and Mass Transfer Modeling

In FLUENT, turbulent heat transport is modeled using the concept of


Reynolds analogy to turbulent momentum transfer. The modeled
energy equation is thus given by the following:

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

"

cp t
k+
Prt

T
+ ui (ij )eff + Sh
xj
(10.3-18)

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

ui
2
eff
ij
3
xi

The term involving (ij )eff represents the viscous heating, and is always
computed in the coupled solvers. It is not computed by default in the
segregated solver, but it can be enabled in the Viscous Model panel. The
default value of the turbulent Prandtl number is 0.85. You can change
the value of Prt in the Viscous Model panel.
Turbulent mass transfer is treated similarly, with a default turbulent
Schmidt number of 0.7. This default value can be changed in the Viscous
Model panel.

c Fluent Inc. November 28, 2001


10-17

Modeling Turbulence

Wall boundary conditions for scalar transport are handled analogously


to momentum, using the appropriate law-of-the-wall.

10.4

The Standard, RNG, and Realizable k- Models

This section presents the standard, RNG, and realizable k- models. All
three models have similar forms, with transport equations for k and .
The major differences in the models are as follows:
the method of calculating turbulent viscosity
the turbulent Prandtl numbers governing the turbulent diffusion
of k and 
the generation and destruction terms in the  equation
The transport equations, methods of calculating turbulent viscosity, and
model constants are presented separately for each model. The features
that are essentially common to all models follow, including turbulent
production, generation due to buoyancy, accounting for the effects of
compressibility, and modeling heat and mass transfer.
10.4.1

The Standard k- Model

The standard k- model [128] 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, it was assumed that the flow is fully
turbulent, and the effects of molecular viscosity are negligible. The standard k- model is therefore valid only for fully turbulent flows.
Transport Equations for the Standard k- Model
The turbulence kinetic energy, k, and its rate of dissipation, , are obtained from the following transport equations:
10-18

c Fluent Inc. November 28, 2001


10.4 The Standard, RNG, and Realizable k- Models

(kui ) =
(k) +
t
xi
xj

"

t
+
k

k
+ Gk + Gb  YM + Sk
xj
(10.4-1)

and

(ui ) =
() +
t
xi
xj

"

t
+




+ C1 (Gk + C3 Gb )
xj
k
C2

2
+ S (10.4-2)
k

In these equations, Gk represents the generation of turbulence kinetic


energy due to the mean velocity gradients, calculated as described in
Section 10.4.4. Gb is the generation of turbulence kinetic energy due
to buoyancy, calculated as described in Section 10.4.5. YM represents
the contribution of the fluctuating dilatation in compressible turbulence
to the overall dissipation rate, calculated as described in Section 10.4.6.
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
(10.4-3)

where C is a constant.
Model Constants
The model constants C1 , C2 , C , k , and  have the following default
values [128]:
C1 = 1.44, C2 = 1.92, C = 0.09, k = 1.0,  = 1.3

c Fluent Inc. November 28, 2001


10-19

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 wall-bounded and free shear flows.
Although the default values of the model constants are the standard ones
most widely accepted, you can change them (if needed) in the Viscous
Model panel.
10.4.2

The RNG k- Model

The RNG-based k- turbulence model is derived from the instantaneous


Navier-Stokes equations, using a mathematical technique called renormalization group (RNG) 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 [36].
Transport Equations for the RNG k- Model
The RNG k- model has a similar form to the standard k- model:

(kui ) =
(k) +
t
xi
xj

k
k eff
xj

+ Gk + Gb  YM + Sk
(10.4-4)

and

(ui ) =
() +
t
xi
xj


 eff
xj


2
C1 (Gk + C3 Gb ) C2 R + S
k
k

(10.4-5)

In these equations, Gk represents the generation of turbulence kinetic


energy due to the mean velocity gradients, calculated as described in

10-20

c Fluent Inc. November 28, 2001


10.4 The Standard, RNG, and Realizable k- Models

Section 10.4.4. Gb is the generation of turbulence kinetic energy due


to buoyancy, calculated as described in Section 10.4.5. YM represents
the contribution of the fluctuating dilatation in compressible turbulence
to the overall dissipation rate, calculated as described in Section 10.4.6.
The quantities 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

(10.4-6)

where

= eff /
C

100

Equation 10.4-6 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-Reynoldsnumber and near-wall flows.
In the high-Reynolds-number limit, Equation 10.4-6 gives
t = C

k2


(10.4-7)

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-Reynolds-number form in Equation 10.4-7. However, there is an
option available that allows you to use the differential relation given in
Equation 10.4-6 when you need to include low-Reynolds-number effects.

c Fluent Inc. November 28, 2001


10-21

Modeling Turbulence

RNG Swirl Modification


Turbulence, in general, is affected by rotation or swirl in the mean flow.
The RNG model in FLUENT provides an option to account for the effects
of swirl or rotation by modifying the turbulent viscosity appropriately.
The modification takes the following functional form:


t = t0 f s , ,

k


(10.4-8)

where t0 is the value of turbulent viscosity calculated without the swirl


modification using either Equation 10.4-6 or Equation 10.4-7. 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 0.05 and cannot be modified. 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 0.6321 + 2.3929 0.3679
mol




=
1.3929
+ 2.3929
eff
0
0

(10.4-9)

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


k =  1.393.
The R Term in the  Equation
The main difference between the RNG and standard k- models lies in
the additional term in the  equation given by

R =

10-22

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

(10.4-10)

c Fluent Inc. November 28, 2001


10.4 The Standard, RNG, and Realizable k- Models

where Sk/, 0 = 4.38, = 0.012.


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

(ui ) =
() +
t
xi
xj


 eff
xj

2


+ C1 (Gk + C3 Gb ) C2

k
k
(10.4-11)

is given by
where C2

C2
C2 +

C 3 (1 /0 )
1 + 3

(10.4-12)

In regions where < 0 , the R term makes a positive contribution, and


becomes larger than C . In the logarithmic layer, for instance, it can
C2
2
2.0, which is close in magnitude to
be shown that 3.0, giving C2
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 neg less than C . In comparison
ative contribution, making the value of C2
2
with the standard k- model, the smaller destruction of  augments ,
reducing k and, eventually, the effective viscosity. As a result, in rapidly
strained flows, the RNG model yields a lower turbulent viscosity than
the standard k- model.
Thus, the RNG model is more responsive to the effects of rapid strain
and streamline curvature than the standard k- model, which explains
the superior performance of the RNG model for certain classes of flows.
Model Constants
The model constants C1 and C2 in Equation 10.4-5 have values derived analytically by the RNG theory. These values, used by default in
FLUENT, are

c Fluent Inc. November 28, 2001


10-23

Modeling Turbulence

C1 = 1.42, C2 = 1.68


10.4.3

The Realizable k- Model

In addition to the standard and RNG-based k- models described in


Sections 10.4.1 and 10.4.2, FLUENT also provides the so-called realizable
k- model [209]. The term realizable means that the model satisfies
certain mathematical constraints on the normal stresses, consistent with
the physics of turbulent flows. To understand this, consider combining
the Boussinesq relationship (Equation 10.2-5) and the eddy viscosity
definition (Equation 10.4-3) to obtain the following expression for the
normal Reynolds stress in an incompressible strained mean flow:
u2 =

U
2
k 2 t
3
x

(10.4-13)

Using Equation 10.4-3 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

(10.4-14)

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 [191],
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.
Another weakness of the standard k- model or other traditional k-
models lies with the modeled equation for the dissipation rate (). The
well-known round-jet anomaly (named based on the finding that the

10-24

c Fluent Inc. November 28, 2001


10.4 The Standard, RNG, and Realizable k- Models

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. [209] 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 [191]
a new model equation for dissipation () based on the dynamic
equation of the mean-square vorticity fluctuation
Transport Equations for the Realizable k- Model
The modeled transport equations for k and  in the realizable k- model
are

(kuj ) =
(k) +
t
xi
xi

"

t
+
k

k
+ Gk + Gb  YM + Sk
xj
(10.4-15)

and

(uj ) =
() +
t
xj
xj

"

C2

t
+



+ C1 S
xj

2

+ C1 C3 Gb + S (10.4-16)
k + 
k

where


C1 = max 0.43,

+5

and

c Fluent Inc. November 28, 2001


10-25

Modeling Turbulence

=S

k


In these equations, Gk represents the generation of turbulence kinetic


energy due to the mean velocity gradients, calculated as described in
Section 10.4.4. Gb is the generation of turbulence kinetic energy due to
buoyancy, calculated as described in Section 10.4.5. YM represents the
contribution of the fluctuating dilatation in compressible turbulence to
the overall dissipation rate, calculated as described in Section 10.4.6. 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 10.4-15) is the same as that in the
standard k- model (Equation 10.4-1) and the RNG k- model (Equation 10.4-4), except for the model constants. However, the form of the 
equation is quite different from those in the standard and RNG-based k-
models (Equations 10.4-2 and 10.4-5). One of the noteworthy features
is that the production term in the  equation (the second term on the
right-hand side of Equation 10.4-16) does not involve the production of
k; i.e., it does not contain the same Gk term as the other k- models. It
is believed that the present form better represents the spectral energy
transfer. Another desirable feature is that the destruction term (the next
to last term on the right-hand side of Equation 10.4-16) does not have
any singularity; i.e., its denominator never vanishes, even if k vanishes
or becomes smaller than zero. This feature is contrasted with traditional
k- models, which have a singularity due to k in the denominator.
This model has been extensively validated for a wide range of flows [116,
209], 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 roundjet anomaly; i.e., it predicts the spreading rate for axisymmetric jets as
well as that for planar jets.

10-26

c Fluent Inc. November 28, 2001


10.4 The Standard, RNG, and Realizable k- Models

Modeling the Turbulent Viscosity


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

t = C

k2


(10.4-17)

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

(10.4-18)

ij
ij
Sij Sij +

(10.4-19)

C =
where
U

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
=

Sij Sjk Ski q


1
, S = Sij Sij
cos1 ( 6W ), W =
3
S
1
Sij =
2

uj
ui
+
xi
xj

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

c Fluent Inc. November 28, 2001


10-27

Modeling Turbulence

and ). C in Equation 10.4-17 can be shown to recover the standard


value of 0.09 for an inertial sublayer in an equilibrium boundary layer.
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
10.4.4

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

(10.4-20)

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


Gk = t S 2

(10.4-21)

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


S
10.4.5

2Sij Sij

(10.4-22)

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 10.4-1, 10.4-4, and 10.4-15), and the
corresponding contribution to the production of  in Equations 10.4-2,
10.4-5, and 10.4-16.

10-28

c Fluent Inc. November 28, 2001


10.4 The Standard, RNG, and Realizable k- Models

The generation of turbulence due to buoyancy is given by


Gb = gi

t T
Prt xi

(10.4-23)

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 10.4-9,
but with 0 = 1/Pr = k/cp . The coefficient of thermal expansion, , is
defined as
1
=

(10.4-24)
p

For ideal gases, Equation 10.4-23 reduces to


Gb = gi

t
Prt xi

(10.4-25)

It can be seen from the transport equations for k (Equations 10.4-1,


10.4-4, and 10.4-15) that turbulence kinetic energy tends to be augmented (Gb > 0) in unstable stratification. For stable stratification,
buoyancy tends to suppress the turbulence (Gb < 0). In 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 10.4-2, 10.4-5, or 10.4-16).
However, you can include the buoyancy effects on  in the Viscous Model
panel. In this case, the value of Gb given by Equation 10.4-25 is used in
the transport equation for  (Equation 10.4-2, 10.4-5, or 10.4-16).
The degree to which  is affected by the buoyancy is determined by the
constant C3 . In FLUENT, C3 is not specified, but is instead calculated
according to the following relation [90]:

c Fluent Inc. November 28, 2001


10-29

Modeling Turbulence

v
C3 = tanh
u

(10.4-26)

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.
10.4.6

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 [267]. 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 [197]:
YM = 2M2t

(10.4-27)

where Mt is the turbulent Mach number, defined as


s

Mt =
where a (

k
a2

(10.4-28)

RT ) is the speed of sound.

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

10-30

c Fluent Inc. November 28, 2001


10.4 The Standard, RNG, and Realizable k- Models

10.4.7

Convective Heat and Mass Transfer Modeling in the k-


Models

In FLUENT, turbulent heat transport is modeled using the concept of


Reynolds analogy to turbulent momentum transfer. The modeled
energy equation is thus given by the following:

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

T
keff
+ ui (ij )eff
xj

+ Sh (10.4-29)

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

ui
2
eff
ij
3
xi

The term involving (ij )eff represents the viscous heating, and is always
computed in the coupled solvers. It is not computed by default in the
segregated solver, but it can be enabled in the Viscous Model panel.
Additional terms may appear in the energy equation, depending on the
physical models you are using. See Section 11.2.1 for more details.
For the standard and realizable k- models, the effective thermal conductivity is given by
keff = k +

cp 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

c Fluent Inc. November 28, 2001


10-31

Modeling Turbulence

where is calculated from Equation 10.4-9, but with 0 = 1/Pr = k/cp .


The fact that varies with mol /eff , as in Equation 10.4-9, is an advantage of the RNG k- model. It is consistent with experimental evidence
indicating that the turbulent Prandtl number varies with the molecular
Prandtl number and turbulence [111]. Equation 10.4-9 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 low-Reynolds-number regions. Equation 10.4-9 smoothly
predicts the variation of effective 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 10.4-9 is 0 = 1/Sc, where Sc is the
molecular Schmidt number.

10-32

c Fluent Inc. November 28, 2001


10.5 The Standard and SST k- Models

The Standard and SST k- Models

10.5

This section presents the standard and shear-stress transport (SST) k models. Both models have similar forms, with transport equations
for k and . The major ways in which the SST model 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.
10.5.1

The Standard k- Model

The standard k- model is an empirical model based on model transport


equations for the turbulence kinetic energy (k) and the specific dissipation rate (), which can also be thought of as the ratio of  to k [267].
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.
Transport Equations for the Standard k- Model
The turbulence kinetic energy, k, and the specific dissipation rate, , are
obtained from the following transport equations:

(kui ) =
(k) +
t
xi
xj

k
k
xj

+ Gk Yk + Sk

(10.5-1)

and

c Fluent Inc. November 28, 2001


10-33

Modeling Turbulence

(ui ) =
() +
t
xi
xj

xj

+ G Y + S

(10.5-2)

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 = +

(10.5-3)

(10.5-4)

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 =

(10.5-5)

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

0 + Ret /Rk
1 + Ret /Rk

(10.5-6)

where
Ret =

10-34

(10.5-7)

c Fluent Inc. November 28, 2001


10.5 The Standard and SST k- Models

Rk = 6
i
0 =
3
i = 0.072

(10.5-8)
(10.5-9)
(10.5-10)

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

(10.5-11)

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


Gk = t S 2

(10.5-12)

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


same way as for the k- model (see Equation 10.4-22).
Production of
The production of is given by

G = Gk
k

(10.5-13)

where Gk is given by Equation 10.5-11.


The coefficient is given by

c Fluent Inc. November 28, 2001


0 + Ret /R
1 + Ret /R

(10.5-14)

10-35

Modeling Turbulence

where R = 2.95. and Ret are given by Equations 10.5-6 and 10.5-7,
respectively.
Note that, in the high-Reynolds-number form of the k- model, =
= 1.
Modeling the Turbulence Dissipation
Dissipation of k
The dissipation of k is given by
Yk = f k

(10.5-15)

where

f =

1+6802k
1+4002k

where
k

k 0

(10.5-16)

k > 0

1 k
3 xj xj

(10.5-17)

and
= i [1 + F (Mt )]

i =

)4

4/15 + (Ret /R
1 + (Ret /R )4

= 1.5

(10.5-18)
(10.5-19)
(10.5-20)

R = 8

(10.5-21)

(10.5-22)

= 0.09

where Ret is given by Equation 10.5-7.


Dissipation of
The dissipation of is given by

10-36

c Fluent Inc. November 28, 2001


10.5 The Standard and SST k- Models

Y = f 2

(10.5-23)

where
1 + 70
1 + 80


ij jk Ski


=
)3
(

f =

1
2

ij =

ui
uj

xj
xi

(10.5-24)
(10.5-25)
!

(10.5-26)

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




= i 1

i
F (Mt )
i

(10.5-27)

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


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

F (Mt ) =

0
Mt Mt0
M2t M2t0 Mt > Mt0

(10.5-28)

where
2k
a2
= 0.25

M2t
Mt0

a =

RT

(10.5-29)
(10.5-30)
(10.5-31)

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

In the incompressible form, = i .

c Fluent Inc. November 28, 2001


10-37

Modeling Turbulence

Model Constants
1

= 0.09, i = 0.072, R = 8
,
9
Rk = 6, R = 2.95, = 1.5, Mt0 = 0.25, k = 2.0, = 2.0
= 1, = 0.52, 0 =

Wall Boundary Conditions


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

w =

(u )2 +

(10.5-32)

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



+
+ = min w
,

where
+
w
=

6
(y + )2

 2
50

ks+ < 25

ks+

ks

100
ks+

ks+

(10.5-33)

(10.5-34)
25

where

ks u
= max 1.0,

(10.5-35)

and ks is the roughness height.


In the logarithmic (or turbulent) region, the value of + is
1 du+
turb
+ = p
dy +
10-38

(10.5-36)

c Fluent Inc. November 28, 2001


10.5 The Standard and SST k- Models

which leads to the value of in the wall cell as


u
=p
y

(10.5-37)

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.
10.5.2

The Shear-Stress Transport (SST) k- Model

In addition to the standard k- model described in Section 10.5.1, FLUENT also provides a variation called the shear-stress transport (SST)
k- model, so named because the definition of the turbulent viscosity
is modified to account for the transport of the principal turbulent shear
stress. It is this feature that gives the SST k- model an advantage in
terms of performance over both the standard k- model and 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.
Transport Equations for the SST k- Model
The SST k- model has a similar form to the standard k- model:

(kui ) =
(k) +
t
xi
xj

k
k
xj

+ Gk Yk + Sk

(10.5-38)

and

(ui ) =
() +
t
xi
xj

xj

+ G Y + D + S (10.5-39)

In these equations, Gk represents the generation of turbulence kinetic


energy due to mean velocity gradients, calculated as described in Section 10.5.1. G represents the generation of , calculated as described in

c Fluent Inc. November 28, 2001


10-39

Modeling Turbulence

Section 10.5.1. 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 10.5.1. 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 = +

(10.5-40)

(10.5-41)

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

k
1
h
i
max 1 , F2

a1

(10.5-42)

where


k =
=

2ij ij

(10.5-43)

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

(10.5-44)
(10.5-45)

ij is the mean rate-of-rotation tensor and is defined in Equation 10.5-6.


The blending functions, F1 and F2 , are given by


F1 = tanh 41
10-40

(10.5-46)

c Fluent Inc. November 28, 2001


10.5 The Standard and SST k- Models


"

1 = min max
"

D+

!
#

k
4k
500
,
,
0.09y y 2
,2 D+ y 2

1 1 k
= max 2
, 1020
,2 xj xj


"

F2 = tanh 22
2 = max 2

k 500
,
0.09y y 2

(10.5-47)

(10.5-48)

(10.5-49)
(10.5-50)

where y is the distance to the next surface and D+ is the positive portion
of the cross-diffusion term (see Equation 10.5-60).
Modeling the Turbulence Production
Production of k
The term Gk represents the production of turbulence kinetic energy,
and is defined in the same manner as in the standard k- model. See
Section 10.5.1 for details.
Production of
The term G represents the production of and is given by
G =

Gk
t

(10.5-51)

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

(10.5-52)

where

c Fluent Inc. November 28, 2001


10-41

Modeling Turbulence

,1 =
,2 =

i,1
2
p

w,1
2
i,2

w,2

(10.5-53)
(10.5-54)

where is 0.41. i,1 and i,2 are given by Equations 10.5-58 and 10.5-59,
respectively.
Modeling the Turbulence Dissipation
Dissipation of k
The term Yk represents the dissipation of turbulence kinetic energy, and
is defined in a similar manner as in the standard k- model (see Section 10.5.1). 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

(10.5-55)

Dissipation of
The term Y represents the dissipation of , and is defined in a similar
manner as in the standard k- model (see Section 10.5.1). 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 10.5-24. For the SST k- model, f is a constant equal to 1. Thus,
Yk = 2

(10.5-56)

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


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

(10.5-57)

where
10-42

c Fluent Inc. November 28, 2001


10.6 The Reynolds Stress Model (RSM)

i,1 = 0.075

(10.5-58)

i,2 = 0.0828

(10.5-59)

and F1 is obtained from Equation 10.5-46.


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 10.5-39). D is defined as
D = 2 (1 F1 ) ,2

1 k
xj xj

(10.5-60)

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


, R , R , R , , and
All additional model constants ( , , 0 ,

k
Mt0 ) have the same values as for the standard k- model (see Section 10.5.1).

10.6

The Reynolds Stress Model (RSM)

The Reynolds stress model [75, 125, 126] involves calculation of the individual Reynolds stresses, u0i u0j , using differential transport equations.
The individual Reynolds stresses are then used to obtain closure of the
Reynolds-averaged momentum equation (Equation 10.2-4).

c Fluent Inc. November 28, 2001


10-43

Modeling Turbulence

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 Reynolds-averaged. Unfortunately, several of the terms in the exact equation are unknown and
modeling assumptions are required in order to close the equations.
In this section, the Reynolds stress transport equations are presented
together with the modeling assumptions required to attain closure.
10.6.1

The Reynolds Stress Transport Equations

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

(uk u0i u0j ) =


( u0i u0j )
+
t
x
k
|
{z
}
|
{z
}
Local Time Derivative Cij Convection





u0i u0j u0k + p kj u0i + ik u0j


xk
{z

DT,ij Turbulent Diffusion




u0i u0k

uj
ui
+ u0j u0k
xk
xk

Pij Stress Production


+

u0j
u0i
+
xj
xi
{z

{z

(gi u0j + gj u0i )


|

{z

Gij Buoyancy Production

2
|

ij Pressure Strain


{z

DL,ij Molecular Diffusion

2k u0j u0m ikm + u0i u0m jkm

(u0 u0 )
xk
xk i j

{z

u0i u0j
xk xk
{z

ij Dissipation


}

Fij Production by System Rotation

Suser

| {z }

User-Defined Source Term


(10.6-1)

10-44

c Fluent Inc. November 28, 2001


10.6 The Reynolds Stress Model (RSM)

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.
10.6.2

Modeling Turbulent Diffusive Transport

DT,ij can be modeled by the generalized gradient-diffusion model of Daly


and Harlow [48]:

DT,ij

= Cs
xk

ku0k u0` u0i u0j


x`

(10.6-2)

However, this equation can result in numerical instabilities, so it has been


simplified in FLUENT to use a scalar turbulent diffusivity as follows [138]:

DT,ij

=
xk

t u0i u0j
k xk

(10.6-3)

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


Lien and Leschziner [138] derived a value of k = 0.82 by applying the
generalized gradient-diffusion model, Equation 10.6-2, to the case of a
planar homogeneous shear flow. Note that this value of k is different
from that in the standard and realizable k- models, in which k = 1.0.
10.6.3

Modeling the Pressure-Strain Term

Linear Pressure-Strain Model


By default in FLUENT, the pressure-strain term, ij , in Equation 10.6-1
is modeled according to the proposals by Gibson and Launder [75], Fu
et al. [71], and Launder [124, 125].
The classical approach to modeling ij uses the following decomposition:
ij = ij,1 + ij,2 + ij,w

c Fluent Inc. November 28, 2001


(10.6-4)

10-45

Modeling Turbulence

where ij,1 is the slow pressure-strain term, also known as the return-toisotropy 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 C1

 0 0
2
u u ij k
k i j 3

(10.6-5)

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

(10.6-6)

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


k3/2
3
3

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


k
2
2
C` d

 3/2
k
3
3
+ C20 km,2 nk nm ij ik,2 nj nk jk,2ni nk
2
2
C` d
(10.6-7)
C10

where C10 = 0.5, C20 = 0.3, nk is the xk component of the unit normal to
3/4
the wall, d is the normal distance to the wall, and C` = C /, where
C = 0.09 and is the von Karm
an constant (= 0.4187).
ij,w is included by default in the Reynolds stress model.

10-46

c Fluent Inc. November 28, 2001


10.6 The Reynolds Stress Model (RSM)

Low-Re Modifications to the Linear Pressure-Strain Model


When the RSM is applied to near-wall flows using the enhanced wall
treatment described in Section 10.8.3, the pressure-strain model needs
to be modified. The modification used in FLUENT specifies the values
of C1 , C2 , C10 , and C20 as functions of the Reynolds stress invariants and
the turbulent Reynolds number, according to the suggestion of Launder
and Shima [127]:
n

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

C2 = 0.75 A
2
C10 = C1 + 1.67
3
"

C20

= max

2
3 C2

C2

1
6

io

(10.6-8)
(10.6-9)
(10.6-10)

,0

(10.6-11)

with the turbulent Reynolds number defined as Ret = (k2 /). The
parameter A and tensor invariants, A2 and A3 , are defined as


9
A 1 (A2 A3 )
8
A2 aik aki

(10.6-12)
(10.6-13)

A3 aik akj aji

(10.6-14)

aij is the Reynolds-stress anisotropy tensor, defined as


aij =

u0i u0j + 23 kij


k

(10.6-15)

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

c Fluent Inc. November 28, 2001


10-47

Modeling Turbulence

Quadratic Pressure-Strain Model


An optional pressure-strain model proposed by Speziale, Sarkar, and
Gatski [228] 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

+ C2  bik bkj


1
bmn bmn ij
3
q

+ C3 C3 bij bij kSij




2
+C4 k bik Sjk + bjk Sik bmn Smn ij
3
+C5 k (bik jk + bjk ik ) (10.6-16)
where bij is the Reynolds-stress anisotropy tensor defined as
u0i u0j + 23 kij
bij =
2k

(10.6-17)

The mean strain rate, Sij , is defined as


1
Sij =
2

uj
ui
+
xi
xj

(10.6-18)

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


1
ij =
2

10-48

ui
uj

xj
xi

(10.6-19)

c Fluent Inc. November 28, 2001


10.6 The Reynolds Stress Model (RSM)

The constants are

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


The quadratic pressure-strain model does not require a correction to
account for the wall-reflection effect in order to obtain a satisfactory
solution in the logarithmic region of a turbulent boundary layer. It
should be noted, however, that the quadratic pressure-strain model is
not available when the enhanced wall treatment is selected in the Viscous
Model panel.
10.6.4

Effects of Buoyancy on Turbulence

The production terms due to buoyancy are modeled as


t
Gij =
Prt

T
T
gi
+ gj
xj
xi

(10.6-20)

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 10.4-24, the following expression is obtained for Gij for ideal
gases:
t
Gij =
Prt
10.6.5

gi
+ gj
xj
xi

(10.6-21)

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:
k=

c Fluent Inc. November 28, 2001


1 0 0
uu
2 i i

(10.6-22)

10-49

Modeling Turbulence

As described in Section 10.6.8, an option is available in FLUENT to solve


a transport equation for the turbulence kinetic energy in order to obtain
boundary conditions for the Reynolds stresses. In this case, the following
model equation is used:
"

t k

(kui ) =
+
+
(k) +
t
xi
xj
k xj
1
(Pii + Gii ) (1 + 2M2t ) + Sk
2

(10.6-23)

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


obtainable by contracting the modeled equation for the Reynolds stresses
(Equation 10.6-1). As one might expect, it is essentially identical to
Equation 10.4-1 used in the standard k- model.
Although Equation 10.6-23 is solved globally throughout the flow domain, the values of k obtained are used only for boundary conditions. In
every other case, k is obtained from Equation 10.6-22. This is a minor
point, however, since the values of k obtained with either method should
be very similar.
10.6.6

Modeling the Dissipation Rate

The dissipation tensor, ij , is modeled as


2
ij = ij ( + YM )
3

(10.6-24)

where YM = 2M2t is an additional dilatation dissipation term according to the model by Sarkar [197]. The turbulent Mach number in
this term is defined as
s

Mt =

10-50

k
a2

(10.6-25)

c Fluent Inc. November 28, 2001


10.6 The Reynolds Stress Model (RSM)

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
+



+
xj

1

2
C1 [Pii + C3 Gii ] C2 + S
2
k
k

(10.6-26)

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 10.4.5, and S is a user-defined source term.
10.6.7

Modeling the Turbulent Viscosity

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


t = C

k2


(10.6-27)

where C = 0.09.
10.6.8

Boundary Conditions for the Reynolds Stresses

Whenever flow enters the domain, FLUENT requires values for individual
Reynolds stresses, u0i u0j , and for the turbulence dissipation rate, . These
quantities can be input directly or derived from the turbulence intensity
and characteristic length, as described in Section 10.10.2.
At walls, FLUENT computes the near-wall values of the Reynolds stresses
and  from wall functions (see Section 10.8.2). FLUENT applies explicit
wall boundary conditions for the Reynolds stresses by using the log-law
and the assumption of equilibrium, disregarding convection and diffusion
in the transport equations for the stresses (Equation 10.6-1). Using
a local coordinate system, where is the tangential coordinate, is

c Fluent Inc. November 28, 2001


10-51

Modeling Turbulence

the normal coordinate, and is the binormal coordinate, the Reynolds


stresses at the wall-adjacent cells are computed from

0
u02
u0 u0
u2
u0 2
= 1.098,
= 0.247,
= 0.655,
= 0.255
k
k
k
k

(10.6-28)

To obtain k, FLUENT solves the transport equation of Equation 10.6-23.


For reasons of computational convenience, the equation is solved globally,
even though the values of k thus computed are needed only near the wall;
in the far field k is obtained directly from the normal Reynolds stresses
using Equation 10.6-22. By default, the values of the Reynolds stresses
near the wall are fixed using the values computed from Equation 10.6-28,
and the transport equations in Equation 10.6-1 are solved only in the
bulk flow region.
Alternatively, the Reynolds stresses can be explicitly specified in terms
of wall-shear stress, instead of k:
0
u02
u0 u0
u2
u02
=
5.1,
=
1.0,
=
2.3,

= 1.0
u2
u2
u2
u2

(10.6-29)

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.
10.6.9

Convective Heat and Mass Transfer Modeling

With the Reynolds stress model in FLUENT, turbulent heat transport is


modeled using the concept of Reynolds analogy to turbulent momentum
transfer. The modeled energy equation is thus given by the following:

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

10-52

"

cp t
k+
Prt

T
+ ui (ij )eff + Sh
xj
(10.6-30)

c Fluent Inc. November 28, 2001


10.7 The Large Eddy Simulation (LES) Model

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

(ij )eff = eff

uj
ui
+
xi
xj

ui
2
eff
ij
3
xi

The term involving (ij )eff represents the viscous heating, and is always
computed in the coupled solvers. It is not computed by default in the
segregated solver, but it can be enabled in the Viscous Model panel. The
default value of the turbulent Prandtl number is 0.85. You can change
the value of Prt in the Viscous Model panel.
Turbulent mass transfer is treated similarly, with a default turbulent
Schmidt number of 0.7. This default value can be changed in the Viscous
Model panel.

10.7

The Large Eddy Simulation (LES) Model

Turbulent flows are characterized by eddies with a wide range of length


and time scales. The largest eddies are typically comparable in size
to the characteristic length of the mean flow. The smallest scales are
responsible for the dissipation of turbulence kinetic energy.
It is theoretically possible to directly resolve the whole spectrum of turbulent scales using an approach known as direct numerical simulation
(DNS). DNS is not, however, feasible for practical engineering problems. To understand the large computational cost of DNS, consider that
the ratio of the large (energy-containing) scales to the small (energy3/4
dissipating) scales is proportional to Ret , where Ret is the turbulent
Reynolds number. Therefore, to resolve all the scales, the mesh size
9/4
in three dimensions will be proportional to Ret . Simple arithmetic
shows that, for high Reynolds numbers, the mesh sizes required for DNS
are prohibitive. Adding to the computational cost is the fact that the
simulation will be a transient one with very small time steps, since the
temporal resolution requirements are governed by the dissipating scales,
rather than the mean flow or the energy-containing eddies.

c Fluent Inc. November 28, 2001


10-53

Modeling Turbulence

As explained in Section 10.2.1, the conventional approach to flow simulations employs the solution of the Reynolds-averaged Navier-Stokes
(RANS) equations. In the RANS approach, all the turbulent motions
are modeled, resulting in a significant savings in computational effort.
Conceptually, large eddy simulation (LES) is situated somewhere between DNS and the RANS approach. Basically large eddies are resolved
directly in LES, while small eddies are modeled. The rationale behind
LES can be summarized as follows:
Momentum, mass, energy, and other passive scalars are transported mostly by large eddies.
Large eddies are more problem-dependent. They are dictated by
the geometries and boundary conditions of the flow involved.
Small eddies are less dependent on the geometry, tend to be more
isotropic, and are consequently more universal.
The chance of finding a universal model is much higher when only
small eddies are modeled.
Solving only for the large eddies and modeling the smaller scales results
in mesh resolution requirements that are much less restrictive than with
DNS. Typically, mesh sizes can be at least one order of magnitude smaller
than with DNS. Furthermore, the time step sizes will be proportional to
the eddy-turnover time, which is much less restrictive than with DNS.
In practical terms, however, extremely fine meshes are still required. It
is only due to the explosive increases in computer hardware performance
coupled with the availability of parallel processing that LES can even be
considered as a possibility for engineering calculations.
The following sections give details of the governing equations for LES,
present the two options for modeling the subgrid-scale stresses (necessary
to achieve closure of the governing equations), and discuss the relevant
boundary conditions.

10-54

c Fluent Inc. November 28, 2001


10.7 The Large Eddy Simulation (LES) Model

10.7.1

Filtered Navier-Stokes Equations

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

(x) =

(x0 )G(x, x0 )dx0

(10.7-1)

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

1
V

Z
V

(x0 ) dx0 , x0 V

(10.7-2)

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

(10.7-3)

Since the application of LES to compressible flows is still in its infancy,


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

(ui ) = 0
+
t
xi

c Fluent Inc. November 28, 2001


(10.7-4)

10-55

Modeling Turbulence

and

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

ui

xj

p
ij

xi
xj

(10.7-5)

where ij is the subgrid-scale stress defined by


ij ui uj ui uj

(10.7-6)

The similarity between the filtered equations, 10.7-4 through 10.7-6, and
the incompressible form of the RANS equations, Equations 10.2-3 and
10.2-4, is obvious. The major difference is that the dependent variables
are now filtered quantities rather than mean quantities, and the expressions for the turbulent stresses differ.
10.7.2

Subgrid-Scale Models

The subgrid-scale stresses resulting from the filtering operation are unknown, and require modeling. The majority of subgrid-scale models in
use today are eddy viscosity models of the following form:
1
ij kk ij = 2t S ij
3

(10.7-7)

where t is the subgrid-scale turbulent viscosity, and S ij is the rate-ofstrain tensor for the resolved scale defined by

S ij

ui
uj
+
xj
xi

(10.7-8)

FLUENT contains two models for t : the Smagorinsky-Lilly model and


the RNG-based subgrid-scale model.

10-56

c Fluent Inc. November 28, 2001


10.7 The Large Eddy Simulation (LES) Model

Smagorinsky-Lilly Model
The most basic of subgrid-scale models was proposed by Smagorinsky [214] and further developed by Lilly [139]. In the Smagorinsky-Lilly
model, the eddy viscosity is modeled by

t = L2s S

(10.7-9)

where Ls is the mixing length for subgrid scales and S 2S ij S ij . Cs


is the Smagorinsky constant. In FLUENT, Ls is computed using


Ls = min d, Cs V 1/3

(10.7-10)

where is the von Karm


an constant, d is the distance to the closest wall,
and V is the volume of the computational cell.
Lilly derived a value of 0.23 for Cs from homogeneous isotropic turbulence in the inertial subrange. However, this value was found to cause
excessive damping of large-scale fluctuations in the presence of mean
shear or in transitional flows. Cs =0.1 has been found to yield the best
results for a wide range of flows, and is the default value in FLUENT.
RNG-Based Subgrid-Scale Model
Renormalization group (RNG) theory can be used to derive a model for
the subgrid-scale eddy viscosity [271]. The RNG procedure results in an
effective subgrid viscosity, eff = + t , given by
eff = [1 + H(x)]1/3

(10.7-11)

H(x) is the Heaviside function:


(

H(x) =

x, x > 0
0, x 0

(10.7-12)

where

c Fluent Inc. November 28, 2001


10-57

Modeling Turbulence

x=

2s eff
C
3

(10.7-13)

and
q

s = (Crng V 1/3 )2 2S ij S ij

(10.7-14)

where V is the volume of the computational cell. The theory gives Crng =
0.157 and C = 100.
In highly turbulent regions of the flow (t  ), eff s , and the
RNG-based subgrid-scale model reduces to the Smagorinsky-Lilly model
with a different model constant. In low-Reynolds-number regions of
the flow, the argument of the ramp function becomes negative and the
effective viscosity recovers molecular viscosity. This enables the RNGbased subgrid-scale eddy viscosity to model the low-Reynolds-number
effects encountered in transitional flows and near-wall regions.
10.7.3

Boundary Conditions for the LES Model

The stochastic components of the flow at the velocity-specified inlet


boundaries are accounted for by superposing random perturbations on
individual velocity components as
ui =< ui > +I |u|

(10.7-15)

where I is the intensity of q


the fluctuation, is a Gaussian random number satisfying = 0, and 0 = 1.
When the mesh is fine enough to resolve the laminar sublayer, the wall
shear stress is obtained from the laminar stress-strain relationship:
u
u y
=
u

10-58

(10.7-16)

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

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
= ln E
u

u y

(10.7-17)

where is the von Karm


an constant and E = 9.793.

10.8
10.8.1

Near-Wall Treatments for Wall-Bounded Turbulent


Flows
Overview

Turbulent flows are significantly affected by the presence of walls. Obviously, the mean velocity field is affected through the no-slip condition
that has to be satisfied at the wall. However, the turbulence is also
changed by the presence of the wall in non-trivial ways. Very close to
the wall, viscous damping reduces the tangential velocity fluctuations,
while kinematic blocking reduces the normal fluctuations. Toward the
outer part of the near-wall region, however, the turbulence is rapidly
augmented by the production of turbulence kinetic energy due to the
large gradients in mean velocity.
The near-wall modeling significantly impacts the fidelity of numerical solutions, inasmuch as walls are the main source of mean vorticity and turbulence. After all, it is in the near-wall region that the solution variables
have large gradients, and the momentum and other scalar transports occur most vigorously. Therefore, accurate representation of the flow in
the near-wall region determines successful predictions of wall-bounded
turbulent flows.
The k- models, the RSM, and the LES model are primarily valid for
turbulent core flows (i.e., the flow in the regions somewhat far from
walls). Consideration therefore 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.

c Fluent Inc. November 28, 2001


10-59

Modeling Turbulence

Numerous experiments have shown that the near-wall region can be


largely subdivided into three layers. In the innermost layer, called the
viscous sublayer, the flow is almost laminar, and the (molecular) viscosity plays a dominant role in momentum and heat or mass transfer. In
the outer layer, called the fully-turbulent layer, turbulence plays a major role. Finally, there is an interim region between the viscous sublayer
and the fully turbulent layer where the effects of molecular viscosity and
turbulence are equally important. Figure 10.8.1 illustrates these subdivisions of the near-wall region, plotted in semi-log coordinates.
U/U = 2.5 ln(U y/ ) + 5.45
inner layer

U/U

U/U = U y/
outer layer

viscous sublayer

buffer layer
or
blending
region

y+
5

fully turbulent region


or
log-law region

60
y+

Upper limit
depends on
Reynolds no.

ln U y/

Figure 10.8.1: Subdivisions of the Near-Wall Region

Wall Functions vs. Near-Wall Model


Traditionally, there are two approaches to modeling the near-wall region.
In one 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.

10-60

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

turbulent core

In another approach, the turbulence models are modified to enable the


viscosity-affected region to be resolved with a mesh all the way to the
wall, including the viscous sublayer. For purposes of discussion, this will
be termed the near-wall modeling approach. These two approaches
are depicted schematically in Figure 10.8.2.

buffer &
sublayer

Wall Function Approach

Near-Wall Model Approach

The viscosity-affected region is not


resolved, instead is bridged by the
wall functions.

The near-wall region is resolved


all the way down to the wall.

High-Re turbulence models can be


used.

The turbulence models ought to be valid


throughout the near-wall region.

Figure 10.8.2: Near-Wall Treatments in FLUENT

In most high-Reynolds-number flows, the wall function approach substantially saves 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-Reynolds-number 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 are valid in the
viscosity-affected region and accordingly integrable all the way to the
wall.

c Fluent Inc. November 28, 2001


10-61

Modeling Turbulence

FLUENT provides both the wall function approach and the near-wall
modeling approach.
Near-Wall Treatments for the Spalart-Allmaras, k-, and LES
Models
See Sections 10.3.6, 10.5.1, and 10.7.3, respectively, for a description of
the near-wall treatments applied by the Spalart-Allmaras, k-, and LES
models.
10.8.2

Wall Functions

Wall functions are a collection of semi-empirical formulas and functions


that in effect bridge or link the solution variables at the near-wall
cells and the corresponding quantities on the wall. The wall functions
comprise
laws-of-the-wall for mean velocity and temperature (or other scalars)
formulas for near-wall turbulent quantities
FLUENT offers two choices of wall function approaches:
standard wall functions
non-equilibrium wall functions
Standard Wall Functions
The standard wall functions in FLUENT are based on the proposal of
Launder and Spalding [129], 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

10-62

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

U =

1
ln(Ey )

(10.8-1)

where
U

1/4 1/2

UP C kP
w /
1/4 1/2

C kP yP
y

and

E
UP
kP
yP

=
=
=
=
=
=

(10.8-2)

(10.8-3)

von Karman constant (= 0.42)


empirical constant (= 9.81)
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 y >
about 30 to 60. In FLUENT, the log-law is employed when y > 11.225.
When the mesh is such that y < 11.225 at the wall-adjacent cells, FLUENT applies the laminar stress-strain relationship that can be written
as
U = y

(10.8-4)

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.
Energy
Reynolds analogy between momentum and energy transport gives a similar logarithmic law for mean temperature. As in the law-of-the-wall for

c Fluent Inc. November 28, 2001


10-63

Modeling Turbulence

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 nearwall 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 [246].
The law-of-the-wall implemented in FLUENT has the following composite
form:

1/4 1/2

(Tw TP ) cp C kP
q

1/4 1/2

+ 1 Pr C kP U 2

Pr
y

2
q i P
h

Pr 1 ln(Ey ) + P +
t

1 C1/4 kP1/2 

Prt UP2 + (Pr Prt )Uc2

q
2

(y < yT )
(10.8-5)
(y

>

yT )

where P is computed by using the formula given by Jayatilleke [104]:


"

P = 9.24

10-64

3/4

1 + 0.28e0.007/t

(10.8-6)

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

and
kf

cp
q
TP
Tw
Pr
Prt
A

E
Uc

=
=
=
=
=
=
=
=
=
=
=
=

thermal conductivity of fluid


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)
26 (Van Driest constant)
0.4187 (von Karman constant)
9.793 (wall function constant)
mean velocity magnitude at y = yT

Note that, for the segregated solver, the terms


1/4 1/2

1
C kP
Pr
UP2
2
q
and

o
1 C kP n
Prt UP2 + (Pr Prt )Uc2

2
q
will be included in Equation 10.8-5 only for compressible flow calculations.
1/4 1/2

The non-dimensional thermal sublayer thickness, yT , in Equation 10.8-5


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

c Fluent Inc. November 28, 2001


10-65

Modeling Turbulence

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

1/4 1/2

(Yi,w Yi ) C kP
Ji,w

Sc yh
i
Sct 1 ln(Ey ) + Pc

(y < yc )
(y > yc )

(10.8-7)

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

(10.8-8)

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

10-66

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

Gk w

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

(10.8-9)

and  is computed from


3/4 3/2

C kP
P =
yP

(10.8-10)

The  equation is not solved at the wall-adjacent cells, but instead is


computed using Equation 10.8-10.
Note that, as shown here, the wall boundary conditions for the solution
variables, 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.
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 [115] is also available. The key elements in the non-equilibrium
wall functions are as follows:

c Fluent Inc. November 28, 2001


10-67

Modeling Turbulence

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
1/4
1/4 k1/2
UC
C k1/2 y
1
= ln E
w /

(10.8-11)

where
"

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

y yv
y2
+ v
+

(10.8-12)

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


yv

yv

(10.8-13)

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 =

10-68

y
0, y < yv
yv
k=
k ,
w , y > y v
P

2

kP , y < yv
=
y > yv

2k
y2 ,
k 3/2
C` y ,

y < yv
y > yv
(10.8-14)

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

3/4

where C` = C , and yv is the dimensional thickness of the viscous


sublayer, defined in Equation 10.8-13.
Using these profiles, the cell-averaged production of k, Gk , and the cellaveraged 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 depthaverage,

Gk
=

yn
U
1
t
dy
yn 0
y
 
1
w2
yn
ln
1/4
1/2
yn C k
yv
P

(10.8-15)

and

 =

1
yn
1
yn

yn

 dy
0

"

1/2

2
yn
k
+ P ln
yv
C`
yv

#

kP

(10.8-16)

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

c Fluent Inc. November 28, 2001


10-69

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).
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 nonequilibrium 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 nearwall modeling approach combined with adequate mesh resolution in the

10-70

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

near-wall region. FLUENT provides the enhanced wall treatment for such
situations. This approach can be used with the three k- models and the
RSM.
10.8.3

Enhanced Wall Treatment

Enhanced wall treatment is a near-wall modeling method that combines a two-layer model with enhanced wall functions. If the near-wall
mesh is fine enough to be able to resolve the laminar sublayer (typically
y + 1), then the enhanced wall treatment will be identical to the traditional two-layer zonal model (see below for details). However, the restriction that the near-wall mesh must be sufficiently fine everywhere might
impose too large a computational requirement. Ideally, then, one would
like to have a near-wall formulation that can be used with coarse meshes
(usually referred to as wall-function meshes) as well as fine meshes (lowReynolds-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.
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

c Fluent Inc. November 28, 2001


10-71

Modeling Turbulence

y k
Rey

(10.8-17)

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

(10.8-18)

~
rw 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 10.4 and 10.6) are employed.
In the viscosity-affected near-wall region (Rey < Rey ), the one-equation
model of Wolfstein [269] is employed. In the one-equation model, the
momentum equations and the k equation are retained as described in
Sections 10.4 and 10.6. However, the turbulent viscosity, t , is computed
from

t,2layer = C ` k

(10.8-19)

where the length scale that appears in Equation 10.8-19 is computed


from [34]


` = yc` 1 eRey /A

(10.8-20)

The two-layer formulation for turbulent viscosity described above is used


as a part of the enhanced wall treatment, in which the two-layer definition
is smoothly blended with the high-Reynolds-number t definition from
the outer region, as proposed by Jongen [106]:

10-72

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

t,enh =  t + (1  )t,2layer

(10.8-21)

where t is the high-Reynolds-number definition as described in Section 10.4 or 10.6 for the k- models or the RSM. A blending function,
 , 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
"

1
 =
1 + tanh
2

Rey Rey
A

!#

(10.8-22)

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)

(10.8-23)

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

=

k3/2
`

(10.8-24)

The length scales that appear in Equation 10.8-24 are again computed
from Chen and Patel [34]:


` = yc` 1 eRey /A

(10.8-25)

If the whole flow domain is inside the viscosity-affected region


(Rey < 200),  is not obtained by solving the transport equation; it
is instead obtained algebraically from Equation 10.8-24. FLUENT uses

c Fluent Inc. November 28, 2001


10-73

Modeling Turbulence

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 10.8-20 and 10.8-25,
are taken from [34]:
c` = C3/4 , A = 70,

A = 2c`

(10.8-26)

Enhanced Wall Functions


To have a method that can extend its applicability throughout the nearwall 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 [108]:
1

u+ = e u+
lam + e uturb

(10.8-27)

where the blending function is given by:

a(y + )4
1 + by +


E
c = exp
1.0
E 00
a = 0.01c
5
b =
c

Similarly, the general equation for the derivative

(10.8-28)
(10.8-29)
(10.8-30)
(10.8-31)
du+
dy +

+
+
1 du
du+
dulam
turb

=
e
+
e
dy +
dy +
dy +

10-74

is
(10.8-32)

c Fluent Inc. November 28, 2001


10.8 Near-Wall Treatments for Wall-Bounded Turbulent Flows

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 [266] and Huang et al. [95]:
i
du+
1 h 0
+
+ 2 1/2
turb
=
S
(1

(u
)
)
dy +
y +
(

where
0

S =

1 + y + for y + < ys+


1 + ys+ for y + ys+

(10.8-33)

(10.8-34)

and


w dp
dp

= 2 3

w u dx
(u ) dx

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

(10.8-35)
(10.8-36)
(10.8-37)

where ys+ is the location at which the log-law slope will remain fixed. By
default, ys+ = 60. The coefficient in Equation 10.8-33 represents the
influences of pressure gradients while the coefficients and represent
thermal effects. Equation 10.8-33 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:

c Fluent Inc. November 28, 2001


10-75

Modeling Turbulence

du+
lam
= 1 + y +
dy +

(10.8-38)

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 10.8-38 results in

+
u+
1+
lam = y

+
y
2

(10.8-39)

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 [108]:

+
+
T + = e Tlam
+ e Tturb

(10.8-40)

where
=

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

(10.8-41)

where Pr is the molecular Prandtl number, and the coefficients a and b


are defined as in Equations 10.8-30 and 10.8-31. Apart from the above
formulation for T + , enhanced thermal wall functions follow the same
logic as previously described for standard thermal wall functions (see
Section 10.8.2). A similar procedure is also used for species wall functions
when the enhanced wall treatment is used. See Section 10.8.2 for details
about species wall functions.
The boundary condition for turbulence kinetic energy is the same as for
standard wall functions (Equation 10.8-8). However, the production of
turbulence kinetic energy Gk is computed using the velocity gradients
that are consistent with the enhanced law-of-the-wall (Equations 10.8-27
and 10.8-32), ensuring a formulation that is valid throughout the nearwall region.
10-76

c Fluent Inc. November 28, 2001


10.9 Grid Considerations for Turbulent Flow Simulations

10.9

Grid Considerations for Turbulent Flow Simulations

Successful computations of turbulent flows require some consideration


during the mesh generation. Since turbulence (through the spatiallyvarying effective viscosity) plays a dominant role in the transport of mean
momentum and other scalars for the majority of complex turbulent flows,
you must ascertain that turbulence quantities are properly resolved, if
high accuracy is required. Due to the strong interaction of the mean
flow and turbulence, the numerical results for turbulent flows tend to be
more susceptible to grid dependency than those for laminar flows.
It is therefore recommended that you resolve, with sufficiently fine meshes,
the regions where the mean flow changes rapidly and there are shear layers with a large mean rate of strain.
You can check the near-wall mesh by displaying or plotting the values of
y + , y , and Rey , which are all available in the postprocessing panels. It
should be remembered that y + , y , and Rey are not fixed, geometrical
quantities. They are all solution-dependent. For example, when you
double the mesh (thereby halving the wall distance), the new y + does
not necessarily become half of the y + for the original mesh.
For the mesh in the near-wall region, different strategies must be used
depending on which near-wall option you are using. In Sections 10.9.1
and 10.9.2 are general guidelines for the near-wall mesh.
10.9.1

Near-Wall Mesh Guidelines for Wall Functions

The distance from the wall at the wall-adjacent cells must be determined
by considering the range over which the log-law is valid. The distance
is usually measured in the wall unit, y + ( u y/), or y . Note that
y + and y have comparable values when the first cell is placed in the
log-layer.
It is known that the log-law is valid for y + > 30 to 60.
Although FLUENT employs the linear (laminar) law when y + <
11.225, using an excessively fine mesh near the walls should be
avoided, because the wall functions cease to be valid in the viscous
sublayer.

c Fluent Inc. November 28, 2001


10-77

Modeling Turbulence

The upper bound of the log-layer depends on, among others, pressure gradients and Reynolds number. As the Reynolds number
increases, the upper bound tends to also increase. y + values that
are too large are not desirable, because the wake component becomes substantially large above the log-layer.
A y + value close to the lower bound (y + 30) is most desirable.
Using excessive stretching in the direction normal to the wall should
be avoided.
It is important to have at least a few cells inside the boundary
layer.
10.9.2

Near-Wall Mesh Guidelines for the Enhanced Wall


Treatment

Although the enhanced wall treatment is designed to extend the validity of near-wall modeling beyond the viscous sublayer, it is still recommended that you construct a mesh that will fully resolve the viscosityaffected 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 nearwall region (Rey < 200) to be able to resolve the mean velocity
and turbulent quantities in that region.
10.9.3

Near-Wall Mesh Guidelines for the Spalart-Allmaras


Model

The Spalart-Allmaras model in its complete implementation is a lowReynolds-number model. This means that it is designed to be used with
10-78

c Fluent Inc. November 28, 2001


10.9 Grid Considerations for Turbulent Flow Simulations

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 10.9.2 for the enhanced wall treatment.
However, as discussed in Section 10.3.6, the boundary conditions for
the Spalart-Allmaras model have been implemented so that the model
will work on coarser meshes, such as would be appropriate for the wall
function approach. If you are using a coarse mesh, you should follow the
guidelines described in Section 10.9.1.
In summary, for best results with the Spalart-Allmaras model, you should
use either a very fine near-wall mesh spacing (on the order of y + = 1) or
a mesh spacing such that y + 30.
10.9.4

Near-Wall Mesh Guidelines for the k- Models

Both k- models available in FLUENT are available as low-Reynoldsnumber models as well as high-Reynolds-number models. If the Transitional Flows option is enabled in the Viscous Model panel, low-Reynoldsnumber variants will be used, and, in that case, mesh guidelines should
be the same as for the enhanced wall treatment. However, if this option
is not active, then the mesh guidelines should be the same as for the wall
functions.
10.9.5

Near-Wall Mesh Guidelines for Large Eddy Simulation

For the LES implementation in FLUENT, the wall boundary conditions


have been implemented using a law-of-the-wall approach as described in
Section 10.7.3. This means that there are no computational restrictions
on the near-wall mesh spacing. However, for best results, it might be
necessary to use a very fine near-wall mesh spacing (on the order of
y + = 1).

c Fluent Inc. November 28, 2001


10-79

Modeling Turbulence

10.10

Problem Setup for Turbulent Flows

When your FLUENT model includes turbulence you need to activate the
relevant model and options, and supply turbulent boundary conditions.
These inputs are described in this section.
The procedure for setting up a turbulent flow problem is described below.
(Note that this procedure includes only those steps necessary for the
turbulence model itself; you will need to set up other models, boundary
conditions, etc. as usual.)
1. To activate the turbulence model, select Spalart-Allmaras, k-epsilon,
k-omega, Reynolds Stress, or (in 3D) Large Eddy Simulation under
Model in the Viscous Model panel (Figure 10.10.1).
Define Models Viscous...
If you choose the k-epsilon model, select Standard, RNG, or Realizable under k-epsilon Model. If you choose the k-omega model, select
Standard or SST under k-omega Model.

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


2. If the flow involves walls, and you are using one of the k- models
or the RSM, choose one of the following options for the Near-Wall
Treatment in the Viscous Model panel:
Standard Wall Functions
Non-Equilibrium Wall Functions
Enhanced Wall Treatment
These near-wall options are described in detail in Section 10.8. By
default, the standard wall function is enabled.
The near-wall treatment for the Spalart-Allmaras, k-, and LES
models is defined automatically, as described in Sections 10.3.6,
10.5.1, and 10.7.3, respectively.
3. Enable the appropriate turbulence modeling options in the Viscous
Model panel. See Section 10.10.1 for details.

10-80

c Fluent Inc. November 28, 2001


10.10 Problem Setup for Turbulent Flows

Figure 10.10.1: The Viscous Model Panel

c Fluent Inc. November 28, 2001


10-81

Modeling Turbulence

4. Specify the boundary conditions for the solution variables.


Define Boundary Conditions...
See Section 10.10.2 for details.
5. Specify the initial guess for the solution variables.
Solve Initialize Initialize...
See Section 10.10.3 for details. Note that Reynolds stresses are
automatically initialized using k, and therefore need not be initialized.
10.10.1

Turbulence Options

The various options available for the turbulence models are described
in detail in Sections 10.3 through 10.7. Instructions for activating these
options are provided here.
If you choose the Spalart-Allmaras model, the following options are available:
Vorticity-based production
Strain/vorticity-based production
Viscous heating (always activated for the coupled solvers)
If you choose the standard k- model or the realizable k- model, the
following options are available:
Viscous heating (always activated for the coupled solvers)
Inclusion of buoyancy effects on 
If you choose the RNG k- model, the following options are available:
Differential viscosity model
Swirl modification
10-82

c Fluent Inc. November 28, 2001


10.10 Problem Setup for Turbulent Flows

Viscous heating (always activated for the coupled solvers)


Inclusion of buoyancy effects on 
If you choose the standard k- model, the following options are available:
Transitional flows
Shear flow corrections
Viscous heating (always activated for the coupled solvers)
If you choose the shear-stress transport k- model, the following options
are available:
Transitional flows
Viscous heating (always activated for the coupled solvers)
If you choose the RSM, the following options are available:
Wall reflection effects on Reynolds stresses
Wall boundary conditions for the Reynolds stresses from the k
equation
Quadratic pressure-strain model
Viscous heating (always activated for the coupled solvers)
Inclusion of buoyancy effects on 
If you choose the enhanced wall treatment (available for the k- models
and the RSM), the following options are available:
Pressure gradient effects
Thermal effects
c Fluent Inc. November 28, 2001

10-83

Modeling Turbulence

If you choose the LES model, the following options are available:
Smagorinsky-Lilly model for the subgrid-scale viscosity
RNG model for the subgrid-scale viscosity
Viscous heating (always activated for the coupled solvers)
It is also possible to modify the Model Constants, but this is not necessary for most applications. See Sections 10.3 through 10.7 for details
about these constants. Note that C1-PS and C2-PS are the constants C1
and C2 in the linear pressure-strain approximation of Equations 10.6-5
and 10.6-6, and C1-PS and C2-PS are the constants C10 and C20 in
Equation 10.6-7. C1-SSG-PS, C1-SSG-PS, C2-SSG-PS, C3-SSG-PS, C3SSG-PS, C4-SSG-PS, and C5-SSG-PS are the constants C1 , C1 , C2 , C3 ,
C3 , C4 , and C5 in the quadratic pressure-strain approximation of Equation 10.6-16.
Including the Viscous Heating Effects
See Sections 11.2.1 and 11.2.2 for information on including viscous heating effects in your model.
Including Turbulence Generation Due to Buoyancy
If you specify a non-zero gravity force (in the Operating Conditions panel),
and you are modeling a non-isothermal flow, the generation of turbulent
kinetic energy due to buoyancy (Gb in Equation 10.4-1) is, by default, always included in the k equation. However, FLUENT does not, by default,
include the buoyancy effects on .
To include the buoyancy effects on , you must turn on the Full Buoyancy
Effects option under Options in the Viscous Model panel.
This option is available for the three k- models and for the RSM.

10-84

c Fluent Inc. November 28, 2001


10.10 Problem Setup for Turbulent Flows

Vorticity- and Strain/Vorticity-Based Production


For the Spalart-Allmaras model, you can choose either Vorticity-Based
Production or Strain/Vorticity-Based Production under Spalart-Allmaras
Options in the Viscous Model panel. If you choose Vorticity-Based Production, FLUENT will use Equation 10.3-8 to compute the value of the
deformation tensor S; if you choose Strain/Vorticity-Based Production, it
will use Equation 10.3-10.
(These options will not appear unless you have activated the SpalartAllmaras model.)
Differential Viscosity Modification
In the RNG turbulence model in FLUENT, you have an option to use
a differential formula for effective viscosity eff (Equation 10.4-6) to
account for low-Reynolds-number effects. To enable this option, turn
on Differential Viscosity Model under RNG Options in the Viscous Model
panel.
(This option will not appear unless you have activated the RNG k-
model.)
Swirl Modification
Once you choose the RNG model, the swirl modification takes effect,
by default, for all three-dimensional flows and axisymmetric flows with
swirl. The default swirl constant (s in Equation 10.4-8) is set to 0.05,
which works well for weakly to moderately swirling flows. However, for
strongly swirling flows, you may need to use a larger swirl constant.
In order to change the value of the swirl constant, you must first turn
on the Swirl Dominated Flow option under RNG Options in the Viscous
Model panel. (This option will not appear unless you have activated the
RNG k- model.)
Once you turn on the Swirl Dominated Flow option, the swirl constant
s is increased to 0.07. You can change its value in the Swirl Factor field
under Model Constants.

c Fluent Inc. November 28, 2001


10-85

Modeling Turbulence

Transitional Flows
If either of the k- models are used, you may enable a low-Reynoldsnumber 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 10.5-6) is equal to 1.
Shear Flow Corrections
In the standard k- model, you also have the option of including corrections to improve the accuracy in predicting free shear 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 [267]. When this option is enabled, FLUENT will calculate f
and f using Equations 10.5-16 and 10.5-24, respectively. If this option
is disabled, f and f will be set equal to 1.
Including Pressure Gradient Effects
If the enhanced wall treatment is used, you may include the 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 10.8-33.
Including Thermal Effects
If the enhanced wall treatment is used, you may include thermal effects
by enabling the Thermal Effects option under Enhanced Wall Treatment
Options. When this option is enabled, FLUENT will include the coefficient in Equation 10.8-33. will also be included in Equation 10.8-33
when the Thermal Effects option is enabled if the ideal gas law is selected
for the fluid density in the Materials panel.
Including the Wall Reflection Term
If the RSM is used with the default model for pressure strain, FLUENT
will, by default, include the wall-reflection effects in the pressure-strain
term. That is, FLUENT will calculate w
ij using Equation 10.6-7 and
10-86

c Fluent Inc. November 28, 2001


10.10 Problem Setup for Turbulent Flows

include it in Equation 10.6-4. Note that wall-reflection effects are not


included if you have selected the quadratic pressure-strain model.

! The empirical constants and the function f used in the calculation of

w
ij are calibrated for simple canonical flows such as channel flows and
flat-plate boundary layers involving a single wall. If the flow involves
multiple walls and the wall has significant curvature (e.g., an axisymmetric pipe or curvilinear duct), the inclusion of the wall-reflection term
in Equation 10.6-7 may not improve the accuracy of the RSM predictions. In such cases, you can disable the wall-reflection effects by turning
off the Wall Reflection Effects under Reynolds-Stress Options in the Viscous
Model panel.

Solving the k Equation to Obtain Wall Boundary Conditions


In the RSM, FLUENT, by default, uses the explicit setting of boundary
conditions for the Reynolds stresses near the walls, with the values computed with Equation 10.6-28. k is calculated by solving the k equation
obtained by summing Equation 10.6-1 for normal stresses. To disable this
option and use the wall boundary conditions given in Equation 10.6-29,
turn off Wall B.C. from k Equation under Reynolds-Stress Options in the
Viscous Model panel. (This option will not appear unless you have activated the RSM.)
Quadratic Pressure-Strain Model
To use the quadratic pressure-strain model described in Section 10.6.3,turn
on the Quadratic Pressure-Strain Model option under Reynolds-Stress Options in the Viscous Model panel. (This option will not appear unless you
have activated the RSM.) The following options are not available when
the Quadratic Pressure-Strain Model is enabled:
Wall Reflection Effects under Reynolds-Stress Options
Enhanced Wall Treatment under Near-Wall Treatment

c Fluent Inc. November 28, 2001


10-87

Modeling Turbulence

Subgrid-Scale Model
If you have selected the Large Eddy Simulation model, you will be able to
choose which of the two subgrid-scale models described in Section 10.7.2
is to be used. You can choose either the Smagorinsky-Lilly or the RNG
subgrid-scale model.
(These options will not appear unless you have activated the LES model.)
Customizing the Turbulent Viscosity
If you are using the Spalart-Allmaras, k-, k-, or LES model, a userdefined 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 user-defined function in the Turbulent Viscosity drop-down list.
10.10.2

Defining Turbulence Boundary Conditions

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 6.2.2 for details about specifying the boundary conditions
for k and  (or k and ) at the inlets.

10-88

c Fluent Inc. November 28, 2001


10.10 Problem Setup for Turbulent Flows

You may want to include the effects of the wall roughness on selected
wall boundaries. In such cases, you can specify the roughness parameters (roughness height and roughness constant) in the panels for the
corresponding wall boundaries (see Section 6.13.1).
The Spalart-Allmaras Model
When you are modeling turbulent flows in FLUENT using the SpalartAllmaras 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 6.2.2 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 6.13.1).
Reynolds Stress Model
The specification of turbulent boundary conditions for the RSM is the
same as for the other turbulence models for all boundaries except at
boundaries where flow enters the domain. Additional input methods are
available for these boundaries and are described here.
When you choose to use the RSM, the default inlet boundary condition
inputs required are identical to those required when the k- model is
active. You can input the turbulence quantities using any of the turbulence specification methods described in Section 6.2.2. FLUENT then
uses the specified turbulence quantities to derive the Reynolds stresses
at the inlet from the assumption of isotropy of turbulence:

c Fluent Inc. November 28, 2001


10-89

Modeling Turbulence

2
k
3
= 0.0

ui2 =
u0i u0j

(i = 1, 2, 3)

(10.10-1)
(10.10-2)

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 6.2.2). To use this method, you
will select K or Turbulence Intensity as the Reynolds-Stress Specification
Method in the appropriate boundary condition panel.
Alternately, you can directly specify the Reynolds stresses by selecting
Reynolds-Stress Components as the Reynolds-Stress Specification Method
in the boundary condition panel. When this option is enabled, you
should input the Reynolds stresses directly.
You can set the Reynolds stresses by using constant values, profile functions of coordinates (see Section 6.25), or user-defined functions (see the
separate UDF Manual).
Large Eddy Simulation Model
It is possible to specify the magnitude of random fluctuations of the
velocity components at an inlet only if the velocity inlet boundary condition is selected. In this case, you must specify a Turbulence Intensity
that determines the magnitude of the random perturbations on individual mean velocity components as described in Section 10.7.3. For all
boundary types other than velocity inlets, the boundary conditions for
LES remain the same as for laminar flows.
10.10.3

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

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 ).

10-90

c Fluent Inc. November 28, 2001


10.10 Problem Setup for Turbulent Flows

Figure 10.10.2: Specifying Inlet Boundary Conditions for the Reynolds


Stresses

c Fluent Inc. November 28, 2001


10-91

Modeling Turbulence

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 22.13 for details.)
For more complex flows (e.g., flows with multiple inlets with different conditions) it may be better to specify the initial values in
terms of turbulence intensity. 510% is enough to represent fullydeveloped turbulence. k can then be computed from the turbulence
intensity and the characteristic mean velocity magnitude of your
problem (k = 1.5(Iuavg )2 ).
You should specify an initial guess for  so that the resulting eddy
2
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 10.10-1 and 10.10-2.

10-92

c Fluent Inc. November 28, 2001


10.11 Solution Strategies for Turbulent Flow Simulations

10.11

Solution Strategies for Turbulent Flow Simulations

Compared to laminar flows, simulations of turbulent flows are more challenging in many ways. For the Reynolds-averaged approach, additional
equations are solved for the 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.
10.11.1

Mesh Generation

The following are suggestions to follow when generating the mesh for use
in your turbulent flow simulation:
Picture in your mind the flow under consideration using your physical intuition or any data for a similar flow situation, and identify
the main flow features expected in the flow you want to model.
Generate a mesh that can resolve the major features that you expect.
If the flow is wall-bounded, and the wall is expected to significantly
affect the flow, take additional care when generating the mesh. You
should avoid using a mesh that is too fine (for the wall function approach) or too coarse (for the enhanced wall treatment approach).
See Section 10.9 for details.
10.11.2

Accuracy

The suggestions below are provided to help you obtain better accuracy
in your results:

c Fluent Inc. November 28, 2001


10-93

Modeling Turbulence

Use the turbulence model that is better suited for the salient features you expect to see in the flow (see Section 10.2).
Because the mean quantities have larger gradients in turbulent
flows than in laminar flows, it is recommended that you use highorder 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.
10.11.3

Convergence

The suggestions below are provided to help you enhance convergence for
turbulent flow calculations:
Starting with excessively crude initial guesses for mean and turbulence quantities may cause the solution to diverge. A safe approach is to start your calculation using conservative (small) underrelaxation parameters and (for the coupled solvers) a conservative
Courant number, and increase them gradually as the iterations
proceed and the solution begins to settle down.
It is also helpful for faster convergence to start with reasonable initial guesses for the k and  (or k and ) fields. Particularly when
the enhanced wall treatment is used, it is important to start with
a sufficiently developed turbulence field, as recommended in Section 10.10.3, to avoid the need for an excessive number of iterations
to develop the turbulence field.
When you are using the RNG k- model, an approach that might
help you achieve better convergence is to obtain a solution with
the standard k- model before switching to the RNG model. Due

10-94

c Fluent Inc. November 28, 2001


10.11 Solution Strategies for Turbulent Flow Simulations

to the additional non-linearities in the RNG model, lower underrelaxation factors and (for the coupled solvers) a lower Courant
number might also be necessary.
Note that when you use the enhanced wall treatment, you may sometimes find during the calculation that the residual for  is reported to
be zero. This happens when your flow is such that Rey is less than 200
in the entire flow domain, and  is obtained from the algebraic formula
(Equation 10.8-24) instead of from its transport equation.
10.11.4

RSM-Specific Solution Strategies

Using the RSM creates a high degree of coupling between the momentum
equations and the turbulent stresses in the flow, and thus the calculation
can be more prone to stability and convergence difficulties than with the
k- models. When you use the RSM, therefore, you may need to adopt
special solution strategies in order to obtain a converged solution. The
following strategies are generally recommended:
Begin the calculations using the standard k- model. Turn on the
RSM and use the k- solution data as a starting point for the RSM
calculation.
Use low under-relaxation factors (0.2 to 0.3) and (for the coupled
solvers) a low Courant number for highly swirling flows or highly
complex flows. In these cases, you may need to reduce the underrelaxation factors both for the velocities and for all of the stresses.
Instructions for setting these solution parameters are provided below. If
you are applying the RSM to prediction of a highly swirling flow, you
will want to consider the solution strategies discussed in Section 8.4 as
well.
Under-Relaxation of the Reynolds Stresses
FLUENT applies under-relaxation to the Reynolds stresses. You can set
under-relaxation factors using the Solution Controls panel.

c Fluent Inc. November 28, 2001


10-95

Modeling Turbulence

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 practicallyconverged solution. However, you may need to apply tighter convergence
criteria (below 104 ) to ensure full convergence.
10.11.5

LES-Specific Solution Strategies

Large eddy simulation involves running a transient solution from some


initial condition, on an appropriately fine grid, using an appropriate time
step size. The solution must be run long enough to become independent
of the initial condition and to enable the statistics of the flow field to be
determined.
The following are suggestions to follow when running a large eddy simulation:
1. Start by running a flow simulation assuming laminar flow or using
a simple Reynolds-averaged turbulence model such as standard k or Spalart-Allmaras. Since this is only an initial condition, you
need run only until the flow field is somewhat converged. This step
is optional.
10-96

c Fluent Inc. November 28, 2001


10.11 Solution Strategies for Turbulent Flow Simulations

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 22.15.1 for guidelines on setting solution parameters for transient calculations in
general.) Use the central-differencing spatial discretization scheme
for all equations.
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 22.15.1.
5. Continue until you get statistically stable data. The duration of
the simulation can be determined beforehand by estimating the
mean flow residence time in the solution domain (L/U , where L is
the characteristic length of the solution domain and U is a characteristic mean flow velocity). The simulation should be run for at
least a few mean flow residence times.
Instructions for setting the solution parameters for LES are provided
below.
Temporal Discretization
FLUENT provides both first-order and second-order temporal discretizations. For LES, the second-order discretization is recommended.
Define Models Solver...
Spatial Discretization
Overly diffusive schemes such as the first-order upwind or power law
scheme should be avoided, because they may unduly damp out the energy

c Fluent Inc. November 28, 2001


10-97

Modeling Turbulence

of the resolved eddies. The central-differencing scheme is recommended


for all equations when you use the LES model.
Solve Controls Solution...

10.12

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
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:

10-98

c Fluent Inc. November 28, 2001


10.12 Postprocessing for Turbulent Flows

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
Turbulence quantities that can be reported for the RSM are as follows:
Turbulent Kinetic Energy (k)
c Fluent Inc. November 28, 2001

10-99

Modeling Turbulence

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 LES model are as
follows:
Subgrid Turbulent Kinetic Energy
Subgrid Turbulent Viscosity
Subgrid Effective Viscosity

10-100

c Fluent Inc. November 28, 2001


10.12 Postprocessing for Turbulent Flows

Subgrid Turbulent Viscosity Ratio


Effective Thermal Conductivity
Effective Prandtl Number
Wall Yplus
All of these variables can be found in the Turbulence... category of the
variable selection drop-down list that appears in postprocessing panels.
See Chapter 27 for their definitions.
10.12.1

Custom Field Functions for Turbulence

In addition to the quantities listed above, you can define your own turbulence quantities using the Custom Field Function Calculator panel.
Define Custom Field Functions...
The following functions may be useful:
Ratio of production of k to its dissipation (Gk /)
Ratio of the mean flow to turbulent time scale, ( Sk/)
Reynolds stresses derived from the Boussinesq formula (e.g., uv =
t u
y )
10.12.2

Postprocessing LES Statistics

As described in Section 10.7, LES involves the solution of a transient


flow field, but it is the mean flow quantities that are of most engineering
interest. If you turn on the Data Sampling for Time Statistics option
in the Iterate panel, FLUENT will gather data for time statistics while
performing a large eddy simulation. You can then view both the mean
and the root-mean-square (RMS) values in FLUENT. See Section 22.15.3
for details.

c Fluent Inc. November 28, 2001


10-101

Modeling Turbulence

10.12.3

Troubleshooting

You can use the postprocessing options not only for the purpose of interpreting your results but also for investigating any anomalies that may
appear in the solution. For instance, you may want to plot contours of
the k field to check if there are any regions where k is erroneously large
or small. You should see a high k region in the region where the production of k is large. You may want to display the turbulent viscosity ratio
field in order to see whether or not turbulence takes full effect. Usually turbulent viscosity is at least two orders of magnitude larger than
molecular viscosity for fully-developed turbulent flows modeled using the
RANS approach (i.e., not using LES). You may also want to see whether
you are using a proper near-wall mesh for the enhanced wall treatment.
In this case, you can display filled contours of Rey (turbulent Reynolds
number) overlaid on the mesh.

10-102

c Fluent Inc. November 28, 2001

You might also like