You are on page 1of 11

44th AIAA Aerospace Sciences Meeting and Exhibit AIAA 2006-395

9 - 12 January 2006, Reno, Nevada


AIAA 2006-0395
th
44 AIAA Aerospace Sciences Meeting
and Exhibit, Reno 2006

Predicting 2D Airfoil and 3D Wind Turbine Rotor


Performance using a Transition Model for General CFD
Codes

R.B. Langtry*, J. Gola†, and F.R. Menter‡


ANSYS CFX Germany, 83624 Otterfing, Germany

A new correlation-based transition model has been developed, which is built strictly on
local variables. As a result, the transition model is compatible with modern CFD techniques
such as unstructured grids and massively parallel execution. The model is based on two
transport equations, one for intermittency and one for a transition onset criterion in terms
of momentum thickness Reynolds number. The proposed transport equations do not
attempt to model the physics of the transition process (unlike e.g. turbulence models), but
form a framework for the implementation of transition correlations into general-purpose
CFD methods. The transition model was initially developed for turbomachinery and
aeronautical flows. The main goal of the present paper is to validate the model for predicting
transition on wind turbines. In this paper, fully turbulent and transitional computations of
the 2D S809 airfoil along with a full 3D wind turbine rotor (that uses the S809 airfoil) have
been accomplished. The transitional results are in good agreement with the experimental
data and the transition model would appear to be well suited for the prediction of wind
turbine aerodynamics.

Nomenclature

AoA = Angle of attack (deg.)


Cl = lift coefficient, lift/(0.5ρUref2Area)
Cd = drag coefficient, drag/(0.5ρUref2Area)
Cp = pressure coefficient, pressure/(0.5ρUref2)
FSTI = freestream turbulence intensity (percent), 100(2k/3)1/2/Uref
k = turbulent kinetic energy
Rex = Reynolds number, ρLUref/µ
Reθ = momentum thickness Reynolds number, ρθU0/µ
Reθt = transition onset momentum thickness Reynolds number (based on freestream conditions), ρθtU0/µ
~ = local transition onset momentum thickness Reynolds number (obtained from a transport equation)
Reθt
RT = viscosity ratio
Ry = wall-distance based turbulent Reynolds number
Rv = vorticity Reynolds number
S = absolute value of strain rate, (2SijSij)1/2
Sij = strain rate tensor, 0.5(∂ui/∂xj + ∂uj/∂xi)
Tu = turbulence intensity, 100(2k/3)1/2/U

*
CFD Development, Staudenfeldweg 12, 83624 Otterfing, Germany.

Intern, Staudenfeldweg 12, 83624 Otterfing, Germany.

Head of CFD Development, Staudenfeldweg 12, 83624 Otterfing, Germany.

1
American Institute of Aeronautics and Astronautics

Copyright © 2006 by ANSYS Inc. . Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
U = local velocity
Uo = local freestream velocity
Uref = inlet reference velocity
u’ = local fluctuating streamwise velocity
x/C = axial distance over axial chord
y = distance to nearest wall
y+ = distance in wall coordinates, ρyµτ/µ
δ = boundary layer thickness
θ = momentum thickness
λθ = pressure gradient parameter, (ρθ2/µ)(dU/ds)
µ = molecular viscosity
µt = eddy viscosity
ρ = density
Ω = absolute value of vorticity, (2ΩijΩij)1/2
Ωij = vorticity tensor, 0.5(∂ui/∂xj - ∂uj/∂xi)
ω = specific turbulence dissipation rate

Subscripts
t = transition onset
s = streamline

I. Introduction

W ind energy is rapidly becoming an economically viable energy source. This is partially due to increases in the
cost of oil, but also due to significant improvements in the performance of modern wind turbines. Like many
aerodynamic devices such as turbine blades and wings, a better understanding of the flow field can result in design
changes that can significantly improve performance. As well, it is critical that the aerodynamic characteristics of the
wind turbine be known during the design phase in order to have accurate economic projections. Because of the
costs associated with performing wind tunnel experiments, there is a significant amount of interest in predicting the
aerodynamic characteristics of a wind turbine using computational fluid dynamics (CFD). However, a survey of the
available literature indicates that there are still some significant issues associated with accurately predicting wind
turbine performance using CFD. According to Wolfe and Ochs (1997) there are two main issues. The first is that
the standard k-ε turbulence model (which is the default in many commercial CFD codes) under predicts the amount
of separation on an airfoil at high angles of attack. This can lead to overly optimistic performance projections. The
second problem is that very few CFD codes attempt to model laminar-turbulent transition, which can also have a
significant effect on the predicted performance. There are a few CFD codes that have been modified to predict
transition using either an empirical correlation or the eN method. However, these approaches are based on non-local
operations such as integrating the boundary layer quantities or tracking the amplification of instabilities along a
streamline. As a result, they are only applicable to structured grids, can only predict transition in 2D flows and often
cannot be parallelized in order to run large cases (e.g. a full 3D wind turbine with 10 million grid nodes).

Recently, a new correlation-based transition model has been developed2, which is based strictly on local variables.
As a result, the transition model is compatible with modern CFD approaches such as unstructured grids, massive
parallel execution and can also predict transition in 3D flows. The model is based on two transport equations, one
for intermittency and one for the transition onset criteria in terms of momentum thickness Reynolds number. The
proposed transport equations do not attempt to model the physics of the transition process (unlike e.g. turbulence
models), but form a framework for the implementation of correlation-based models into general-purpose CFD
methods. One of the central variables in the formulation of model is the strain-rate (or sometimes vorticity3)
Reynolds number:

2
American Institute of Aeronautics and Astronautics
ρy 2 ∂u ρy 2
Rev = = S (1)
µ ∂y µ

where y is the wall distance, ρ is the density, µ is the molecular viscosity and S is the absolute value of the strain
rate. The importance of ReV lies in the relation of its maximum value inside the boundary layer to the momentum
thickness Reynolds number ReΘ, of Blasius (or more generally Falkner-Skan) profiles4:

Re V ( x , ~y ) max ~ 2 . 193 Re Θ ( x ) (2)

where ~ y is the location where ReV has its maximum. The function Reν can be used on physical reasoning, by
arguing that the combination of y2S is responsible for the growth of disturbances inside the boundary layer, whereas
ν = µ / ρ is responsible for their damping. As y2S grows with the thickness of the boundary layer and ν stays
constant, transition will take place once a critical value of Reν is reached. The connection between the growth of
disturbances and the function Reν was shown by Van Driest and Blumer (1963) in comparison with experimental
data. The transition model was first detailed in Menter et al (2004) and has been extensively validated for predicting
transition in both turbomachinery6 and aeronautical flows2 under the influence of freestream turbulence, pressure
gradients and separation. It has been combined with the SST turbulence model7, which has been shown to be very
accurate for predicting turbulent separation on airfoils.

The first part of the paper will compute the 2D aerodynamic coefficients of the well-known S809 wind turbine
airfoil1 using the SST turbulence model along with the new transition model. The numerical results for this airfoil
will be compared to the experimental data obtained by Somers (1989,1997). The second part of the paper will detail
the numerical results obtained for the NREL Phase VI test case9, which was a 3D wind turbine that used the S809
airfoil for the blade profile. It will be shown that the transition model results in a significant improvement of the
predicted power output, particularly at the higher angles of attack. The differences between the 2D and 3D airfoil
characteristics will also be investigated, particularly at the higher angles of attack were there is a significant amount
of separated flow and the 3D effects are substantial.

II. Transition Model Formulation

The present transition model formulation is described very briefly for completeness, a detailed description of the
model can be found in Menter et al. (2004) along with some improvements for Natural and separation induced
transition which are detailed in Langtry and Menter (2005). The present transition model is summarized in the
following pages.

The transport equation for the intermittency, γ, reads:

∂ (ργ ) ∂ (ρU j γ ) ∂  µ  ∂γ 
+ = Pγ 1 − Eγ 1 + Pγ 2 − Eγ 2 +  µ + t   (3)
∂t ∂x j ∂x j  σ γ  ∂x j 

The transition sources are defined as follows:

Pγ 1 = Flength c a1 ρ S [γFonset ] α ;
c
Eγ 1 = ce1 Pγ 1γ (4)

where S is the strain rate magnitude. Flength is an empirical correlation that controls the length of the transition
region. The destruction/relaminarization sources are defined as follows:

3
American Institute of Aeronautics and Astronautics
Pγ 2 = c a 2 ρ Ω γ Fturb ; E γ 2 = c e 2 Pγ 2 γ (5)

where Ω is the vorticity magnitude. The transition onset is controlled by the following functions:

ρy 2 S ρk
ReV = ; RT = (6)
µ µω

Fonset 1 =
Re v
2 . 193 ⋅ Re θ c
; ( (
Fonset 2 = min max Fonset 1 , Fonset 1 , 2 . 0
4
) ) (7)

4
  R 3  R 
− T 
Fonset 3 = max 1 −  T  ,0 ; Fonset = max (Fonset 2 − Fonset 3 ,0 ); Fturb = e  4  (8)
  2.5  
 

Reθc is the critical Reynolds number where the intermittency first starts to increase in the boundary layer. This
~
occurs upstream of the transition Reynolds number, R eθt , and the difference between the two must be obtained
~
from an empirical correlation. Both the Flength and Reθc correlations are functions of R eθt .

The constants for the intermittency equation are:

ce1 = 1.0; ca1 = 2.0; cα = 0.5; ce 2 = 50; ca 2 = 0.06; σ γ = 1.0; (9)

The modification for separation-induced transition is:


4
  Re v    R 
− T 
γ sep = min  2 ⋅ max   − 1,0  Freattach , 2  Fθt ; Freattach = e  20  ; γ eff = max (γ , γ sep ) (10)
 
  3 .235 Re θc   

The boundary condition for γ at a wall is zero normal flux while for an inlet γ is equal to 1.0.
~
The transport equation for the transition momentum thickness Reynolds number, R eθt , reads:

~
~
(
∂ ρ R e θt ) (
∂ ρ U j R e θt ∂  ) ~
∂ R e θt 
+ = Pθ t +  σ θ t (µ + µ t )  (11)
∂t ∂x j ∂ x j  ∂ x j 

The source term is defined as follows:

ρ ~ 500 µ
Pθ t = c θ t
t
(Re θt )
− R e θ t (1 . 0 − F θ t ); t =
ρU 2
(12)

   y
4

 γ − 1 / ce 2   
2
  − 
Fθt = min max Fwake ⋅ e δ 
,1.0 −   ,1.0 (13)
 
 1.0 − 1 / ce 2   
   
~
R eθ t µ 15 50 Ω y
θ BL = ; δ BL = θ BL ; δ = ⋅ δ BL (14)
ρU 2 U

4
American Institute of Aeronautics and Astronautics
2
 Re 
ρωy 2 − ω

Reω = ; Fwake = e  1E +5  (15)
µ

~
The model constants for the R eθt equation are:

cθt = 0 .03; σ θt = 2 .0 (16)

~ ~
The boundary condition for R eθt at a wall is zero flux. The boundary condition for R eθt at an inlet should be
calculated from the empirical correlation based on the inlet turbulence intensity.

The model contains three empirical correlations. ReΘt is the transition onset as observed in experiments and it is used
in Eq.12. Flength is the length of the transition zone and goes into Eq. 4. ReΘc is the point where the model is activated
in order to match both, ReΘt and Flength, and it goes into Eq. 7. At present these empirical correlations are
proprietary and are not given in the paper.

~ ~
( )
ReΘt = f (Tu, λ ); Flength = f ReΘt ; ReΘc = f ReΘt ( ) (17)

The first empirical correlation is a function of the local turbulence intensity, Tu, and the Thwaites’ pressure gradient
coefficient λθ defined as:

λθ = (θ2/ν)dU/ds (18)

where dU/ds is the acceleration in the streamwise direction.

The transition model interacts with the SST turbulence model7, as follows:

∂ ∂ ∂  
( ρk ) +
~ ~
( ρu j k ) = Pk − Dk +  (µ + σ k µ t ) ∂k  (19)
∂t ∂x j ∂x j  ∂x j 
 
~ ~
Pk = γ eff Pk ; Dk = min(max(γ eff ,0.1),1.0)Dk (20)

8
 Ry 
− 
ρy k 120 
Ry = ; F3 = e   ; F1 = max (F1orig , F3 ) (21)
µ

where Pk and Dk are the original production and destruction terms for the SST model and F1orig is the original SST
blending function. Note that the production term in the ω-equation is not modified.

In order to capture the laminar and transitional boundary layers correctly, the grid must have a y+ of approximately
one. If the y+ is too large (i.e. > 5) than the transition onset location moves upstream with increasing y+. All
simulations have been performed with the CFX CFD code using a bounded 2nd order upwind biased discretisation
for the mean flow, turbulence and transition equations. For the case where system rotation is present (i.e. a 3D wind
turbine rotor) a moving frame approach is used to solve the equations. The equations where solved in steady-state
mode and the forces (i.e. lift and drag) where monitored in order to judge convergence. A relatively large time step
of 1/5 the airfoil characteristic time (i.e. chord/relative freestream velocity/5) was specified and this was possible
due to the implicit nature of the CFX code. This time step was found to be small enough to give relatively good
convergence for the equation residuals while being large enough to damp out any unsteadiness, which would have
been significant in the high angle of attack cases where a large amount of separated flow was present

5
American Institute of Aeronautics and Astronautics
III. Results and Discussion

A. 2D S809 Airfoil Results

The S809 airfoil is a 21% thick, laminar-flow airfoil


that was designed specifically for horizontal-axis wind
turbine (HAWT) applications8a,b. The airfoil profile is
shown in Figure 1. The experimental results were
obtained in the low-turbulence wind tunnel at the Delft
University of Technology.

In this study the experimental results are compared to


the numerical results obtained with the Navier-Stokes
code CFX using both the transition model and fully
turbulent computations. The CFD grid for this case
consisted of a C-type mesh with approximately 750
nodes around the airfoil (300 on each side and 150 in
the wake), 100 nodes normal to the airfoil and a Figure 1 S809 Airfoil Profile
maximum y+ of one. The farfield boundary was
located ten chord lengths away from the airfoil. Also
included for comparison are results obtained with the
well-known X-Foil coupled panel method/boundary
layer code10 that is often used in the wind energy
Laminar
community to evaluate airfoil performance
Separation
parameters. X-Foil uses an eN method for transition
prediction and is widely regarded as one of the best
tools available for predicting transition on 2D airfoils. Turbulent
All calculations were made assuming incompressible Reattachment
flow and a Reynolds number of 2×106. For the CFD
calculations an inlet turbulence level was selected so
that the freestream turbulence level around the airfoil
was approximately 0.2%, which would correspond to
the turbulence level in a typical low-turbulence wind
tunnel. The X-Foil results were computed with the
default N-factor of 9.
Figure 2 Pressure Distribution (Cp) for AoA = 1°
The predicted pressure distribution around the airfoil
for angles of attack (AoA) of 1°, 9°, 14° and 20° are
shown in Figures 2 – 5. For the 1° AoA case the flow
is laminar for the first 0.5 chord of the airfoil on both
the suction and pressure sides. The boundary layers
then undergo a laminar separation and reattach as a
turbulent boundary layer and this is clearly visible in
the experimental pressure distribution plateaus. The
fully turbulent computation obviously does not capture
this phenomenon, as the turbulent boundary layers
remain completely attached. Both the transitional
CFD and X-Foil solutions do predict the laminar
separation bubble. However, X-Foil appears to over-
predict the reattachment location while the transitional
CFD is in very good agreement with the experiment.
A very small laminar separation/turbulent
reattachment is also visible on the pressure side for
Figure 3 Pressure Distribution (Cp) for AoA = 9°
6
American Institute of Aeronautics and Astronautics
both the 9° and 14° cases. On the suction side
transition occurs in the adverse pressure gradient
before separating and as a result the pressure
distribution is virtually identical between the turbulent
and transitional CFD results for the 9° and 14° cases.
For the 20° case there is a large turbulent trailing edge
separation on the suction side that starts at the 0.45
chord location. Upstream of the turbulent separation
there is a significant difference between the
computations and the experiment and this could be
caused by 3D wind tunnel effects or alternatively
large scale unsteadiness and turbulence modeling
errors in the massively separated region.

The predicted transition locations as a function of


angle of attack are shown in Figure 6. The Figure 4 Pressure Distribution (Cp) for AoA = 14°
experimental transition locations were obtained using
a stethoscope method8b. In general the present
transition model would appear to be in somewhat
better agreement with the experiment than the X-Foil
code, particularly around 14° angle of attack.
However, at the moderate angles of attack all of the
results appear be to within approximately 5% chord of
each other. The X-Foil transition locations appear to
change quite rapidly over a few degrees angle of
attack while the transition model has a much smoother
change in the transition location. The experimental
data would appear to confirm that the smooth change
in transition location is more physical, however this
observation is based primarily on the 10° and 14°
angle of attack cases.

The results obtained for the lift and drag polars are Figure 5 Pressure Distribution (Cp) for AoA = 20°
shown in Figures 7 and 8. Between 0° and 9° the lift
coefficients (Cl) predicted by the transitional CFD
results are in very good agreement with the
experiment while both the fully turbulent CFD and X-
Foil results appear to under-predict the lift curve by
approximately 0.1. As well, between 0° and 9° the
drag coefficient (Cd) predicted by the transitional
CFD and X-Foil results are in very good agreement
with the experiment while the fully turbulent CFD
over-predicts the drag. For angles of attack greater
than 9° all of the numerical results significantly over-
predict the lift coefficient and under-predict the drag
coefficient. One potential explanation for the poor
numerical results could be the presence of a
significant amount of 3D flow in the separated regions
of the wind tunnel model, which the 2D numerical
results obviously cannot resolve. As well, it is well
known that the RANS turbulence models typically fail
in massively separated regions and it is possible that a
hybrid RANS-LES model such as DES or SAS may Figure 6 Transition location (xt/c) vs angle of attack
be needed to capture the separated flow physics for the S809 airfoil
correctly11.

7
American Institute of Aeronautics and Astronautics
Figure 7 Lift Coefficient (Cl) Polar (top) and enlarged Figure 8 Drag Coefficient (Cd) Polar (top) and
view from 0° to 10° (bottom) enlarged view from 0° to 10° (bottom)

B. 3D NREL Phase VI Wind Turbine (S809 Airfoil


section) Results

The NREL Phase VI experiment consisted of a two-


bladed twisted and tapered 10-meter diameter wind
turbine that used the S809 airfoil section. It was tested
in the NASA Ames Research Center 80-by-120 Foot
Wind Tunnel and a large matrix of performance and load
data was collected12. In this study the main focus was
only to compare the predicted torque output to the
measured value and to investigate the differences
between fully turbulent and transitional CFD solutions.
The CFD grid for this case consisted of an O-type mesh
of approximately 10 million nodes with a maximum y+
of one and was created using the ICEM grid generator.
The grid used to compute the NREL wind turbine is
illustrated in Figure 9. The second blade was modeled
using a periodic boundary condition at the hub and
farfield boundary conditions were located at
approximately twice the rotor diameter. The inlet
turbulence intensity specified for the CFD computations
Figure 9 Grid used to compute the NREL Wind Turbine.
was 0.2%. Fully turbulent and transitional CFD runs
Surface grid on the rotor (left) and close up of the volume grid
were made for a wind speeds between 7 m/s and 25 m/s
near the tip (right).

8
American Institute of Aeronautics and Astronautics
(i.e. increasing angle of attack) and each run was
computed overnight on a 16 CPU Linux cluster.

The suction side streamlines and intermittency


predicted by the transitional CFD computations for
wind speeds of 7m/s, 10m/s and 20m/s are shown in
Figures 10 and 11. As the wind speed is increased the
effective angle of attack of the wind turbine increases.
For the 7 m/s wind speed the flow is largely attached
on the suction side. As well, a significant amount of
laminar flow is predicted near the tip as well as in the
hub region. In the tip region transition occurs at the
0.5 chord position where as in the middle of the blade
span transition occurs near the leading edge. This is
most likely caused by the fact that the smaller radius
results in an increased effective angle of attack. For Figure 10 NREL wind turbine, suction surface streamlines
the 10 m/s case the inner hub region stalls while the computed with the transition model for wind speeds of 7m/s
tip region remains attached up until the 0.5 chord (top), 10m/s (middle) and 20 m/s (bottom). Arrows indicate
position. As well, due to the increased angle of attack flow direction.
the transition location near the tip moves to the
leading edge. Finally, at 20 m/s the suction side of the Laminar
blade is completely separated and the intermittency
contours indicate that the flow is almost completely
turbulent.

Figure 12 shows the predicted shaft torque in Turbulent


comparison with the experimental measurements. In
general the agreement between the transitional CFD
and experiment is quiet good considering how
complex the flow field is on the suction side of the
wind turbine. At a wind speed of 20 m/s there is a
large difference between the transitional and fully
turbulent shaft torque. Figure 13 is a contour plot of
the local torque on the suction side computed for the
fully turbulent and transitional solutions at this wind
speed. The fully turbulent solution predicts much Figure 11 NREL wind turbine, suction surface intermittency
higher torque levels near the leading edge and also in computed by the transition model for wind speeds of 7m/s
the hub region. (top), 10m/s (middle) and 20 m/s (bottom).

The most likely explanation for this difference is the


state of the boundary layer near the leading edge. At
the 20 m/s wind speed the stagnation point is located
on the pressure side and the suction side boundary
layer must first travel up to and around the leading
edge, at which point after the leading edge the
boundary layer separates. For the transitional cases
the boundary layer on the pressure side was laminar as
it traveled from the stagnation point around the
leading edge. It is well known that a laminar
boundary layer will separate earlier than a turbulent
boundary layer in a strong adverse pressure gradient.
Because of this, it would appear that as the boundary
layer traveled from the stagnation point around the
leading edge, the laminar boundary layer was not able
to sustain as high a suction peak as the turbulent
boundary layer. This would seem to be confirmed in Figure 12 NREL wind turbine, predicted shaft torque
at different wind speeds.
9
American Institute of Aeronautics and Astronautics
Figure 13, which is a plot of the pressure distribution (Cp)
at the 30 percent span location near the hub. The
maximum suction peak Cp near the leading edge for the
turbulent case was 9 while for the transitional case it was
5. After the suction peak the fully turbulent solution
significantly over-predicts the Cp while the transitional
solution is flat (indicating a strong separation after the
leading edge) and is in relatively good agreement with the
experimental measurements. It would appear that this
could explain the 80% difference in output torque between
the fully turbulent and transitional cases. It should be
noted that other researchers have performed fully turbulent
CFD computations of the NREL wind turbine and obtained
Figure 13 NREL wind turbine, local suction surface torque
relatively good results (e.g. Ref. 12). The separation near
computed fully turbulent (top) and with the transition model
the leading edge is probably highly sensitive to the choice
(bottom) for a wind speed of 20 m/s.
of turbulence model. The present fully turbulent results
are thought to be accurate because they have been obtained
with the well-known SST turbulence model, which is
generally regarded as one of the best two-equation models
available for predicting turbulent separation.

In Figure 15 the lift coefficient extracted from the 3D


NREL wind turbine (which used the S809 airfoil) at 90%
span is compared to the 2D S809 lift coefficient results.
The angle of attack for the NREL results was computed
from the CFD solution based on the angle between the
blade and the relative velocity one blade chord upstream of
the NREL wind turbine. What is particularly surprising is
that there is a very large difference between the 3D and 2D
lift coefficients, even at 9° angle of attack. The differences
could be related to finite span effects or alternatively, the
3D nature of the separated flow in the case of the NREL
turbine. If it is due to the separated flow this would seem
to raise a lot of issues with approaches such as the Blade Figure 14 Pressure Distribution (Cp) for the NREL wind
Element Method (BEM) that are based on using 2D airfoil turbine at the 30 percent span location for a wind speed of
data in order to predict 3D wind turbine performance, 20 m/s.
particularly near stall conditions.

IV. Conclusion

Recently, a new correlation based transition model has


been developed, which is strictly formulated based on
local variables and thereby compatible with modern CFD
methods such as unstructured grids and massive parallel
execution. In this paper, fully turbulent and transitional
computations of the 2D S809 airfoil along with a full 3D
wind turbine (that uses the S809 airfoil) have been
accomplished. The transitional results are in good
agreement with the experimental data and the transition
model would appear to be well suited for the prediction of
wind turbine aerodynamics. It is the author’s opinion that
the present transition model will allow the 1st order effects
of transition to be included into everyday industrial
computations of wind turbines. Figure 15 Lift Coefficient (Cl) Polar for the S809 airfoil
and the NREL wind turbine (extracted at 90% span near
the tip).
10
American Institute of Aeronautics and Astronautics
Acknowledgments

The model development and validation at ANSYS CFX was funded by GE Aircraft Engines and GE Global
Research. Prof. G. Huang and Dr. B. Suzen from the University of Kentucky have supported the original model
development with their extensive know-how and their in-house codes.

References
1
Wolfe, W.P. and Ochs, S.S. “CFD Calculations of S809 Aerodynamic Characteristics”, AIAA Paper AIAA-97-0972, 1997.
2
Langtry, R.B., and Menter, F.R., “Transition Modeling for General CFD Applications in Aeronautics”, AIAA Paper 2005-
522, 2005.
3
Van Driest, E.R. and Blumer, C.B., “Boundary Layer Transition: Freestream Turbulence and Pressure Gradient Effects,”
AIAA Journal, Vol. 1, No. 6, June 1963, pp. 1303-1306.
4
Menter, F. R., Esch, T. and Kubacki, S., “Transition Modelling Based on Local Variables”, 5th International Symposium on
Turbulence Modeling and Measurements, Spain, 2002.
5
Menter, F.R., Langtry, R.B., Likki, S.R., Suzen, Y.B., Huang, P.G., and Völker, S., "A Correlation based Transition Model
using Local Variables Part 1- Model Formulation”, ASME-GT2004-53452, ASME TURBO EXPO 2004, Vienna, Austria.
6
Langtry, R.B., Menter, F.R., Likki, S.R., Suzen, Y.B., Huang, P.G., and Völker, S., "A Correlation based Transition Model
using Local Variables Part 2 – Test Cases and Industrial Applications”, ASME-GT2004-53454, ASME TURBO EXPO 2004,
Vienna, Austria.
7
Menter, F.R., Two-equation eddy-viscosity turbulence models for engineering applications, AIAA Journal, Vol. 32, No. 8,
1994, pp. 1598-1605.
8a
Somers, D.M., “Design and Experimental Results for the S809 Airfoil”, Airfoils, Inc., State College, PA, 1989.
8b
Somers, D.M., “Design and Experimental Results for the S809 Airfoil”, NREL/SR-440-6918, January 1997.
9
Simms, D., Schreck, S., Hand, M, and Fingersh, L.J. , “NREL Unsteady Aerodynamics Experiment in the NASA-
Ames Wind Tunnel: A Comparison of Predictions to Measurements.”, NREL Technical report, NREL/TP-500-29494, 2001.
10
Drela, M., and Giles, M. B., “Viscous-Inviscid Analysis of Transonic and Low Reynolds Number Airfoils”, AIAA Journal,
Vol. 25, No. 10, October 1987, pp. 1347 – 1355.
11
Menter, F.R., and Egorov, Y., “A Scale-Adaptive Simulation Model using Two-Equation Models” AIAA Paper 2005-1095,
2005.
12
Duque, E.P.N., Burklund, M.D., Johnson, W., “Navier-Stokes and Comprehensive Analysis Performance Predictions of the
NREL Phase VI Experiment”, AIAA Paper 2003-0355, 2003.

11
American Institute of Aeronautics and Astronautics

You might also like