You are on page 1of 16

Tips & Tricks: Turbulence Part 1 - Introduction to Turbulence Modelling

Posted By LEAP CFD Team on May 25, 2012 | 6 comments

We will now focus on Turbulence Modelling, which is a critical area for any engineer involved with
industrial CFD. There are a number of different approaches so it is important that you have solid
grounding in this area to enable you to choose the appropriate model for your simulation
requirements. It is worth noting that in August 2012, LEAP will be hosting Dr. Florian Menter to run a
series of Advanced Turbulence Training courses in Melbourne, Sydney and Perth. Dr. Menter is a world
recognised expert in turbulence modelling, and more information on his visit to Australia can be found
here.

The ANSYS CFD Solvers solve the Navier Stokes and conservation equations, but as direct solutions are
not possible to resolve for any flows of an industrial Reynolds number then we need to do some
modelling, as opposed to resolving the values directly.

The equations that we used are not closed and so we need to use Turbulence Modelling to close the
equation set and then iterate towards a solution. We can use what is called a Reynolds Averaged Navier
Stokes (RANS) approach, or we can use an Eddy Simulation technique which resolves the larger eddies in
the flow and is only really required when you have separation or large recirculating regions.

The most commonly used models are the RANS models due to their low cost in terms of compute power
and run times. The Eddy Simulation methods can be quite mesh sensitive but will yield much better
results for separated and recirculating flow, albeit over much longer run times.

The RANS models apply a Reynolds decomposition technique to the Navier Stokes equations which
breaks the velocity down into its mean and fluctuating components. This decomposition leaves us with
one unknown value, which is termed the Reynolds Stress. We use Turbulence Models to resolve the
Reynold’s Stress and close the equation set. There are two ways we can go about resolving this, the first
(and most commonly used approach) is to use an isotropic value for the turbulent viscosity value which
is called the an Eddy Viscosity Model, the other way is to solve using the Reynolds Stress Model (RSM)
for the 6 separate Reynolds Stresses, which results in an anisotropic solution.
EDDY VISCOSITY MODELS

The limitation with Eddy Viscosity models is that they use an isotropic value which may not be
appropriate and hence can increase the diffusion in your result. Obviously solving for the 6 Reynolds
Stresses and dissipation will be more accurate, but you are then solving extra equations which will
increase your run time considerably. There are further modifications to the 2 equation Eddy Viscosity
models that yield similar results to RSM which I will elaborate on soon.

I have listed below the most commonly used Eddy Viscosity turbulence models over the past ten or so
years and the intended use behind their development.

The common options are:

 Spalart-Allmaras – One equation model for attached aerodynamic analysis.

 k-epsilon – Two equation model for free shear and non-wall bounded flow behaviour. Was the
previous industrial standard.

 k-omega – Two equation model for wall bounded flows, not commonly used.

 SST (Shear Stress Transport) - Two equation model blending the freestream advantages of the k-
epsilon model with the wall bounded advantages of the k-omega model. This is the new
industrial standard and should be the default choice for most applications.

Note: Modifications for Curvature Correction to the SST model give comparable results to RSM and the
SST model also works very well with the Transition Model now available in CFX and FLUENT.

Options like Curvature Correction are very useful in tackling problems where previously you would have
required an RSM model. The Transition Models can be of benefit also, depending on your application
and give much better drag prediction as they will maintain laminar flow along a body and develop
natural transition points, as well as calculate regions of bypass transition. I won’t discuss these here, but
there is extensive information in the ANSYS User Guides. These additional options are available in
ANSYS CFX and ANSYS FLUENT.

REYNOLDS STRESS MODELS

The Reynolds Stress models were used widely for high swirling flows, for example, the flow in a cyclone
separator. In the past we would use the RSM models for this type of flow, but nowadays you should use
the SST model with the Curvature Correction model enabled.
EDDY SIMULATION

The more advanced options for resolving turbulence are the simulation based models. These generally
require much better mesh resolution and involve much longer run times as they can only be run as an
unsteady simulation. These are used primarily for resolving large scale separations and recirculating
regions.

The options in order of least expensive, to most expensive are:

- Scale Adaptive Simulation

- Detached Eddy Simulation

- Large Eddy Simulation

The Scale Adaptive Simulation is the best point to start when stepping up from a RANS modelling
analysis. The Scale Adaptive Simulation is based on an unsteady SST RANS model but calculates the local
length scales and resolves accordingly. The advantage of this is that it can be run on a good quality
RANS mesh without any additional meshing requirements. For some flow behaviours this has been
shown to give very similar results to Large Eddy Simulation (LES) and Detached Eddy Simulation
(DES). This is why we would recommend this as a starting point for unsteady simulations with separated
or recirculating flows, and as a good alternative to an unsteady RANS SST simulation.

SAS solution in a Swirl Combustor

The Detached Eddy Simulation resolves the boundary layer and the smaller eddies by using a RANS
solution in those regions. This can produce quite good results for a fraction of the cost of LES, but is very
sensitive to the mesh resolution. You need to ensure a very good quality mesh in the boundary layer
and a good transition region from the boundary layer to the freestream to help the solver switch from
the RANS solution in the boundary layer to an LES solution for the larger eddies. This adds complexity
and has been known to cause erroneous results.

DES solution in a Swirl Combustor

Large Eddy Simulation is the most expensive of the three simulation options and requires a very fine grid
to accurately resolve the eddies, especially in the boundary layer. Once the eddies are too small to be
resolved by the grid, they fall back into a sub-grid model. As this sub-grid model is used only to resolve
the smallest eddies which tend to have more universal properties, this yields very accurate results, but
because of the large mesh requirements this is not a commonly used approach in industrial settings.

We can now use LES in specific regions of the mesh only using Embedded LES. This means that we can
run an unsteady RANS in the bulk of the flow and a full LES calculation in only the required separated
regions hence making it more practical for industrial use.

Each of these simulation methods when used correctly will yield a better result than the steady state or
unsteady RANS models for separated flow or recirculating flow behaviour but it is more expensive to do
so.

As far as Turbulence Models are concerned, the SST is a good default choice for most flow
problems. Please contact our Technical Support Team if you would like to discuss your flow problem
and the options available to you. If you are interested in attending the Advanced Turbulence
Training courses by Dr. Florian Menter, more information can be found here.

In the next blog post we will look at the importance of the y+ value and discuss the use of appropriate
values for the different turbulence models.
Previously we have discussed the importance of an inflation layer mesh and how to implement
one easily in ANSYS Meshing. We also touched upon the concept of mesh y+ values and how
we can estimate them during the inflation meshing process. In other posts, we also discuss the
different turbulence models and eddy simulation methods available to ANSYS CFD users. In
today's post, we'll talk in more detail about y+ values apply to the most commonly used
turbulence models.

From our earlier discussions, we now understand that the placement of the first node in our near-
wall inflation mesh is very important. The y+ value is a non-dimensional distance (based on
local cell fluid velocity) from the wall to the first mesh node, as you can see in the image
below. To use a wall function approach for a particular turbulence model with confidence, we
need to ensure that our y+ values are within a certain range.

y+ definition

Looking at the image above, we need to be careful to ensure that our y+ values are not so large
that the first node falls outside the boundary layer region. If this happens, then the Wall
Functions used by our turbulence model may incorrectly calculate the flow properties at this first
calculation point which will introduce errors into our pressure drop and velocity results. The
upper range of applicability will vary depending on the flow physics and the extent of the
boundary layer profile.

For instance, flows with very high Reynolds numbers (typically aircraft, ships, etc) will
experience a logarithmic boundary layer that extends to several thousand y+ units, whereas low
Reynolds number flows such as turbine blades may have an upper limit as little as 100 y+
units. In practice, this means that the use of wall functions for these class of flows should be
avoided as their use will limit the overall number of mesh nodes that can be sensibly placed
within the boundary layer. In general, it is recommended that you endeavour to place sufficient
inflation layer cells within the boundary layer, rather than simply focusing on achieving any
particular y+ value. This will be covered in detail in a future post

In addition to the concern about having a mesh with y+ values that are too large, you need to be
aware that if the y+ value is too low then the first calculation point will be placed in the viscous
sublayer (logarithmic) flow region and the Wall Functions will also be outside their validity
(below about y+ < 11). You can imagine that this would become an issue if a mesh intended to
be used with wall functions is then refined near the wall. Fortunately, the use of scalable wall
functions in ANSYS CFD products now takes care of these problems and produces consistent
results for grids of varying y+. Without any further user involvement, the scalable wall functions
activate the local usage of the log law in regions where the y+ is sufficiently small,
in conjunction with the standard wall function approach in coarser y+ regions.

So, where should you start? We have learnt that the wall function approach and y+ value
required is determined by the flow behaviour and the turbulence model being used. If you have
an attached flow, then generally you can use a Wall Function approach, which means a larger
initial y+ value, smaller overall mesh count and faster run times. If you expect flow separation
and the accurate prediction of the separation point will have an impact your result, such as the
drag or lift forces experienced by the ellipse below, then you would be advised to resolve the
boundary layer all the way to the wall with a finer mesh. Please refer to this post for a more
detailed explanation of appropriate turbulent wall function and modelling approaches.

Wall Function applicability

Once we know our preferred approach, we can estimate the thickness for our first inflation layer
cell using the equation below, which can be used to calculate the distance value for a specific
velocity fluid and the required y+ value (based on the flow over a flat plate). This is usually a
good initial estimate and the y+ value we aim for will depend on our turbulence model selection.

Note that Δy is the distance of the first node from the wall, L is the flow characteristic length
scale, y+ is the desired y+ value, Re_L is the Reynolds Number based on your problem's
characteristic length scale.

Unfortunately, as the y+ value is dependent on the local fluid velocity which varies across the
wall significantly for most industrial flow applications, it is not possible to know your exact y+
prior to running an initial simulation. For this reason, it is important that you get into the habit of
checking your y+ values as part of your normal post-processing in ANSYS CFD-Post so that you
can make sure you are in the valid range for your flow physics and turbulence model selection.

Our next post in this series concentrates on the feasibility and selection of different wall
functions, based on the applied turbulence modelling strategy.

This is still an area of active research and is a hot topic for many of our CFD users. If you have
any questions or comments, please leave a message below or contact our CFD Technical Support
team for more detailed technical information on these topics.
In recent posts in our series of Turbulence Modelling posts, we have covered boundary layer theory and
touched on some useful meshing and post-processing guidelines to check you are appropriately
resolving the boundary layer profile. Today we will consider three critical questions that are often asked
by CFD engineers when developing or refining a CFD simulation:

- Am I using the correct turbulence model for the type of results I am looking for?

- Do I have an appropriate Y+ value and a sufficient number of inflation layers?

- Am I using the right wall function for my problem?

This topic is so important because we know that in turbulent flows the velocity fluctuations within the
turbulent boundary layer can be a significant percentage of the mean flow velocity, so it is critical that
we capture these effects with accuracy. A Reynolds averaging approach using turbulence models will
provides us with an estimate of the increased levels of stress within the boundary layer, termed the
Reynolds stresses. In order to appreciate the use of wall functions and the influence of walls on the
turbulent flowfield, we should first gain familiarity with the composite regions of the turbulent
boundary layer:

Composite regions of the turbulent boundary layer


In the laminar sub-layer region (Y+ < 5) inertial forces are less domineering and the flow exhibits laminar
characteristics, which is why this is known as the low-Re region. Low-Re turbulent models (e.g. the SST
model) aim to resolve this area and therefore require an appropriate mesh resolution to do this with
accuracy. This is most critical for flows with a changing pressure gradient where we expect to see
separation, as observed below.

Predicting separation in a diffuser-type geometry

In the law of the wall region, inertial forces strongly dominate over viscous forces and we have a high
presence of turbulent stresses (this is known as the high-Re composite region). If using a low-Re model,
the whole turbulent boundary layer will be resolved including the log-law region. However, it possible to
use semi-empirical expressions known as wall functions to bridge the viscosity-affected region between
the wall and the fully-turbulent region.

Contours of the eddy viscosity ratio on a low-Re grid illustrating high turbulent viscosity in the log-law
region as opposed to the laminar sub-layer
The main benefit of this wall function approach lies in the significant reduction in mesh resolution and
thus reduction in simulation time. However, the shortcoming lies in numerical results deteriorating
under subsequent refinement of the grid in wall normal direction (thus reducing the Y+ value into the
buffer layer zone). Continued reduction of Y+ to below 15 can gradually result in unbounded errors in
wall shear stress and wall heat transfer (due to the damping functions inherent within the wall function
approach).

Bearing all of the above in mind, and keeping our eye on finding the right balance between accuracy,
stability and speed, we can tackle a wide variety of CFD problems using the following guidelines:

What results am I interested in and am I using the right turbulence model?

If our aim is to accurately predict the boundary layer velocity or thermal profile, or if the developing
boundary layer will tend to separate (due to a changing pressure gradient – and not because of sharp
edges or discontinuities in the geometry), then we recommend the use of a low-Re model. Low-Re
models are also required for accurate pressure-drop or drag calculations. We highly recommend the use
of the Shear Stress Transport (SST) model, but all ω-based models or ε-based models with enhanced wall
treatment may be used. For high speed external aerodynamic flows, the one-equation Spalart-Allmaras
model (with Y+ < 2) may also be considered to reduce the computational time. Alternatively, for flows
where wall-bounded effects are not a priority, or if separation is expected to occur only due to sharp
changes in the geometry, an ε-based wall function approach is more than sufficient. In ANSYS CFD, all ω-
based models and the SST model are capable of resorting to a wall function formulation (automatic wall
treatment) in the presence of coarse mesh resolutions near the wall without any further user input. Wall
function models are also useful for calibrating our CFD models, due to the decreased simulation time.

Flow pattern with separation and reattachment on a rotor blade

What is my Y+ value and do I have a sufficient number of prism layers?

When using low-Re models or any models with enhanced wall treatment, the average Y+ value should
be on the order of ~1 to ensure we are capturing the laminar sub-layer. When using wall function
models, the Y+ value should ideally be above 15 to avoid erroneous modelling in the buffer layer and the
laminar sub-layer. High quality numerical results for the boundary layer will only be obtained if the
overall resolution of the boundary layer is sufficient. This requirement in some cases is more important
than achieving certain Y+ values. The minimum number of cells to cover a boundary layer accurately is
around 10, but values of 20 are desirable. The total thickness of the prisms should be implemented such
that around 15 or more nodes are actually covering the boundary layer. Our next post in this series on
the turbulent boundary layer will cover a very useful and practical technique to post-process the
resolution of the boundary layer, and offer insight into modifications required to improve accuracy.

Boundary layer velocity profile modeled with standard k-e for three different mesh densities using
Enhanced Wall Treatment

Am I using the right wall function?

In ANSYS CFD, all turbulence models are y-plus independent. However selecting the most appropriate
wall function is dependent on level of refinement of our wall adjacent mesh, or the relative scales in our
flow. Use of the standard wall function (ε-based models) implies that our boundary layer mesh lies
entirely within the log-law region of the boundary layer. For industrial applications, this in fact might be
difficult to achieve due to varying geometrical and velocity scales associated with our model – and
therefore grids inherently designed with arbitrary refinement. We highly recommend the use of the
scalable wall function, which offers an elegant solution to this ambiguity often encountered. This wall
function virtually displaces the mesh to a Y+ ~ 11.225 (transition to the log-law composite layer)
irrespective of the level of refinement, thereby avoiding the erroneous modelling of the laminar sub-
layer and buffer region. It is also important to note that for grids designed with a Y+ > 11.225, the
scalable wall function will provide identical results to the standard wall function. Enhanced wall
treatment may further be selected for ε-based models on refined low-Re grids, and is also formulated
such that it can perform well for meshes of intermediate resolution. However the use of enhanced (or
non-equilibrium) wall treatment for low-Re modelling of the turbulent boundary layer is generally not
recommended and more confidence in our solution can be obtained by selecting a suitable ω-based
formulation, such as the SST model.
In recent posts we have comprehensively discussed inflation meshing requirements for resolving
or modeling wall-bounded flow effects due to the turbulent boundary layer. We have identified
the y-plus value as the critical parameter for inflation meshing requirements, since it allows us to
determine whether our first cell resides within the laminar sub-layer, or the logarithmic region.
We can then select the most suitable turbulence model based on this value. Whilst this theoretical
knowledge is important regarding composite regions of the turbulent boundary layer and how it
relates to y-plus values, it is also useful to conduct a final check during post-processing to ensure
we have an adequate number of prism layers to fully capture the turbulent boundary layer profile,
based on the turbulence model used (or more precisely, whether we aim to resolve the boundary
layer profile, or utilize a wall function approach). In certain cases, slightly larger y-plus values
can be tolerated if the boundary layer resolution is sufficient.

How can I check in CFD-Post that I have adequately resolved the boundary layer?

For the majority of industrial cases, it is recommended to use the two-equation turbulence
models, or models which utilize the turbulent viscosity concept and the turbulent viscosity ratio
(i.e. the turbulent viscosity over the molecular viscosity). We can make use of this concept to
visualize the composite regions of the turbulent boundary layer, and ultimately visualize how
well we are resolving the boundary layer profile. Consider the conceptual case-study of the
turbulent flow over an arbitrarily curved wall. Prism layers are used for inflation, and tetra
elements in the free-stream. Once we have calculated the solution, within CFD-Post we can
create an additional variable for the eddy viscosity ratio. Then by plotting this variable on a
suitable plane, and superimposing our mesh in the near-wall region, we can visualize the
boundary layer resolution.

Figure 1: RKE with standard wall function – y-plus of 75 with 10 prism layers

Figure 1 provides an example of a reasonable wall function mesh. There is a good cell transition
from the prisms to the free stream tetra elements. The y-plus we have prescribed at the first cell
indicates we are in the logarithmic composite region of the turbulent boundary region, which is
the region largely dominated by inertial forces and thus we have high levels of turbulence. The
turbulence gradually dissipates as we approach free stream conditions (where the levels of
turbulence are governed by inlet conditions), which is expected. At this stage, we could even
reduce the number of cells in the inflation layer as we are clearly capturing the logarithmic
region layer before approaching the free stream. Correspondingly, we could aim to reduce the y-
plus value (y-plus ~ 20) to better capture the increase in turbulent viscosity as we move from the
inner layer to the outer layer of the logarithmic region.

Figure 2: SST low-Re model – y-plus of 1 with 20 prism layers

Figure 2 provides a good mesh for a low-Re turbulence model. We observe that the transition in
size from the final prism layer to the free stream tetra elements has been regulated well. Since we
have prescribed a y-plus value of 1 we are within the laminar sub-layer, which exhibits laminar
flow characteristics (thereby resulting in no turbulent viscosity). As we gradually move through
the buffer region and into the logarithmic region we see a large rise in the viscosity ratio before it
dissipates into the free stream. This maximum value will generally occur near the middle of the
boundary layer, which also gives us an indication of the physical boundary layer thickness (twice
the location of the maximum eddy viscosity ratio gives the boundary layer edge). As per the
example given in Figure 2, it is essential that the prism layer is thicker than the boundary layer as
otherwise there is a danger that the prism layer confines the growth of the boundary layer.

Figure 3: SST low-Re model – y-plus of 0.25 with 20 prism layers


Figure 3 ia an example of a poor quality mesh for a low-Re turbulence model (such as SST k-
omega). In this case we have unnecessarily prescribed a very low y-plus value yet we have not
compensated by appropriately allowing for more prism layers in the inflation layer. Therefore,
we are capturing the laminar sub-layer to an excessive detail, and the boundary layer does not
transition to the logarithmic region until we are well inside the free stream. Consequently there
are cells which are not aligned with the direction of the flow and thus our boundary layer profile
will not be well resolved (its growth may be confined by the extent of the prism layers), hence
affecting our drag or pressure-drop calculations.

Figure 4: RKE with standard wall function – y-plus of 1 with 20 prism layers

Figure 4 is an example of a poorly defined mesh for a standard wall function turbulence model.
The accuracy of wall function or high-Re turbulence models (e.g. the k-epsilon variants) cannot
be confirmed modeling the laminar sub-layer and thus should be avoided. In ANSYS Fluent, the
laminar stress-strain relationship is employed when the mesh is below a y-plus of 11.225 (noted
as the transition to the logarithmic region). After which, the logarithmic wall functions are
employed. This is an example where a low-Re turbulence model should be used (c.f. Figure 2),
or alternatively we could aim to increase our y-plus value such that it resides in the logarithmic
region (c.f. Figure 1).
Figure 5: RKE with scalable wall function – y-plus of 1 with 20 prism layers

The problems in Figure 4 can be overcome using scalable wall functions, as shown in Figure 5
using the same mesh. The purpose of scalable wall functions is to force the usage of the
logarithmic law. Here we can see a turbulent viscosity distribution which is analogous to the case
presented in Figure 1 (with the exception that we are now capturing the increase in turbulent
viscosity as we move from the inner layer to the outer layer of the logarithmic region). For this
simple case, we could potentially save simulation time by coarsening the mesh immediately
adjacent to the wall, or alternatively we could opt for a low-Re turbulence model. The real
advantages of the scalable wall functions arise for complex flows on grids of arbitrary refinement
(or correspondingly flows with various boundary layer scales) since it will provide consistent
modelling.
In previous posts we have stressed the importance of using an appropriate value in
combination with a given turbulence modelling approach. Today we will help you calculate the
correct first cell height ( ) based on your desired value. This is an important first step as
the global mesh resolution parameters will also be influenced by this near-wall mesh as well as
the Reynolds number.

Let's review the two main choices we have in choosing a near-wall modelling strategy:

Resolving the Viscous Sublayer

 Involves the full resolution of the boundary layer and is required where wall-bounded
effects are of high priority (adverse pressure gradients, aerodynamic drag, pressure drop,
heat transfer, etc.)
 Wall adjacent grid height must be order
 Must use an appropriate low-Re number turbulence model (i.e. Shear Stress Transport)

Adopting a Wall Function Grid

 Involves modelling the boundary layer using a log-law wall function. This approach is
suitable for cases where wall-bounded effects are secondary, or the flow undergoes
geometry-induced separation (such as many bluff bodies and in modern automotive
vehicle design).
 Wall adjacent grid height should ideally reside in the log-law region where
 All turbulence models are applicable (e.g. Shear Stress Transport or k-epsilon with
scalable wall functions)

During the pre-processing stage, we need to estimate the first cell height ( ) so that
our falls within the desired range. The computed flow-field will dictate the actual value
which in reality will vary along the wall. In some cases, we may need to locally refine our mesh
to achieve the desired value in all regions.

So how to calculate the First Cell Height for a desired Y+ value?

Firstly, we should calculate the Reynolds number for our model based on the characteristic
scales of our geometry such that:

where and are the fluid density and viscosity respectively, is the freestream velocity,
and is the characteristic length (e.g. pipe diameter, body length, etc.).
The definition of the value is such that:

The target value and fluid properties are known a priori, so we need to calculate the
frictional velocity , which is defined as:

The wall shear stress, can be calculated from skin friction coefficient, , such that:

The ambiguity in calculating surrounds the value for . Empirical results have been used
to provide an estimate to this value:

Flow Type Empirical Estimate


Internal Flows
External Flows

We can then input these known values into the above equations to estimate our value for .

When considering simple flows and simple geometry, we might find this correlation is highly
accurate. However, when considering complex geometry, refinement in the boundary layer may
be required to ensure the desired value is achieved. In these cases, you can choose to re-
mesh in ANSYS Meshing or use anisotropic mesh adaption (ie. adaption of local cells only in the
wall-normal direction) to achieve your desired Y+ value across the entire model. Please leave a
comment below or contact our support line if you have any questions.

For the lazy CFD-ers out there, we have written an applet for you to estimate the cell height
quickly.

You might also like