Langtry 2009
Langtry 2009
Robin B. Langtry∗
The Boeing Company, Seattle, Washington 98124-2207
and
Florian R. Menter†
ANSYS Germany, 83624 Otterfing, Germany
DOI: 10.2514/1.42362
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 computational fluid dynamics 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. A number of validation papers have
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
been published on the basic formulation of the model. However, until now the full model correlations have not been
published. The main goal of the present paper is to publish the full model and release it to the research community so
that it can continue to be further validated and possibly extended or improved. Included in this paper are a number of
test cases that can be used to validate the implementation of the model in a given computational fluid dynamics code.
The authors believe that the current formulation is a significant step forward in engineering transition modeling, as it
allows the combination of transition correlations with general-purpose computational fluid dynamics codes. There is
a strong potential that the model will allow the first-order effects of transition to be included in everyday industrial
computational fluid dynamics simulations.
(which may or may not reattach). As well, a turbulent boundary layer that none of these methods can meet all of the aforementioned
can relaminarize under the influence of a strong favorable pressure requirements.
gradient [4]. Although the importance of transition phenomena for The only transition models that have historically been compatible
aerodynamic and heat transfer simulations is widely accepted, it is with modern CFD methods are the low-Re models [10,11]. However,
difficult to include all of these effects in a single model. they typically suffer from a close interaction with the transition
The second complication arises from the fact that conventional capability and the viscous sublayer modeling, and this can prevent an
Reynolds-averaged Navier–Stokes (RANS) procedures do not lend independent calibration of both phenomena [12,13]. At best, the low-
themselves easily to the description of transitional flows, where both Re models can only be expected to simulate bypass transition which
linear and nonlinear effects are relevant. RANS averaging eliminates is dominated by diffusion effects from the freestream. This is because
the effects of linear disturbance growth and is therefore difficult to the standard low-Re models rely exclusively on the ability of the wall
apply to the transition process. Although methods based on the damping terms to capture the effects of transition. Realistically, it
stability equations such as the en method of Smith and Gamberoni [5] would be very surprising if these models that were calibrated for
and van Ingen [6] avoids this limitation, they are not compatible with viscous sublayer damping could faithfully reproduce the physics of
general-purpose CFD methods as typically applied in complex transitional flows. It should be noted that there are several low-Re
geometries. The reason is that these methods require a priori models where transition prediction was considered specifically
knowledge of the geometry and the grid topology. In addition, they during the model calibration [14–16]. However, these model formu-
involve numerous nonlocal operations (e.g., tracking the disturbance lations still exhibit a close connection between the sublayer behavior
growth along each streamline) that are difficult to implement into and the transition calibration. Recalibration of one functionality also
today’s CFD methods [7]. This is not to argue against the stability
changes the performance of the other. It is therefore not possible to
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
easily computed at each grid point in an unstructured, parallel For strong adverse pressure gradients, the difference between the
Navier–Stokes code. momentum-thickness and vorticity Reynolds number can become
A scaled profile of the vorticity Reynolds number is shown in significant, particularly near separation (H 3:5). However, the
Fig. 1 for a Blasius boundary layer. The scaling is chosen to have a trend with experiments is that adverse pressure gradients reduce the
maximum of one inside the boundary layer. This is achieved transition momentum-thickness Reynolds number. In practice, if a
by dividing the Blasius velocity profile by the corresponding constant transition momentum-thickness Reynolds number is
momentum-thickness Reynolds number and a constant of 2.193. In specified, the transition model is not very sensitive to adverse
other words, the maximum of the profile is proportional to the pressure gradients and an empirical correlation such as that of Abu-
momentum-thickness Reynolds number and can therefore be related Ghannam and Shaw [17] is necessary to predict adverse pressure
to the transition correlations [22] as follows: gradient transition accurately. In fact, the increase in vorticity
Reynolds number with increasing shape factor can actually be used
maxRev to predict separation-induced transition. This is one of the main
Re (2) advantages of the present approach because the standard definition
2:193 of momentum-thickness Reynolds number is not suitable in
separated flows.
Based on this observation, a general framework can be built, which The function Rev can be used on physical reasoning, by arguing
can serve as a local environment for correlation-based transition that the combination of y2 S is responsible for the growth of
models. disturbances inside the boundary layer, whereas = is res-
When the laminar boundary layer is subjected to strong pressure ponsible for their damping. As y2 S grows with the thickness of the
gradients, the relationship between momentum-thickness and boundary layer and stays constant, transition will take place once a
vorticity Reynolds number described by Eq. (2) changes due to the critical value of Rev is reached. The connection between the growth
change in the shape of the profile. The relative difference between of disturbances and the function Rev was shown by Van Driest and
momentum-thickness and vorticity Reynolds number, as a function Blumer [21] in comparison with experimental data. As well, Langtry
of shape factor H, is shown in Fig. 2. For moderate pressure gradients and Sjolander [15] found that the location in the boundary layer
(2:3 < H < 2:9), the difference between the actual momentum- where Rev was largest correspondeds surprisingly well to the loca-
thickness Reynolds number and the maximum of the vorticity Rey- tion where the peak growth of disturbances was occurring, at least for
nolds number is less than 10%. Based on boundary-layer analysis, a bypass transition. The models proposed by Langtry and Sjolander
shape factor of 2.3 corresponds to a pressure gradient parameter of [15] and Walters and Leylek [16], use Rev in physics-based
approximately 0.06. Because the majority of experimental data on arguments based on these observations of disturbance growth in the
transition in favorable pressure gradients falls within that range (see, boundary layer during bypass transition. These models appear
for example, [17]), the relative error between momentum-thickness superior to conventional low-Re models, as they implicitly contain
and vorticity Reynolds number is not of great concern under those information of the thickness of the boundary layer. Nevertheless, the
conditions. close integration of viscous sublayer damping and transition
prediction does not easily allow for an independent calibration of
both submodels.
In the present approach first described in [22–25], the main idea is
to use a combination of the strain-rate Reynolds number with
experimental transition correlations using standard transport equa-
tions. Because of the separation of viscous sublayer damping and
transition prediction, the new method has provided the flexibility for
introducing additional transition effects with relative ease. Currently,
the main missing extensions are crossflow instabilities and high-
speed flow correlations and these do not pose any significant obsta-
cles. The concept of linking the transition model with experimental
data has proven to be an essential strength of the model and this is
difficult to achieve with closures based on a physical modeling of
these diverse phenomena.
The present transition model is built on a transport equation for
intermittency, which can be used to trigger transition locally. In
addition to the transport equation for the intermittency, a second
Fig. 1 Scaled vorticity Reynolds number Rev profile in a Blasius transport equation is solved for the transition onset momentum-
boundary layer. thickness Reynolds number. This is required to capture the non-
LANGTRY AND MENTER 2897
local influence of the turbulence intensity, which changes due to the The transport equation for the intermittency reads
decay of the turbulence kinetic energy in the freestream, as well as
due to changes in the freestream velocity outside the boundary layer. @ @Uj @ @
P E t (3)
This second transport equation is an essential part of the model as it @t @xj @xj f @xj
ties the empirical correlation to the onset criteria in the intermittency
equation. Therefore, it allows the model to be used in general The transition sources are defined as follows:
geometries and over multiple airfoils, without additional information
P1 Flength ca1 SFonset 0:5 1 ce1 (4)
on the geometry. The intermittency function is coupled with the shear
stress transport (SST) k-!-based turbulence model [26]. It is used to
where S is the strain-rate magnitude. Flength is an empirical cor-
turn on the production term of the turbulent kinetic energy
relation that controls the length of the transition region, and FOnset
downstream of the transition point based on the relation between
controls the transition onset location. Both are dimensionless
transition momentum-thickness and strain-rate Reynolds number. As
functions that are used to control the intermittency equation in the
the strain-rate Reynolds number is a local property, the present
boundary layer. The destruction/relaminarization source is defined as
formulation avoids another very severe shortcoming of the
follows:
correlation-based models, namely their limitation to 2-D flows. It
therefore allows the simulation of transition in 3-D flows originating E ca2 Fturb ce2 1 (5)
from different walls. The formulation of the intermittency has also
been extended to account for the rapid onset of transition caused by where is the vorticity magnitude. The transition onset is controlled
separation of the laminar boundary layer [Eq. (17)]. In addition, the by the following functions:
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
8
> 1 4 ~ 6 ~ 2 ~ t < 400
> 398:189 10 119:270 10 Ret 132:567 10 Ret ;
>
<
Re
2 ~ 5 ~ 2 ~ 3t ; ~ t < 596
263:404 123:939 10 Ret 194:548 10 Ret 101:695 108 Re 400 Re
Flength (12)
>
> ~ 4
0:5 Ret 596:0 3:0 10 ; 596 Re ~ t < 1200
>
: 0:3188; 1200 Re ~ t
number given in [23] was called CFX-v-1.0. Based on this naming In certain cases, such as transition at higher Reynolds numbers, the
convention, the present model with the preceding modifications will ~ t scalar will often decrease to very small values in the boundary
Re
be referred to as the -Re model, CFX-v-1.1. The present transition ~ t , this can
layer shortly after transition. Because Flength is based on Re
model is briefly summarized in the following pages. result in a local increase in the source term for the intermittency
2898 LANGTRY AND MENTER
equation, which in turn can show up as a sharp increase in the skin Outside the boundary layer, the source term Pt is designed to force
friction. The skin friction does eventually return back to the fully ~ t to match the local value of Ret calculated
the transported scalar Re
turbulent value, however, this effect is unphysical. It appears to be from the empirical correlation [Eqs. (35) and (36)]. The source term
caused by a sharp change in the y in the viscous sublayer where the is defined as follows:
intermittency decreases back to its minimum value due to the
destruction term [Eq. (5)]. The effect can be eliminated by forcing Pt ct Ret Re ~ t 1:0 Ft (22)
Flength to always be equal to its maximum value (in this case 40.0) in t
the viscous sublayer. The modification for doing this is shown next.
The modification does not appear to have any effect on the predicted 500
t (23)
transition length. An added benefit is that, at higher Reynolds num- U2
bers, the model now appears to predict the skin friction overshoot
measured by experiments: where t is a time scale, which is present for dimensional reasons. The
time scale was determined based on dimensional analysis with the
R! 2 main criteria being that it had to scale with the convective and
Fsublayer e 0:4 (13)
diffusive terms in the transport equation. The blending function Ft is
used to turn off the source term in the boundary layer and allow the
y2 ! ~ t to diffuse in from the freestream. Ft is equal
transported scalar Re
R! (14)
500 to zero in the freestream and one in the boundary layer. The Ft
blending function is defined as follows:
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
8
>
> ~ t 396:035 102 120:656 104 Re~ t 868:230 106 Re
~ 2t
< Re
Rec 9 ~ 3 12 ~ 4
696:506 10 Ret 174:105 10 Ret ; ~ t 1870
Re (16)
>
>
: Re
~ t 593:11 Re
~ t 1870:0 0:482; ~ t > 1870
Re
dU 1 2 1 du dv dw A minimum turbulence intensity of 0.027% results in a transition
u v2 w2 2 2u 2v 2w (31) momentum-thickness Reynolds number of 1450, which is the largest
dx 2 dx dx dx
experimentally observed flat-plate transition Reynolds number based
on the Sinclair and Wells [29] data. For cases where larger transition
Reynolds number are believed to occur (e.g., aircraft in flight) this
dU 1 2 1 du dv dw limiter may need to be adjusted downward.
u v2 w2 2 2u 2v 2w (32)
dy 2 dy dy dy The empirical correlation is used only in the source term [Eq. (22)]
of the transport equation for the transition onset momentum-
thickness Reynolds number. Equations. (35–38) must be solved
iteratively because the momentum thickness t is present in the left-
dU 1 2 1 du dv dw hand side of the equation and also in the right-hand side in the
u v2 w2 2 2u 2v 2w (33)
dz 2 dz dz dz pressure gradient parameter . In the present work, an initial guess
for the local value of t was obtained based on the zero pressure
gradient solution of Eqs. (35) and (36) and the local values of U, ,
and . With this initial guess, Eqs. (35–38) were solved by iterating
dU dU dU dU
u=U v=U w=U (34) on the value of t and convergence was obtained in less then 10
ds dx dy dz iterations using a shooting point method.
The transition model interacts with the SST turbulence model [26]
The use of the streamline direction is not Galilean invariant.
However, this deficiency is inherent to all correlation-based models, as follows:
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
Table 1 Inlet condition for the flat-plate test cases at 0.04 m upstream of plate leading edge
Case Inlet velocity, m=s Turbulence intensity, % t = Density, kg=m3 Dynamic viscosity, kg=ms
Inlet/leading edge value
T3A 5.4 3.3 12.0 1.2 1:8 105
T3B 9.4 6.5 100.0 1.2 1:8 105
T3A- 19.8 0.874 8.72 1.2 1:8 105
Schubauer and Klebanof [28] 50.1 0.3 1.0 1.2 1:8 105
T3C2 5.29 3.0 11.0 1.2 1:8 105
T3C3 4.0 3.0 6.0 1.2 1:8 105
T3C4 1.37 3.0 8.0 1.2 1:8 105
T3C5 9.0 4.0 15.0 1.2 1:8 105
Relaminarization 1.4 5.5 15 1.2 1:8 105
viscosity ratio (i.e., the ! inlet condition) to match the experimentally boundary layer with a turbulence intensity of 3%. The same effect
measured turbulence levels at various downstream locations. As the could have been accomplished with a small step or gap in the CFD
freestream turbulence increases, the transition location moves to geometry. Downstream of the trip, the boundary layer slowly rela-
lower Reynolds numbers. minarizes due to the strong favorable pressure gradient.
The T3C test cases consist of a flat plate with a favorable and For all of the flat-plate test cases, the agreement with the data is
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
adverse pressure gradient imposed by the opposite converging/ generally good, considering the diverse nature of the physical phe-
diverging wall. The wind-tunnel Reynolds number was varied for the nomena computed, ranging from bypass transition to natural tran-
four cases (T3C5, T3C3, T3C2, T3C4), thus moving the transition sition, separation-induced transition, and even relaminarization.
location from the favorable pressure at the beginning of the plate to
the adverse pressure gradient at the end. The cases are used to
demonstrate the transition models ability to predict transition under B. Turbomachinery Test Cases
the influence of various pressure gradients. Figure 5 details the results This section describes a few of the turbomachinery test cases that
for the pressure gradient cases. The effect of the pressure gradient on have been used to validate the transition model including a com-
the transition length is clearly visible with favorable pressure gra- pressor blade, a low-pressure turbine, and a high-pressure turbine. A
dients increasing the transition length and adverse pressure gradients summary of the inlet conditions is shown in Table 2.
reducing it. For the T3C4 case, the laminar boundary layer actually For the Zierke and Deutsch [31] compressor blade, transition on
separates and undergoes separation-induced transition. the suction side occurs at the leading edge due to a small leading-edge
The relaminarization test case is shown in Fig. 6. For this case, the separation bubble on the suction side. On the pressure side, transition
opposite converging wall imposes a strong favorable pressure occurs at about midchord. The turbulence contours and the skin
gradient that can relaminarize a turbulent boundary layer. In both the friction distribution are shown in Fig. 7. There appears to be a signi-
experiment and in the CFD prediction, the boundary layer was ficant amount of scatter in the experimental data; however, overall,
tripped near the plate leading edge. In the CFD computation, this was the transition model is predicting the major flow features correctly
accomplished by injecting a small amount of turbulent air into the (i.e., fully turbulent suction side, transition at midchord on the
Fig. 4 Results for flat-plate test cases with different freestream turbulence levels.
LANGTRY AND MENTER 2901
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
Fig. 5 Results for flat-plate test cases where variation of the tunnel Reynolds number causes transition to occur in different pressure gradients (dp=dx).
Case Rex cUo =, 106 Mach Uo =a where speed of Chord c, m FSTI, % t =
sound a RT0:5
Zierke and Deutsch [31] compressor 0.47 0.1 0.2152 0.18 2.0
incidence 1:5 deg
Pak-B low-pressure turbine blade 0.05, 0.075, 0.1 0.03 0.075 0.08, 2.35, 6.0 6.5–30
VKI MUR transonic guide vane 0.26 Inlet: 0.15 0.037 1.0, 6.5 11, 1000
Outlet: 1.06
2902 LANGTRY AND MENTER
the ability of the model to capture the effects of Reynolds number and
turbulence intensity variations on the size of a laminar separation
bubble and the subsequent turbulent reattachment.
The surface heat transfer for the transonic von Karman Institute
(VKI) MUR 241 (FSTI 6:0%) and MUR 116 (FSTI 1:0%) test
cases [34] is shown in Fig. 10. The strong acceleration on the suction
side for the MUR 241 case keeps the flow laminar until a weak shock
at midchord, whereas for the MUR 116 case, the flow is laminar until
right before the trailing edge. Downstream of transition there appears
to be a significant difference between the predicted turbulent heat
transfer and the measured value. It is possible that this is the result of a
Mach number (inlet Mach number Mainlet 0:15, Maoutlet 1:089)
effect on the transition length [35]. At present, no attempt has been
made to account for this effect in the model. It can be incorporated in
future correlations, if found consistently important.
The pressure side heat transfer is of particular interest for this
case. For both cases, transition did not occur on the pressure side,
however, the heat transfer was significantly increased for the high
turbulence intensity case. This is a result of the large freestream
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
Fig. 9 Blade loading for the Pak-B low-pressure turbine at various FSTI and Reynolds numbers Re.
measured using hot films on the upper surface of the slat and flap and
on both the upper and lower surfaces of the main element. The skin
friction was also measured at various locations using a Preston tube
[39]. For the present comparison, the Reynolds number Re 9
106 and an angle of attack AoA 8 deg was selected. The free-
stream conditions for k and ! were selected to match the transition
location at the suction side of the slat. The other transition locations
are an outcome of the simulation.
A contour plot of the predicted turbulence intensity around the flap
is shown in Fig. 16. Also indicated are the various transition locations
that were measured in the experiment (Exp.) as well as the locations
predicted by the present transition model (CFD). In the compu-
tations, the onset of transition was judged as the location where the
skin friction first started to increase due to the production of turbulent
kinetic energy in the boundary layer. In general, the agreement
Fig. 10 Heat transfer for the VKI MUR 241 (FSTI 6:0%) and between the measured and predicted transition locations is very
MUR 116 (FSTI 1:0%) test cases. good. The largest error was observed on the lower surface of the main
element where the predicted transition location was too far down-
stream by approximately 6% of the cruise-airfoil chord.
changes in pressure gradient and the varying local freestream The DLR F-5 geometry is a 20 deg swept wing with a symmetrical
turbulence intensity around the different lifting surfaces. airfoil section that is supercritical at a freestream Mach number of
The experiment was performed in NASA Langley’s low- 0.82. The experiment was performed at the DLR by Sobieczky [40]
turbulence pressure tunnel and the transition locations were and consists of a wing mounted to the tunnel sidewall (which is
IV. Conclusions
In this paper, various methods for transition prediction in general-
purpose CFD codes have been discussed. In addition, the require-
ments that a model has to satisfy to be suitable for implementation
into a general-purpose CFD code have been listed. The main criterion
is that nonlocal operations must be avoided. A new concept of
transition modeling termed local correlation-based transition model
(LCTM) was introduced. It combines the advantages of locally
formulated transport equations with the physical information
contained in empirical correlations. The Req transition model is
representative of that modeling concept. The model is based on two
new transport equations (in addition to the k and ! equations), one for
Fig. 13 Transition location (xt=c) vs angle of attack for the S809 airfoil.
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.
An overview of the -Req model formulation has been given along
with the publication of the full model including some previously
undisclosed empirical correlations that control the predicted tran-
sition length. The main goal of the present paper was to publish the
full model and release it to the research community so that it can
continue to be further validated and possibly extended. Included in
this article are a number of test cases that can be used to validate the
implementation of the model in a given CFD code.
The present transition model accounts for transition due to free-
stream turbulence intensity, pressure gradients, and separation. It is
fully CFD-compatible and does not negatively affect the conver-
gence of the solver. Current limitations of the model are that
Fig. 14 Lift coefficient Cl polar for the S809 airfoil. crossflow instability or roughness are not included in the correlations
LANGTRY AND MENTER 2905
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362
Fig. 16 Contour of turbulence intensity Tu around the McDonnell Douglas 30P-30N flap as well as the measured (Exp.) and predicted (CFD) transition
locations (x=c) as a function of the cruise-airfoil chord (c 0:5588 m). Also indicated is the relative error between the experiment and the predicted
transition locations.
Fig. 17 DLR-F5 wing with transition: simulations (left), experiment (middle and right).
[2] Morkovin, M. V., “On the Many Faces of Transition,” Viscous Drag [22] Menter, F. R., Esch, T., and Kubacki, S., “Transition Modelling Based
Reduction, edited by C. S. Wells, Plenum, New York, 1969, pp. 1–31. on Local Variables,” Proceedings of the 5th International Symposium
[3] Malkiel, E., and Mayle, R. E., “Transition in a Separation Bubble,” on Engineering Turbulence Modelling and Measurements, Elsevier,
Journal of Turbomachinery, Vol. 118, No. 4, 1996, pp. 752–759. Amsterdam, 2002, pp. 555–564.
[4] Mayle, R. E., “The Role of Laminar-Turbulent Transition in Gas [23] Menter, F. R., Langtry, R. B., Likki, S. R., Suzen, Y. B., Huang, P. G.,
Turbine Engines,” Journal of Turbomachinery, Vol. 113, No. 4, 1991, and Völker, S., “A Correlation Based Transition Model Using Local
pp. 509–537. Variables Part 1: Model Formulation,” Journal of Turbomachinery,
doi:10.1115/1.2929110 Vol. 128, No. 3, 2006, pp. 413–422.
[5] Smith, A. M. O., and Gamberoni, N., “Transition, Pressure Gradient [24] Langtry, R. B., Menter, F. R., Likki, S. R., Suzen, Y. B., Huang, P. G.,
and Stability Theory,” Douglas Aircraft Co., Rept. ES 26388, Long and Völker, S., “A Correlation Based Transition Model Using Local
Beach, CA, 1956. Variables Part 2: Test Cases and Industrial Applications,” Journal of
[6] van Ingen, J. L., “A Suggested Semi-Empirical Method for the Turbomachinery, Vol. 128, No. 3, 2006, pp. 423–434.
Calculation of the Boundary Layer Transition Region,” Univ. of Delft, doi:10.1115/1.2184353
Dept. Aerospace Engineering, Rept. VTH-74, Delft, The Netherlands, [25] Menter, F. R., Langtry, R. B., and Völker, S., “Transition Modelling for
1956. General Purpose CFD Codes,” Flow, Turbulence and Combustion,
[7] Stock, H. W., and Haase, W., “Navier-Stokes Airfoil Computations with Vol. 77, Nos. 1–4, 2006, pp. 277–303.
eN Transition Prediction Including Transitional Flow Regions,” AIAA doi:10.1007/s10494-006-9047-1
Journal, Vol. 38, No. 11, 2000, pp. 2059–2066. [26] Menter, F. R., “Two-Equation Eddy-Viscosity Turbulence Models for
doi:10.2514/2.893 Engineering Applications,” AIAA Journal, Vol. 32, No. 8, 1994,
[8] Drela, M., and Giles, M. B., “Viscous-Inviscid Analysis of Transonic pp. 1598–1605.
and Low Reynolds Number Airfoils,” AIAA Journal, Vol. 25, No. 10, doi:10.2514/3.12149
[27] Langtry, R. B., “A Correlation-Based Transition Model Using Local
Downloaded by MONASH UNIVERSITY on November 20, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.42362