You are on page 1of 10

Choosing the Right Turbulence Model for Your

CFD Simulation
Shawn Wasserman posted on November 22, 2016 | Comment 24126 views

Google Bookmark Facebook Twitter Print More

Since the 19th century, finding a simulation model to describe turbulence


perfectly has proven to be a bumpy ride.

Despite this, engineers need ways to simulate turbulent fluid flow to optimize
their designs for the real world. Various empiric or semi-derived turbulence
models have been created to help engineers to find the best model to fit their

system of study, but this process could take a lot of trial, error and physical
Turbulent CFD simulation of the air
velocity around landing gear. (Image testing.
courtesy of CD-adapco/Siemens.)

“To make the selection of a turbulence model easier for end users,” suggested
David Corson, director of program management at Altair, “[here are] what are widely accepted to be the most accurate
general-purpose models: Spalart-Allmaras, SST and k-omega. For the majority of engineering applications, these models
provide a good trade-off between [computational] cost and accuracy.”

Unfortunately, engineers need more than just a short list to make a correct selection. MIT professor Emilio Baglietto noted the
importance of understanding the fundamental challenges, myths, fallacies, successes and failures of computational fluid
dynamics (CFD) to determine a model with accuracy.

The Difficulty Defining Turbulent Fluid Flow


Baglietto explained that the mission to find a general solution to turbulence is known as the turbulence closure problem. The
aim is to close the Navier-Stokes and Reynolds stress equations that describe turbulent flow. The solution has remained
elusive, as averaging nonlinear occurrences of fluctuating quantities will only create new unknowns without governing
equations.
Simulation of the fluid flow in a washing machine using SIMULIA. (Video courtesy of Dassault Systèmes.)

Turbulence models attempt to close the system of equations that describe turbulent flows by devising new equations through
experimentation or derivations for specific applications.

Corson noted that in making a turbulent model, many assumptions are made to reduce the computational costs of
the simulation. Based on the type of flow being modeled, different assumptions will be made.

This has created a ballooning number of available turbulent models. This can make choosing a CFD simulation
software solution a considerable challenge for engineering teams because while more is not always an advantage, if your
software has too few turbulence models then you might miss the one you need.

“When someone shops for a CFD code, they might think that it would be an advantage to have many turbulence models,” said
Paul Malan, director of fluids applications for SIMULIA R&D. “Let’s say that they make the purchase of their perfect code
with, say, 50 different models. They are thrilled, because surely at least one of these will give the right answer. But when he
starts to solve a real problem, he has to choose one of the 50. Which should he choose? And once he has made the choice, how
does he know it is giving the right answer?”

The key to choosing the right model is to understand its strengths, weaknesses and definitions. According to Corson, “Until
there is a single model of turbulence developed, CFD engineers will always be faced with the challenge of selecting the right
model for the right job.”

The following is a list of turbulent model families and how they compare.

Reynolds-averaged Navier-Stokes Models


The family of Reynolds-averaged Navier-Stokes (RANS) models is the largest in the field of turbulence. These models attempt
to close the turbulence equations using viscosity terms. A common variable calculated in these models is k, or the kinetic
energy per unit mass of turbulent fluctuations.

Baglietto explained that there are numerous ways to perform these closures, but some are much more common and
instructive than others. Typically, algebraic models have been used with either one or two equations.

“That loss of degrees of freedom bakes in an inherent assumption that the turbulence is isotropic and not stretched by the
proximity of the wall, strong shear, or swirling flow,” said David Mann, product manager for STAR-CCM+ at CD-adapco. “We
should look for extra treatments in RANS models to overcome these limitations, or they will perform badly for these flows.”

There are some limitations with RANS models as they are based on the definition of turbulent viscosity. These limitations are:

Lack of physical description


Turbulence-induced secondary flows
Streamlined curvatures
Swirling flows or flows with rotations
Transitional flows between turbulent and laminar
Unsteady flows like internal combustion engines
Stagnant regions in flows

RANS Single-Equation Model: Spalart-Allmaras


“Spalart-Allmaras (SA) is a one-equation turbulence model that has been
developed specifically for aerodynamic flows such as transonic flow over
airfoils,” said Baglietto.

The model is based on kinematic eddy viscosity and mixing length. This mixing
length defines the transport of the turbulent viscosity.

Simulation of the turbulent flow Baglietto noted that this popularity is in large part due to the model’s
around a NACA profile calculated
using Spalart-Allmaras within robustness and fast implementation when modeling specialized flows. Spalart-
COMSOL. (Image courtesy of Allmaras is not memory-intensive and has good convergence but it has no wall
COMSOL.)
functions. The model is also a popular addition to various CFD codes.

“When we look at the benefits and drawbacks, the Spalart-Allmaras model has historically been a strength … due to its speed
and robustness,” said Corson.
“Because we are only solving a single equation for turbulence,” Corson added, “the non-linear convergence is outstanding and
the model is very forgiving of poor quality mesh, particularly in the near wall region. The drawback is that it does have some
limitations due to the single-equation formulation. The turbulence length and time scales are not as well defined as they are
in other models such as SST.”

Limitations of Spalart-Allmaras include:

Shear flows
Under predicting separation
Decaying turbulence

RANS Two-Equation Model: Standard k-epsilon, Realizable k-epsilon, RNG k-epsilon


“In [the standard k-epsilon model] we solve for two variables, the turbulent
kinetic energy, k, and the rate of dissipation of kinetic energy, epsilon [ε],” said
Valerio Marra, marketing director at COMSOL.

Marra explained that the model uses wall functions to analytically account for
the fluid velocity in the viscous sublayer near the wall.

The technique offers good convergence and isn’t memory-intensive. Marra also
Turbulent flow around a car-like
explained that the model is typically used for external flows with complex
model calculated in COMSOL using a
k-epsilon model. (Image courtesy of geometry. However, it is also a good general-purpose model.
COMSOL.)

Baglietto noted that the equation for epsilon is postulated, so it isn’t perfect. Nonetheless, the model is used for the largest
number of applications. This is partly because many of the model’s limitations are well-known.

Limitations of k-epsilon include:

No-slip walls
Adverse pressure gradients
Strong curvatures
Jet flows
Difficulty solving for epsilon

Despite this, the model is reliable due to its predictability and numerous variants that aim to improve the model for several
applications.

Perhaps the most famous variation of the model is the realizable k-epsilon model. This variation modifies the equation for
epsilon and introduces the effect of the mean flow distortion on turbulent dissipation.
“[Realizable k-epsilon] is the default recommendation in mainstream commercial packages, therefore represents the most
proven, well-quantified and widely-documented of all closures,” said Baglietto. “The model has improved performance for
planar surfaces, round jets, rotation, recirculation and streamline curvature. It also improves the boundary layer under strong
adverse pressure gradients or separation. But it cannot do magic as it’s still based on [turbulent] viscosity.”

Malan clarifies that k-epsilon has also become the “de facto” standard two-equation model because its two-layer formulation
has improved its applicability to well-resolved boundary layers. It also has improved results for complex separated industrial
flows.

Another popular modification is the renormalization group (RNG) k-epsilon model. The model was originally derived by
attempting to solve for epsilon using the Navier-Stokes equation. The result was very much like the original equation.
However, an update of the method added a term to the epsilon equation that accounts for the mean flow distortion of
turbulence dissipation.

The result is that RNG produces lower turbulence levels and can underestimate the value of k. This produces a less viscous
flow that creates more realistic flow features with complex geometry. Though the method is popular, Baglietto notes that it
gets on the nerves of many modeling veterans as it is more accurate for the wrong reasons.

“It is the production of k that is overestimated by the EVM (eddy viscosity models) and not the level of epsilon,” explained
Baglietto, “so the cure should be found in a better representation of anisotropy and essentially of the normal stresses.”

Though the standard, realizable and RNG variations of k-epsilon are all popular with CFD vendors, Baglietto is correct that
the RNG model does have its detractors. This has caused at least one vendor to take action. “Although [SIMULIA] used to
provide a version of the RNG k-epsilon model, we will not be supporting it for the R2017x release,” explained Malan. “We feel
that it offers little or no advantage over the realizable k-epsilon model which has superseded it and we cannot convincingly
articulate why one would select it.”

RANS Two-Equation Model: Standard k-omega and SST k-omega

Another popular two-equation model pairs k with the specific rate of dissipation of kinetic energy, or omega (ω). Baglietto
explained that the aim of the standard k-omega model is to model near-wall interactions more accurately than k-epsilon
models.

However, he noted that k-omega can over-predict shear stresses of adverse pressure gradient boundary layers and that the
model has issues with free stream flows. The model is also very sensitive to inlet boundary conditions, which is a disadvantage
not seen in k-epsilon.
Left: Simulation of a turbulent flow modeled with the shear stress transport
(SST) k-omega turbulence model in Altair AcuSolve. Right: Comparison of the
convergence rate for the model solved using Spalart-Allmaras, SST k-omega
and standard k-omega models. (Image courtesy of Altair.)

“The most significant advantage of the k-omega model is that it may be applied throughout the boundary layer without further
modification,” said Baglietto. “Furthermore, the standard k-omega model can be used in this mode without requiring the
computation of wall distance.”

“[k-omega] is a popular model for turbomachinery simulations and for simulations where strong vortices are present such as
those originating from wing tips,” said Mann. “[It] performs well for swirling flows and in the near wall region, but it over-
predicts separation.”

Limitations of k-omega include:

Difficulty of convergence compared to k-epsilon


Sensitivity to initial conditions

One variant of k-omega that has gained popularity, especially in the aeronautics area, is the shear stress transport (SST)
model. The model has gained this popularity based on its ability to predict separation and reattachment better when
compared to k-epsilon and the standard k-omega.

“The SST k-omega model is an enhancement of the original k-omega model and addresses some specific flaws of the base
model, such as the sensitivity to freestream turbulence levels,” explained Malan. “It has the advantage that it can be applied to
the viscous-affected region without further modification, which is one reason it has become a popular choice in aerospace
applications where the flow is deemed too complex for Spalart-Allmaras.”

The SST model accounts for cross-diffusion which better marries the k-epsilon and k-omega models. Using a blended function
based on wall distance, engineers can include cross-diffusion when away from the wall but not near it. In other words, using
the wall distance as a switch, SST works like k-epsilon in the far field and k-omega near the target geometry.
“Purists may object strongly that the blending function crossover location is arbitrary and could obscure some critical feature
of the turbulence,” noted Baglietto. Clearly the model isn’t perfect; it also requires limiters to improve the prediction of
stagnant regions of the flow. Additionally, it has issues predicting turbulence levels and complex internal flows and it doesn’t
take buoyancy into account.

Malan added, “Some people claim that the model has superior performance to the k-epsilon model in simulating boundary
layers with adverse pressure gradients. Ultimately, though, the performance of SST k-omega is not very different from the
realizable k-epsilon two-layer model. The choice between the two will typically be made based on user preference.”

It seems that many engineers do prefer k-omega as all the CFD vendors interviewed have the SST model and most have the
standard k-omega within their code.

Large-Eddy Simulation and Detached Eddy Simulation Models


RANS models simulate all scales of turbulence and resolve none. Large-eddy
simulation (LES) and detached-eddy simulation (DES) models, on the other
hand, resolve the largest scales of turbulence and model the rest by use of sub-
grid turbulence models or by blending with a RANS model.

The LES model is used to predict large turbulent eddy structures when solving
a CFD model system with a fine mesh. However, since turbulent scales are
Simulation of a turbulent flow around
a cylinder using Altair’s Acusolve LES small near the wall, the model is unable to predict these regions with accuracy.
turbulence model. (Image courtesy of
Altair.)
“LES and DES simulations are being carried out more and more often for
applications such as aeroacoustics or combustion and again there are several variants of these models,” explained Mann. “DES
is a hybrid RANS-LES method which combines the benefits of LES for resolving the large turbulent structures away from the
wall, with the benefits of RANS near the wall where the turbulent eddies are too small to resolve. It is important to remember
that the RANS portion of DES models is still responsible for the prediction of separation, heat transfer and other near-wall
effects.”

The biggest limitations with both the LES and DES models are their high computational and programming costs. This likely
explains why LES and DES models are not that popular with CFD software vendors. So if you need to use one for your
application, choose your CFD software wisely.

“All RANS models [are] limited in accuracy for highly separated flows,” explained Corson. “For these types of applications, or
those that require explicit resolution of turbulent structures, it is necessary to move towards a scale resolving simulation. DES
models fulfil this requirement for users, but come at the expense of increased compute time.”
“When it comes to scale resolving simulations, Spalart-Allmaras-based DES—more specifically, delayed detached eddy
simulation—is by far the most popular among our users,” said Corson. “This model is very stable and provides excellent
accuracy for highly separated flows. For attached flows in which the smaller scales of turbulence are important, users typically
choose the dynamic subgrid scale LES model. This model has excellent accuracy, and has little or no drawbacks in comparison
to the fixed coefficient version.”

Reynolds Stress Model


“Reynolds Stress Model (RSM) is the most complete physical representation of turbulent flows,” said Baglietto. “It is useful for
new challenges and is able to capture complex strains like swirling flows and secondary flows. For swirling flows, such as
cyclones, RSM is the only accurate closure.”

These models attempt to model the flow and terms directly in RANS equations. These models are based on the six equations
that represent turbulent stresses. They represent the flow very well but at the cost of high computational work. They are
typically reserved for flows that are extremely complex or have never been studied before.

Limitations of RSM include:

Computational expense
Sensitivity to initial conditions
Amount of modeling required
Requirement of high-quality mesh

Due to the difficulty in using these models, they are not that popular with CFD vendor software. Therefore, engineers looking
to use RSM will need to do their research, or read this eBook.

So, How Do I Choose My Turbulence Model Again?


Now that you’ve learned all this information about the turbulence model families, you might be asking yourself, “So, how do I
choose my turbulence model again?”

“Before choosing a model we need to ask ourselves what question is it that I am looking for an answer to,” said Mann. “Then
we need to understand the strengths and more importantly weakness of each model so that we can be sure that the strengths
of the model we choose is aligned with the type of problem and that we are not asking a model to do something it is weak at.”

Mann explained this with a great example; let’s say you want to look at the air flow around an airplane. Spalart-Allmaras
would be a great choice in this instance because it’s tested and well-known for this sort of application. However, if you want to
dig into your design further and determine the angle of attack that will cause
the airfoil to stall, then Spalart-Allmaras is no longer the model of choice.

“It will tell you the flow is still attached long after it has separated in reality,”
explained Mann. “The reason being is that although the model was designed
for attached aerospace flows it simply does not have enough degrees of
freedom to predict stall adequately.”

Other factors affect the choice of model such as the mesh resolution near the
wall. This is because turbulent flow near the wall is different from that in the
bulk. Normal to the wall, the flow is constrained and eddies become
Simulation of a 19.7-ft (6-m) ozone
reactor calculated in COMSOL anisotropic; near the wall, the flow becomes laminar at the viscous sub-layer.
Multiphysics. Proper assessment of the
turbulence allows for estimations of This doesn’t fit many turbulent models that assume the flow is completely
the residence times of each chemical
species. (Image courtesy of COMSOL.) turbulent and isotropic.

If the mesh is fine near the wall, the model will need to be compatible with near-wall turbulent flow. “Knowing how your
chosen turbulence model deals with the anisotropy in the near-wall flow and in other features such as swirling flow is key to
getting the best out of your model choice,” said Mann.

Marra agreed that certain models treat the viscous sublayer and buffer layer differently through the usage of wall functions.
These models will differ based on the number of variables solved, what these variables represent and the velocity and pressure
values.

“Each turbulence model has strength and weaknesses. Being aware of their range of applicability is of the essence in picking
the right one,” noted Marra. “Some models are well-suited for internal flows, others for external flow around complex
geometries. Some engineers might be interested in separated flows, jets, or need to compute lift, drag, heat fluxes [and more]
with high accuracy. Once a model that meets the criteria for the job at hand is picked, the next step is to use a mesh able to
capture all details of the flow.”

Corson explained that best practices at Altair include identifying the dominant feature of a turbulent flow and basing the
choice of model on this feature. The engineer can then study how the model performs with situations and canonical flows
where these features are dominant and compare the results to experimental data.

“Once you've identified which model performs best for the canonical flows of interest, you can apply that to your more
complex application,” mentioned Corson. “We can't guarantee that the models will provide the same level of accuracy on the
complex case, but this approach provides a good starting point that should result in reasonable levels of accuracy.”
However, Marra also suggests that other factors can affect why an engineer would choose a certain turbulence model. These
factors are:

1. Accuracy of the model when used in their original scope


2. Model’s ability to produce appropriate results in applications it isn’t intended for
3. Computational cost of the model and its ability to produce quick preliminary results to rule out early design options

Sometimes, an engineer will need to still use a computationally expensive model with limited computational power. In these
situations, Marra suggests a best practice of using boundary layer meshes at the wall and adaptive mesh refinements within
the bulk of the fluid. This will help engineers to balance the accuracy they need with the computational power they have.

But in the end, choosing the right model comes down to practice. A seasoned CFD simulation expert might be able to look at
an application and name off a few models of choice. They can then verify the correct model from the shortlist based on the
convergence of the solution and mesh.

However, no matter who you are or what you are simulating, it is always a good practice to verify that the turbulence model is
producing results in line with experimental data. Even the best of us can get it wrong and it’s best to find that error early in
development cycles.

“For applications that demand highly accurate resolution of specific flow features, the only way to determine the best
modeling approach is to rely on comparison to experimental data that is specific to that application,” noted Corson. “In that
case, a turbulence model sensitivity study is necessary to identify which model produces the best results in comparison to
experimental data. Once the best practice for that application is desired, similar applications in the future can rely on the
same modeling guidelines.”

Once you have chosen a few potential turbulence models for your application, you will then need to ensure that these models
are available in the CFD software you have access to. Otherwise, you might need to look to source a new CFD platform.

To see what models are available from various CFD vendor software, check out the eBook: Turbulence Models Offered by CFD
Simulation Vendors.

You might also like