Turbulence Models Commonly Used in CFD
Turbulence Models Commonly Used in CFD
5,400
Open access books available
132,000
International authors and editors
160M Downloads
154
Countries delivered to
TOP 1%
most cited scientists
12.2%
Contributors from top 500 universities
Abstract
1. Introduction
1
Computational Fluid Dynamics
modeling into the CFD solutions. The first turbulence models accounted for turbu-
lence effects through a concept termed the “eddy viscosity”. Essentially, the eddy
viscosity (or turbulent viscosity) reflects an apparent increase in viscosity caused by
small-scale chaotic motions in a fluid. The simulations do not attempt to actually
capture small scale turbulent motions, rather they approximate their effect with an
increase in the fluid viscosity. As we will discuss, the concept of turbulence viscos-
ity plays a central role in Reynolds Averaged Navier Stokes (RANS) models. As we
will also show, other approaches do not rely extensively on the turbulent viscosity
concept.
The first turbulent viscosity “eddy viscosity” models were developed in the
1960s and are classified as algebraic [2, 3], one-Equation [4], or two-Equation [5–7].
The basis for two equation models was the relationship between the turbulent
viscosity and local values of the turbulent kinetic energy k and turbulent
dissipation, ε. Since this approach soon became the dominant method (even for
today), it is worthwhile to discuss it in some detail. In essence, this group of
turbulence models neglect small scale and rapid turbulent motions and use an
average flow field (timewise average values in the velocities and pressure values) to
estimate the effects of turbulence.
The first major effort to simulate turbulence in the context of CFD was the so-
called k-ε model [5, 6]. This approach utilizes the fluctuating components of the
turbulent velocity in the three coordinate directions to obtain a turbulent kinetic
energy, from:
1 02 02 02
k¼ u þv þw (1)
2
That is, k is the additional turbulent energy that results from the time-
fluctuating turbulent motions. Accompanying the turbulent kinetic energy is a
turbulent dissipation ε which can be calculated as
κ 3=2
ε¼ (2)
0:3D
for flows in pipes with diameter D [7, 8]. The connection of turbulence kinetic
energy and turbulent dissipation will be provided, following the equations of
motion. In essence, the governing equations of motion are conservation of mass,
which under steady conditions is:
∂ui
¼0 (3)
∂xi
2
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
∂ðui kÞ ∂ μt ∂k
ρ ¼ Pk þ Pb ρε þ μþ (5)
∂xi ∂xi σ k ∂xi
ε2
∂ðui εÞ ∂ μt ∂ε ε
ρ ¼ μþ þ C1 ðPk þ C3 Pb Þ C2 ρ (6)
∂xi ∂xi σ ε ∂xi k k
k2
μt ¼ ρCμ (7)
ε
The Pk is the production of turbulent kinetic energy from the shear strain rate
and Pb is the production of turbulent kinetic energy from buoyancy effects. The
production of turbulent kinetic energy is obtained from the time-averaged velocity
field from:
∂ui ∂uj ∂ui ∂uκ ∂uκ
Pκ ¼ μt þ 3μt þ ρκ (8)
∂xj ∂xi ∂xj ∂xκ ∂xκ
The σ terms are corresponding Prandtl numbers for the transported variables.
The values of the constants and turbulent Prandtl numbers are specific to a partic-
ular k-ε model. The k-ε approach is likely the most widely used turbulent model,
even today. It is generally sufficient for flows that are wall bounded, with limited
adverse pressure gradients or separation.
Traditionally, the elements are not used to capture steep velocity and
temperature gradients near the wall. Rather, wall functions are employed to inter-
polate to the wall. Of course, the accuracy of this approach depends on the suitabil-
ity of a particular wall function to a problem. For example, wall functions often fail
when the flow experiences adverse pressure gradients and/or separation. On the
other hand, when small elements are deployed near the wall and/or when damping
equations are used to limit fluid motion in the boundary layer, integration can be
performed up to the wall. In our experience, if integration is to be performed up to
the wall (and wall function interpolation is avoided), the near-wall element should
have a size of y+1 for models that resolve the boundary layer. This guidance is not
used for models that use the law-of-the-wall to interpolate to the wall.
A popular modification of the traditional k-ε model is the RNG
(Renormalization Group) model. It was developed by [9] in an effort to handle
small flow phenomenon. The mechanism of multiple scale motions is achieved by
modifying the turbulent dissipation equation production term. In our experience, it
has somewhat better performance than the standard k-ε particularly for rotating
flows. The differences between the RNG and standard models is in the relationship
between the turbulent kinetic energy, turbulent dissipation, and turbulent viscosity.
With the RNG approach the turbulent viscosity is found from:
κ2
μt ¼ CμRNG ρ (9)
ε
3
Computational Fluid Dynamics
η
η 1 4:38
Cε1RNG ¼ 1:42 (11)
ð1 þ βRNG η3 Þ
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Pκ
η¼ (12)
ρCμRNG ε
While the k-ε model has experienced success in computational modeling, it has
deficiencies in some situations. In particular, the k-ε model performs suitably away
from walls, in the main flow. However, it has issues in the boundary layer zone,
particularly with low Reynolds numbers. Here, Reynolds numbers refer to local
Reynolds numbers that decrease as one moves closer to the wall and the no-slip
condition exerts its influence (rather than to the Reynolds number based on mac-
roscopic dimensions such a pipe diameter or plate length).
A significant development in CFD was brought forward by the development of
k-ω model that replaced the transport equation for ε with a specific rate of turbu-
lence dissipation, ω [10]. The new equations are:
∂ðui kÞ ∂ μt ∂k
ρ ¼ ρPk þ ρPb ρβωk þ μþ (13)
∂xi ∂xi σ k ∂xi
∂ðui ωÞ ∂ μ ∂ω αω
ρ ¼ μþ t þ Pk βρω2 (14)
∂xi ∂xi σ ω ∂xi k
Recognizing that the k-ε and k-ω model each have strengths and weaknesses, a
new model was proposed that uses both of these approaches in a way that harnesses
their strengths [11]. This new approach, termed the Shear Stress Transport model
(SST), smoothly transitions from the k-ω model near the wall to the k-ε model in
the main flow. With the SST model, the governing equation for turbulent dissipa-
tion is recast into an ω form. The governing equations are:
∂ðρui kÞ ∂ μt ∂k
¼ Pk β1 ρkω þ μþ (16)
∂xi ∂xi σ k ∂xi
∂ðρui ωÞ ω 2 ∂ μt ∂ω 1 ∂k ∂ω
¼ α3 Pκ β2 ρω þ μþ þ 2ð1 F 1 Þρ (17)
∂xi κ ∂xi σ ω ∂xi σ ω2 ω ∂xi ∂xi
aρk
μt ¼ (18)
max ðaω, SF 2 Þ
4
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
away from the wall from the wall. The S term is the magnitude of the shear strain
rate.
While ostensibly, the SST model is used for fully turbulent flows, it has shown
ability to capture both laminar and turbulent flow regimes [12]. However, in the
next section we discuss a set of modifications to the SST models that are specifically
designed to handled laminar/transitional/turbulent flow regimes that are
recommended.
The already discussed turbulent models were largely developed based on corre-
lations of canonical fully turbulent flow situations (such as flows over flat plates,
airfoils, Falkner-Skans flows, and flows in tubes and ducts). Of course, researchers
and engineers often experience situations where the flow is partially turbulent or
other situations where the flow changes so that for part of the time it is laminar and
other times turbulent. Consider for example pulsatile flow wherein the fluid veloc-
ity changes sufficiently so that for parts of the flow period, different flow regimes
occur. There are a number of approaches to handle these situations but with respect
to the RANS models, the approaches generally utilize the concept of turbulent
intermittency. Intermittency was originally defined as the percentage of time that a
flow was turbulent. However, more recently, turbulent intermittency has been used
as a multiplier on the rate of turbulent kinetic production [13–15].
Here we will set forth two current transitional models, both based on the SST
turbulence approach. The first method involves two extra transport equations. One
for the intermittency, γ, which is a multiplier to the turbulent production. The
transport equation for turbulent intermittency is:
∂ðργ Þ ∂ðρui γ Þ ∂ μt ∂γ
þ ¼ Pγ,1 Eγ,1 þ Pγ,2 Eγ,2 þ μþ (19)
∂t ∂xi ∂xi σ γ ∂xi
Together, solution to Eqs. (19) and (20) determine the local state of turbulence.
They result in an intermittency that takes values between 0 and 1. For fully
laminar flow, γ = 0 and the model reverts to a laminar solver. When γ = 1, the flow
is fully turbulent. The turbulent production then is then multiplied by the local
value of the intermittency, γ. Interested readers are invited to review the develop-
ment of this model, including implementation for problems that involve heat
transfer [16–22].
Recently, the above two-equation model was modified to reduce the two transi-
tional transport equations to a single Equation [23] and that approach was later
adapted by [24] to accurately solve for situations in confined pipe/duct/tube flows.
Essentially, Eqs. (19) and (20) are replaced by a single intermittency equation
which is:
∂ðρui γ Þ ∂ μt ∂γ
¼ Pγ Eγ þ μþ (21)
∂xi ∂xi σ γ ∂xi
5
Computational Fluid Dynamics
As we have already noted, the term S is the shear strain rate. A new term that
appears in Eq. (23) is the so-called onset transition term (Fonset) which is calculated
using the following set of equations.
" 3 ! #
ReV RT
Fonset ¼ MAX min ,2 max 1 ,0 ,0 (24)
2:2 Re θc 3:5
We have already noted that these transitional turbulence models were initially
developed for external boundary layer flows (flat plate boundary layers, airfoil
flows, Falkner-Skans flows, etc.). Insofar as we have adopted them for internal
flow, some modification was required. We recommend, at least for flows through
pipes, tubes, and ducts, that the initial constants determined in [23] be replaced by
alternative values from [24].
While we recommend the above approach for solving transitional flow problems,
this area of research is also heavily studied by other researchers who have provided
alternative approaches to handle such flows. We cite them here for readers who are
interested in those alternative but complementary viewpoints [25–33].
Reynolds stress models (RSM) are quite different from the RANS approach that
was just discussed. For RSMs, transport equations are used for all components of the
Reynolds stress tensor and an eddy viscosity is not utilized. These models are
expected to be superior for situations with non-isotropic turbulence and flows with
6
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
∂p0
∂u j ∂ ∂ ∂u j ∂ui ∂
ρ þρ ui u j ¼ þμ þ ρ ui u j þ B j (30)
∂t ∂xi ∂x j ∂xi ∂xi ∂xi ∂xi
The second-to-last term on the right-hand side represents the Reynolds stresses.
There is a pseudo-pressure term p’ that is calculated from the local static pressure p
and local velocity gradient from the following expression.
2 ∂uk
p0 ¼ p þ μ (31)
3 ∂xk
The Reynolds stresses are calculated by a collection of six equations for all
directional possibilities. The transport equations for Reynolds stresses are:
" ! #
2
∂u j ui ∂ ∂ 2 k ∂ui u j 2
ρ þρ uk ui u j μ þ ρCs ¼ Pij δij ρε þ Φij þ Pij,b
∂t ∂xk ∂xk 3 ε ∂xk 3
(32)
We note that a turbulence dissipation term, ε, appears in Eq. (32) and it has to be
solved from its own transport equation. We refer readers to [34, 35] for more
details.
A modification to the above is realized from the Baseline RSM (BSL RSM)
model. It differs from the SSG RSM in that the transport equation for ε is replaced
by a transport equation for ω. The new equation is:
∂ω ∂ðuk ωÞ ∂ μt ∂ω ω 2ρ 1 ∂k ∂ω
ρ þρ ¼ μþ þ α 3 Pk β3 ρω2 þ ð1 F1 Þ
∂t ∂xk ∂xk σ ω3 ∂xk k σ ω2 ω ∂xk ∂xk
þ Pωb
(33)
This approach blends between two different models that are used near the wall
and alternatively away from the wall. The modeling is accomplished using a
weighting function, similar to the SST:
ϕ3 ¼ F 1 ϕ1 þ ð 1 F1 Þ ϕ2 (34)
Where the symbols ϕ correspond to any particular transport variable in the near
wall and far wall regions. Various constants change their values in the two regions,
so that:
The constants near the wall:
The last RSM version to be discussed is the Explicit Algebraic RSM (EARSM).
This approach includes a non-linear relationship between the local values of the
Reynolds stresses and the vorticity tensors. It is focused on flows with secondary
7
Computational Fluid Dynamics
motions and curvature [36]. The local values of the Reynolds stresses are calculated
using an anisotropy tensor which is based on algebraic equations [36]. This is
contrasted with RSM approaches that solve for the Reynolds stress components
using differential transport equations. The approach is to use higher order terms for
many of the flow phenomena. It was designed to handle secondary flow situations
and flows with extensive curvature and rotation. The governing equations are
complex and lengthy and for brevity sake, we refer interested readers to [36].
One of the primary decisions that models are faced with is whether to perform
calculations in steady or unsteady mode. Typically with numerical simulation,
unsteadiness is driven by either timewise changes in boundary conditions or it is
related to unsteady phenomena that occur in an otherwise steady scenario. A classic
example is the Karmen Vortex Street that occurs in a wake region of a blunt object.
Figure 1, shown below, illustrates this phenomenon.
Researchers have often conjectured that if a RANS model is performed with
sufficiently small elements and time steps, the unsteady features of the flow would
naturally be resolved. But in fact, this is not true. It is important to note that steady
state calculations using RANS models will often provide very accurate information
8
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
Figure 1.
Unsteady wake region, even though oncoming flow is steady state.
about averaged quantities (like drag), these simulations will miss details in the
rapidly fluctuating downstream wake region. This issue was explored in depth
in [35] where time-averaged results of drag obtained from unsteady RANS
simulations were compared with calculations from steady RANS calculations
(using the SST transitional model that was previously described). It was found
that the steady state calculations were able to accurately capture drag forces
but were only partially adept at capturing vortex movement in the downstream
wake region.
With this discussion as background, it is now time to turn attention to the
governing equations of scale-adaptive RANS models. The model to be discussed
here uses the SST approach for the underlying governing equations (in the
literature it is often termed the SAS-SST model). The scale-adaptive approach
modifies the ω transport equation based on [37]. In particular, a new transport
equation is presented that incorporates the turbulent length scale L and is set
forth here:
pffiffiffi
Φ¼ kL (37)
and
2 !
∂Φ ∂U i Φ ΦPk Lt ∂ μt ∂Φ
þρ ¼ C1 C2 ρC3 k þ (38)
∂t ∂xi k Lk ∂xi σ Φ ∂xi
Values of the various constants can be found in [34, 37] and are not repeated
here for brevity. The term Lt is a novel modification; it refers to the von Karmen
length scale. Figures 2 and 3 are provided that show a comparison of downstream
wake regions for an unsteady RANS calculation using the SST model (Figure 2) and
a simulation using the scale-adaptive SST modification. Results are obtained from
[34]. It can be seen that the standard SST model does capture a periodic release of
eddies from the downstream side of a circular cylinder (shown in blue). In both
images, the flow is left-to-right. The color legend is keyed to the local values of the
turbulent length scale. Clearly the scale-adaptive approach provides a much wider
range of turbulent eddy sizes.
9
Computational Fluid Dynamics
Figure 2.
Calculations of turbulent length scale for flow over a circular cylinder, based on an unsteady SST model.
Figure 3.
Calculations of turbulent length scale for flow over a circular cylinder, based on a scale-adaptive unsteady SST
model.
∂ρui u j
∂ ∂p ∂ ∂ui ∂u j ∂τij
ðρui Þ þ ¼ þ μ þ þ (39)
∂t ∂xi ∂x j ∂xi ∂x j ∂xi ∂xi
10
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
1
τij ¼ ρui u j þ ρui u j ¼ 2μsgs Sij þ δij τkk (40)
3
And the Sij term indicates the strain rate tensor for large scale motions. The
small-scale eddy viscosity μsgs is found from
3=2
Sdij Sdij
μsgs ¼ ρðCw ΔÞ2 5=4 (41)
5=2
Sij Sij þ Sdij Sdij
The term Cw is a constant and the symbol ∆ = (element volume)1/3. The tensor
Sijd is calculated from the strain-rate and vorticity tensors, as shown here
1
Sdij ¼ Sik Skj þ Ωik Ωkj δij Smn Smn Ωmn Ωmn (42)
3
Now that the main CFD models have been presented, we turn attention to
comparisons of the results from different models. There are comparisons available
in [7, 8, 34, 35, 37, 39–46] and a very small subset of those comparisons will be
provided here. We have selected the classic problem of flow over a square blockage.
This canonical problem has the features that elucidate the strengths and weaknesses
of the particular models. For instance, some important parameters relate to the
time-averaged interactions between the fluid and the solid structure (drag force).
Also, there are significant unsteady phenomena, particularly in the wake region that
provide a challenging test for the models. In addition, this is a problem with
extensive experimental work that will serve as the basis for evaluating the results.
To begin we refer to Figure 4 which shows the solution domain (similar to [35]).
Figure 4.
Geometry for flow over a square cylinder.
11
Computational Fluid Dynamics
Figure 5.
Computational mesh used for square cylinder simulation.
Figure 6.
Drag coefficients for flow over a square cylinder and comparison with experiments. Reynolds numbers range
from 1 to 10,000,000.
12
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
Figure 7.
Drag coefficients for flow over a square cylinder and comparison with experiments, Reynolds numbers ranging
between 10 and 10,000.
13
Computational Fluid Dynamics
Figure 8.
(upper image) velocity contour and vectors for SST model and (lower image) side-by-side comparison of
leading-edge flow for SST and WALE LES models.
14
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
Figure 9.
Streamline patterns and velocity contours for Re = 100,000 flow over a square cylinder. Images at a sequence of
time instances, using SST model.
attained from the unsteady calculations. On the other hand, the period is very
different between the two.
4. Concluding remarks
15
Computational Fluid Dynamics
Figure 10.
SST solution that began as steady state and then was changed to unsteady.
Author details
© 2021 The Author(s). Licensee IntechOpen. This chapter is distributed under the terms
of the Creative Commons Attribution License (http://creativecommons.org/licenses/
by/3.0), which permits unrestricted use, distribution, and reproduction in any medium,
provided the original work is properly cited.
16
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
References
17
Computational Fluid Dynamics
18
Turbulence Models Commonly Used in CFD
DOI: http://dx.doi.org/10.5772/intechopen.99784
International Journal of Heat and Mass enhancement in tube flow with twisted
Transfer. 2018;117:1231-1250. tape insert, Progress in Computational
Fluid Dynamics, an International
[33] Everts M, Meyer J. Heat transfer of Journal, 2017; 17(3): 193-197.
developing and fully developed flow in
smooth horizontal tubes in the [42] S Bhattacharyya, H Chattopadhyay,
transitional flow regime. International S Bandyopadhyay, S Roy, A Pal, S
Journal of Heat and Mass Transfer. Bhattacharjee, Experimental
2018;117:1331-1351. investigation on heat transfer
enhancement by swirl generators in a
[34] Menter F. Best Practices: Scale- solar air heater duct, International
Resolving Simulations in ANSYS CFD, Journal of Heat and Technology, 2016;
Published by ANSYS, 2015. 34(2): 191-196.
19