Sei sulla pagina 1di 20

AIAA-2005-0086

On the Prediction of Stall Onset on Airfoils at Moderately High Reynolds Number Flows
Zvi Rusak
*
and Wallace J. Morris II
**
Department of Mechanical, Aerospace, and Nuclear Engineering
Rensselaer Polytechnic Institute, Troy, NY 12180-3590, USA
Abstract
The inception of leading-edge stall on two-dimensional smooth thin airfoils at moderately
high Reynolds number flows is investigated by an asymptotic approach and numerical
simulations. The asymptotic theory is based on the work of Rusak (1994) and demonstrates
that a subsonic flow about a thin airfoil can be described in terms of an outer region,
around most of the airfoil chord, and an inner region, around the nose, that asymptotically
match each other. The flow in the outer region is dominated by the classical thin airfoil
theory. Scaled (magnified) coordinates and a modified (smaller) Reynolds number are used
to correctly account for the non linear behavior and extreme velocity changes in the inner
region, where both the near stagnation and high suction areas occur. It results in a model
(simplified) problem of a uniform flow past a semi-infinite parabola with a far-field
circulation governed by a parameter
~
A
that is related to the airfoils angle of attack, nose
radius of curvature, and camber and to the flow Mach number. The model parabola
problem consists of a compressible and viscous flow described by the Navier-Stokes
equations that is solved numerically for various values of
~
A
to determine
~
A
s
where a large
separation bubble first appears in the nose flow and the maximum suction suddenly drops.
The change of
~
A
s
with the modified Reynolds number is determined. These values indicate
the stall onset on the airfoil at various flow conditions. The predictions according to this
approach show good agreement with results from both full numerical simulations and much
available experimental data of the stall of thin airfoils. This simplified approach provides
for a universal criterion to determine the stall angle of airfoils with a parabolic nose and the
effect of airfoils thickness ratio, nose radius of curvature, camber and flaps and flow
compressibility on the onset of stall. This approach also presents an analysis method that
can be used to predict the stall of airfoils with its respective nose geometry.
I. Introduction
The stall of airfoils is a basic phenomenon in aerodynamics which refers to the significant and
sudden loss of lift, increase of drag, and change in pitching moment at a certain angle of attack (known
as the stall angle) and above. The occurrence of stall was first observed in flight and in the pioneering
wind-tunnel experiments of Jones
1
(see also the review paper of extensive experimental work on stall by
Tani
2
). For thin airfoils at low to moderately high Reynolds number flows, this phenomenon is associated
with the separation of the flow due to viscous effects and the adverse pressure gradient along the upper
surface of the airfoil near the leading edge. The flow separation results in the establishment of a
relatively large, re-circulating dead flow region above the upper surface. There, the pressure is higher
than in the attached flow case, thereby creating the loss of lift (McCormick,
3
pp 61-66, Anderson,
4
pp
282). The stall phenomenon limits the maximal lift of the airfoil and significantly
____________________________________
*
Professor, AIAA Associate Fellow.
**
Graduate Student, AIAA Student member.
affects the stability of flight and control of airplanes. The typical design objective is to delay separation
and stall to higher angles of attack and thereby increase the maximal lift (Liebeck and Ormsbee
5
and
Lieback
6
) and widen the operational speed envelope of flying vehicles and of helicopters, propellers and
turbines blades.
The study of the onset of stall was primarily based on many experimental investigations of airfoils at
moderate to relatively high Reynolds number flows, see, for example, the works of Jones,
1
Jacobs and
Sherman,
7
the classical book by Abbott and von Doenhoff,
8
Tani,
2
and the modern studies by Selig et al.
9
The experimental data provides the stall angle of attack for many airfoils with various geometries. It can
be seen that for most airfoils the stall angle increases with Reynolds number and slowly grows at very
high Reynolds number flows. Also, for thin airfoils with thickness ratio up to 15%, increasing the
thickness ratio and/or the nose radius of curvature increases the stall angle, thereby increasing the
maximum lift coefficient. Increasing the camber also increases the stall angle. Deflecting a trailing-edge
flap decreases the stall angle, but with a net gain in maximum lift. The experimental measurements also
show the existence of hysteresis loops between attached and separated flow states at near stall angles of
attack when the flow Reynolds number is sufficiently high (see Ref. 3, pp 141-146). The range of angles
of attack for these loops also increase with Reynolds number. At relatively low Reynolds numbers this
behavior disappears.
The experimental studies and flow visualizations demonstrate two types of the stall phenomenon
(see, for example, Anderson,
4
pp 330-337 and the flow pictures of Ito shown in Nakayama
10
). The first is
the leading-edge stall that characterizes relatively thin airfoils with thickness ratios up to 15% of the
airfoil chord where the flow separation zone suddenly evolves from the nose as the angle of attack is
increased. Numerical simulations demonstrate that leading-edge stall is preceded by the appearance of a
short recirculation bubble in the boundary layer at a certain angle of attack. With a small increase of
angle of attack, this bubble grows and than bursts into a large separation zone, causing the stall of the
airfoil. Here the loss of lift is rather rapid and abrupt. The second is the trailing-edge stall that
characterizes relatively thick airfoils with thickness ratio above 16% where the flow separation zone
develops gradually from the tail as the angle of attack is increased. Here the loss of lift is progressive and
slower. The experimental results for high Reynolds number flows also show that the maximal lift
coefficient increases with the thickness ratio up to about 15% and gradually decrease as the thickness
ratio is further increased. It is understood that the leading-edge stall is dominated by the interplay
between the relatively sharp nose geometry, viscous effects and adverse pressure gradients that lead to
boundary layer separation near the leading edge. On the other hand, the trailing-edge stall may be related
to flow instabilities that move upstream along the upper surface from the tail toward the nose and exhibit
hysteresis phenomena. For a given airfoil, the leading-edge stall occurs at moderate Reynolds numbers
and the trailing-edge stall at higher Reynolds numbers. As the thickness ratio is increased the transition
between the two types of stall occurs at a lower Reynolds number. In the present paper we focus on the
prediction of the onset of the leading-edge stall. This is still considered a difficult and challenging flow
problem to study using theoretical tools.
The experimental data also resulted in empirical guidelines that were used in the design of high lift
airfoils with leading-and trailing-edge flap systems (McCormick,
3
pp 85-109). Based on variational
calculus and a criterion for a separation pressure recovery in turbulent boundary layers, Liebeck and
Ormsbee
5
and Liebeck
6
designed high lift airfoils with prescribed optimum pressure distributions on the
upper and lower surfaces, such as the turbulent rooftop design. Only recently, with the advancement of
computational fluid dynamics, the phenomenon of stall can be computed and explored by very intensive
computations (see, for example, Hung et al,
11
Moulton and Steinhoff,
12
and Grove et al.
13
). The flow
simulations can predict the onset of stall and the separated flow field over single airfoils, multi airfoils
with flaps and other high lift devises, and over full aircraft configurations. These simulations become a
useful tool in the design of modern airfoils and wings.
In addition, the recent increased interest in micro-air vehicles (MAVs) raised the need for
understanding the stall characteristics of small scale airfoils at relatively low to moderate Reynolds
number flows. In such flows, where viscous effects dominate, flow separation and stall occur at lower
angles of attack and the maximal lift is smaller than in the large-scale airfoils (OMeara & Mueller
14
).
This significantly limits the performance of MAVs. It is clear that a theoretical approach to predict the
onset of stall in such flows will provide an essential design tool of small-scale airfoils. Note that for
very thin airfoils at very low Reynolds number flows the effects due to viscosity may change the
traditional trends known from the larger scales. For example, the numerical computations of Kunz and
Kroo
15
show that the increase of Reynolds number or the increase of thickness ratio may result in a
lower maximal lift coefficient. This regime of flow presents additional scientific and design
challenges.
Theoretical studies of the leading-edge separation at the nose of an airfoil were presented by Werle
and Davis,
16
Cebeci, Khattab and Stewartson,
17
Stewartson, Smith and Kaups,
18
Ruban
19
and Cheng and
Smith.
20
Werle and Davis
16
numerically solved the classical boundary-layer equations for an
incompressible flow past a parabola at angle of attack where the inviscid solution of the speed on the
parabola surface is prescribed as the outer speed of the boundary layer. They found that the shear stress
vanishes along the parabola surface and a Goldstein-type singularity of the stress and the boundary-layer
thickness are displayed when the vertical displacement of the stagnation point on the parabola exceeds a
certain value, K > 1.157, which is independent of the flow Reynolds number (Cebeci et al.
17
and
Stewratson et al.
18
found a similar result, K > K
c
=1.155). This value corresponds to a pressure gradient
parameter

below -0.2 (which also characterizes the separation of a laminar boundary layer on a flat
plate under an adverse pressure gradient). Note that according to our analysis, the matching of the
parabola flow with the classical solution of the thin airfoil theory (which satisfies the Kutta condition at
the trailing edge) requires that c R K
c
2 / / (here

is the angle of incidence, c the airfoils chord,


and
c
R
is the airfoils nose radius of curvature). This means that the condition for onset of nose
separation is c R
c c
2 / 157 . 1 . The approach of Werle and Davis
16
is limited to no interaction
between the boundary-layer flow and the imposed outer flow thereby it forces the singularity of the shear
stress and can not accurately describe the structure of a separated flow state. Hence, it results in a too low
prediction of the angle
c

for the onset of separation with respect to experimental findings.


7-9

An elegant and more detailed analysis of the nose flow was conducted by Stewartson et al
18
(see also
a similar analysis by Ruban
19
). They extended the classical boundary-layer study of Werle and Davis
16
by
using an asymptotic analysis of the Navier-Stokes equations at high values of the Reynolds number
where Re=Uc/ >>1 and
16 / 1
Re /

<< c R
c
. They formulated a special version of the triple-deck theory
of the boundary-layer flow about the separation point, also known as the Marginal Separation theory.
This approach allows a weak interaction between the inner, nearly separated flow close to the wall and
the outer inviscid flow above the separation point. They identified the scales of the flow parameters near
the wall and its power law dependence on the modified Reynolds number Re
M
based on
c
R
, Re
M
=Re R
c
/c.
They also constructed an integral equation for the solution of the velocity structure near the wall and its
relationship to the pressure changes in the outer flow. The computed results show that attached-flow
states exist when

(=K in Werle and Davis


16
) is slightly greater than
157 . 1
c

. For
5 / 2
) (Re

+
M c
m , where m is a positive number, small closed separation bubbles appear when
4 . 2 > , they grow in size until
75 . 2
c
, where a fold in the solution branch occurs and larger
closed separation bubbles are found for a certain range of
c
<
. For
75 . 2 >
c
no solution is
found, suggesting the bursting of the separation bubble and that no steady separated flow state can exist
when the angle of attack is above c R m
c M c c s
2 / ] ) (Re [
5 / 2
+ . The existence of two separation
bubble states for a certain range of
c
<
suggests the possible hysteresis of the leading-edge separation
for a certain range of angles of attack below the stall angle, which may be related to the observed
hysteresis loops in the experiments.
3

The approach of Stewartson et al.
18
is limited to a weak interaction between the separated flow near
the wall and the outer flow and therefore can not predict stronger steady-state interactions, specifically
when
c
>
or

is above the predicted angle of stall


s

. Also, the formula for


s

predicts that the


stall angle and the range of the hysteresis loop decrease with Reynolds number but this is in direct
opposition to established data and observations which show the increase of angle of stall and the
hysteresis loops with Reynolds number.
3,7-9
To the best of our understanding, the formula for
s

does not
provide accurate predictions of the stall angle of practical airfoils. It also does not account for
geometrical attributes of practical airfoils such as camber and the deflection of flaps. Additional
arguments can be raised about the expected stability of the separated flow states along the branch of
solutions that folds back as decreases from
c

. Typically, fold points along a branch of steady state


solutions is also a point of exchange of stability of the states and it is expected that the states past the fold
point are unstable under realistic global boundary conditions and can not represent near-steady physical
states.
Despite being a basic and classical phenomenon in aerodynamics that limits the envelope of flight,
the onset of stall and its mechanism is not yet fully understood. To the best of our knowledge, there is no
theoretical approach that can yet accurately predict the stall angle of practical airfoils as function of the
flow Reynolds number. Moreover, the changes of the stall angle due to changes in the airfoils geometry,
such as camber and flap deflection as are described above from the experimental data of Jacobs and
Sherman,
7
Abbott and von Doenhoff,
8
and Selig et al.,
9
were never explained by any basic theoretical
study or model.
In this paper, we extend the work of Rusak
21
to study the leading-edge stall of thin airfoils at
subsonic speeds and moderately high Reynolds numbers. The asymptotic theory
21
extends Van Dyke
22
analysis to include compressible subsonic flows with a finite upstream Mach number. It also
demonstrates that a subsonic flow about a thin airfoil can be described in terms of an outer region,
around most of the airfoil chord, and an inner region, around the nose, that asymptotically match each
other in an intermediate overlap region all around the nose, unlike Van Dyke
22
approach that focused
only on the matching of the surface speed. The matching results in a model (simplified) problem of a
uniform stream described by the Navier-Stokes compressible flow equations past a semi-infinite parabola
with a far-field circulation governed by a parameter
~
A
(same as K from Werle and Davis
16
and from
Stewartson et al.
18
) that is related to the angle of attack, nose radius of curvature, and airfoils camber and
flow Mach number. The model parabola problem is solved numerically for various values of
~
A
to
determine
~
A
s
where a large separation zone first appears in the nose flow and the maximum suction
suddenly drops. The change of
~
A
s
with Reynolds number is computed. These values indicate the
leading-edge stall onset on the airfoil at various flow conditions. The predictions according to this
approach show good agreement with results from full numerical simulations and much available
experimental data of the inception of stall. The present simplified approach provides a universal criterion
for the onset of stall on thin airfoils and an analysis method that can be used to predict the onset of stall
on other airfoils using its respective nose geometry. Note that the theory is limited to thin airfoils at
moderate to high Reynolds number flows. The criterion does not apply to airfoils with thickness greater
than 15% where the trailing-edge separation dominates.
II. Theoretical Study
To investigate the onset of stall on thin airfoils, we developed a multi-scale matched asymptotic
analysis for a moderate Reynolds number (Re >> 1) compressible flow of a thermally perfect gas around
a thin airfoil at low and moderate angles of attack. The smooth airfoil geometry is given by y =
cF
u,l
(x/c) for 0 x c where c is the chord, the upper and lower surface shape functions F
u,l
(x/c) are
described by F
u,l
(x/c) = C
a
(x/c) - A x/c t(x/c) with 0 < << 1 and A=/. Here, C
a
(x/c) is the camber-
line function, t(x/c) is the thickness function and the airfoil nose is parabolic. The airfoil is given in a
uniform stream of air with subsonic Mach number M below critical for transonic effects, temperature

T
and pressure

p . The free-stream flow speed is

RT M U (

is ratio of specific heats and R is the


specific gas constant for air), density is ) RT /( p

, viscosity is ) T (

, specific heat is
p
C
,
and heat conduction coefficient is ) T (

.
Asymptotic expansions of the velocity components and the thermodynamic properties are
constructed in an outer region around most of the airfoil and in an inner region near the nose, in terms of
the airfoil small thickness ratio . The analysis extends the work of Rusak
21
on the inviscid subsonic flow
around a thin airfoil.
II.1. Outer expansion. In the outer region, we use the reference parameters c,

U ,

T ,

p ,

and

to scale the axial and vertical distances, velocity components, temperature, pressure, density,
viscosity, respectively. The steady-state dimensionless equations governing the flow are:

( ) ( )
.
,
Re
) 1 ( ) (
Pr
1
Re
1 1
, ) (
3
2
2
Re
1 1
, 0 ) (
2
2
T p
M
T p V T V
V V V p
M
V V
V
T


1
]
1

'

+ +


I
(1)
Here
V

is the dimensionless velocity vector,


,
p T,
,
are the dimensionless density, pressure,
temperature, viscosity, and heat conduction coefficient,

/ Re c U is the Reynolds number,

/ Pr
p
C
is the Prandtl number, I is the unit tensor, and is the dimensionless dissipation
function. Typically 1 Re >> and the effect of viscosity is localized to the very thin boundary layers and
plays only a secondary (or even smaller) role in determining the pressure distribution along the airfoil.
We also assume that the boundary layers thickness is much small than the airfoils thickness such that
2
/ 1 Re >> . Then, except for the very thin boundary layers near the airfoil surfaces and the airfoils
nose region, the flow in the outer region around most of the airfoil is dominated by the inviscid version
of Eqs. (1),

.
, 0
1
, 0
1
, 0 ) (
2
T p
p V T V
p
M
V V
V

(2)
Actually, since the upstream flow is uniform, the flow is also irrotational and potential. The solution of
(2) with the no penetration condition along the airfoil surfaces and the upstream uniform flow conditions
at the far field provides the far-field conditions for the boundary layers. Specifically, the velocity and
pressure distributions along the airfoil surfaces from this inviscid solution are also the velocity and
pressure distributions at the edge of the boundary layers.
Moreover, since typically the airfoils thickness ratio is small (0 < 1 << ), it can be shown (see
Rusak
21
) that the solution of (2) in the outer region (around most of the airfoil except for the nose region)
is dominated by the classical subsonic linear airfoil theory. Then, the pressure coefficient along most of
the airfoils upper (u) and lower (l) surfaces (in the range < x/c < 1) may be approximated in the
leading-order terms by:
( )
(3)

1
]
1


+
+

'

1
]
1


+
1
]
1


+ +

1 n
n
n
0
m
2 k
k 2
2 k
1 k 2 l , u
) n sin( w ) 1 (
sin
cos 1
w A ) 1 (
sin
) 2 sin( ) k 2 sin(
A 2
sin
sin ) 1 k 2 sin(
A 2 cos 2 1 h 2 ) c / x ( Cp

where ,
2
cos 1
c
x
0 < < ,

d n sin
c
x
t
n 2
A
0
n


,
_

d
c
a C
A
w


,
_


0
0
1
1
,

d n cos
c
a C
A
n 2
w
0
n


,
_


and
2
1 M
(the Prandle-Glauert compressibility parameter). Also,
m=1 for the upper surface and m=-1 for the lower surface. Note that the linear theory solution (3)
exhibits a singular behavior as x/c approaches the leading edge, i.e.

x c w 2
) 1 ( ~ ) c / x ( Cp
0 m
l , u
as
x/c approaches zero. Also note that according to the thin airfoil theory the airfoils lift coefficient is
( ) / 2
0l l
C
where the zero lift angle is given by
,
_

2 / 1
_
0
_
0
w w
l

and

d
c
a C w


,
_


0
0
_
1
,

d
c
a C w cos
2
0
1
_


,
_


.
II.2. Inner expansion. In the inner region around the airfoils nose (0 x/c < R
c
/c), scaled
coordinates and flow parameters are used to correctly describe the local non-linear behavior of the flow
which includes the extreme velocity changes due to the near-stagnation and suction areas around the
nose. There, we use the nose radius of curvature c h 2 R
2 2
c
to scale the axial and vertical coordinates,
x
*
= x/R
c
and y
*
= y/R
c
. The scaling parameters

U ,

T ,

p ,

and

are used for the velocity


components, temperature, pressure, density, and viscosity, respectively. The problem in the nose region
becomes in the leading order the flow of a compressible, viscous stream described by the dimensionless
Navier-Stokes equations:

( ) ( )
.
,
Re
) 1 ( ) (
Pr
1
Re
1 1
, ) (
3
2
2
Re
1 1
, 0 ) (
* * *
*
2
* * * * * * * * * *
* * * * * * * * *
2
* * * *
* * *
T p
M
T p V T V
V V V p
M
V V
V
M M
T
M


1
]
1

'

+ +


I
(4)
Equations (4) show that the inner flow problem is characterized by the same Prandtl number (Pr) but
with a modified Reynolds number Re
M
= Re R
c
/c which is much smaller than Re. In this way, the local
viscous effects around the airfoil nose are correctly accounted for, specifically when the flow tends to
separate and indicate the onset of stall. Moreover, in the inner region, to the leading orders the airfoil is
described by a canonic smooth parabola
* *
2x y t
. Along the parabola upper and lower surfaces the
no-penetration and no-slip conditions are satisfied. Also, in far field of the inner region viscous effects
decay and the velocity vector is described by a velocity potential,

*
V
, where
) , (log
2
cos 2
/
2
sin 2
1
* *
*
* 0
*
* *

r O r
h Aw
r x + + + . (5)
Here,

,
_

+
*
*
* 2 * 2 2 * *
arctan ,
x
y
y x r


. For more details of this expansion see Rusak.
21
Equation
(5) shows that the far-field flow is near uniform with Mach number M (of the flow around the airfoil)
,
temperature

T , pressure

p , and with higher-order correction terms that relate to a symmetric flow due
to the symmetric nose curvature and an asymmetric circulatory flow with the circulation parameter
h Aw A /
0
~

. Equations (4) and (5) show that in the inner region, the flow is governed by the modified
Reynolds number Re
M
, the Mach number M, and the circulation parameter
~
A
.
In the present work, we look for the specific values of
~
A
as function of Re
M
and M,
( ) M A
M
s , Re
~
, at
which the minimum of pressure coefficient (-Cp
min
) along the parabola suddenly drops with the increase
of
~
A
,
indicating the onset of stall on the airfoil. In this sense the present results are universal. From such
values of
( ) M A
M
s , Re
~
we can predict the stall of a given airfoil at given conditions of flow, Re and M.
Note that the circulation parameter can be given by
c R
w
h
w
A
c
l
2 /
2 / 1
_
0
0
_
~


,
_

,
_

and then the


predicted angle of stall and maximum lift coefficient are given by:
0
_ ~
_
1
0
~
2 2 2
w
c
R
A
w
c
R
A
c
s
l
c
s
s
+ + +
(6a)
and
( ) 1
_ ~
0 max
2
2
2
w
c
R
A C
c
s
l s l

+ . (6b)
Here we used the relationship c R h
c
2 / . Formulae (6a) and (6b) provide predictions of order and
the expected error is O(
2
). This means that the theory is limited to airfoils with small thickness ratio and
as the thickness increases the error also increases. Our numerical studied shown later shown that for
practical airfoils the theory provides nice predictions when <0.15.
It can be seen that according to the present theory, for given flow conditions and when there is no
camber (
0 1
_
0
_
w w
), the stall angle and the maximum lift coefficient increase linearly with the
thickness ratio and/or the square root of the nose radius of curvature. For example, it is predicted that
doubling the thickness ratio and/or increasing the nose radius of curvature by four times results in
doubling the stall angle and the maximum lift coefficient. Also, increasing the camber also increases the
stall angle and the maximum lift since typically both
0 0
_
> w
and
0 1
_
> w
. Deflecting a trailing-edge
flap decreases the stall angle, but since typically
0 1
_
> w
, it also decreases the zero-lift angle much more
than the stall angle and there is a net gain of
1
_
w
in the maximum lift coefficient. For a split trailing-
edge flap it can be shown that
f
h
w


,
_

1 0
_
and
f
h
w

sin 2
1
_
where
c
x
h
h
2 1 cos ,
h
x
is
the axial distance of the flap leading edge form the airfoils leading edge, and
f

is the flap deflection


angle. It is predicted that for given flow conditions and the airfoils nose radius of curvature,
s


decreases linearly and
max l
C
grows linearly with the increase of
f

, i.e.
f h l l l f
h
s s s
f f
C C C



) sin 2 ( , 1
0 max, max max 0 ,

,
_



. (6c)
In addition, it is interesting to observe that deflecting a leading-edge flap increases the stall angle. All of
these qualitative predictions of stall characteristics are similar to those found from the experimental data
for many thin airfoils described in Abbot and von Doenhoff.
8
It can also be seen that increasing the
subsonic flow Mach number causes the decrease of the angle of stall and reduces only the contributions
of camber or a trailing-edge flap to
max l
C
.
II.3. Matching. The inner viscous flow model problem is solved numerically using the
commercial code Fluent. The solution results in velocity field around the parabola and the distribution of
the pressure coefficient Cp
*
(x
*
;
~
A
, M, Re
M
) along the parabola surfaces. Both are strongly affected by
Re
M
, specifically when the Reynolds number is relatively low. The solutions are used to determine the
angle of onset of separation and the value of
( ) M A
M
s , Re
~
. It should also be noted that for most relevant
applications Re
M
is sufficiently large (Re
M
>>1/
4
) such that very thin boundary layers form along the
parabola surfaces, the thickness of which is much smaller than the parabola radius of curvature. Then, the
flow around the parabola and outside the boundary layers is approximately inviscid and potential.
Therefore, following Rusak,
21
the pressure coefficient Cp
*
(x
*
;
~
A
, M, Re
M
) along the parabola surfaces
decays like
h
x Aw
m

*
0
/ 2
) 1 ( as x
*
increases.
For a complete analysis, an asymptotic matching between the inner and outer solutions for the
pressure distribution in an overlap region between the two regions (R
c
<
2 2 2
y x + < c) is also
established (see details in Rusak
21
). The expansions have a common part of the pressure distribution
along the airfoil surface, which is approximated by:

x c w
A M c x Cp
m
cp
0
2
) 1 ( ) , , ; / ( (7)
where m=1 for the upper surface and m=-1 for the lower surface. Note that the common part given by
Eq. (5) coincides with the singular behavior of
l , u
Cp
as x approaches zero and with the decay of Cp
*
as
x
*
increases. The asymptotic matching results also in a uniformly valid asymptotic composite solution of
the pressure distribution along the entire airfoil chord that can be approximated by:

[ ] ). ( ) ( ) , , ; / ( ) / (
) Re , , ; ( Re) , , , ; / (
* * * *
,
~
* *
x x V A M c x Cp c x Cp
M A x Cp M c x Cp
cp l u
M


+

(8)
Here,

1
1
]
1

,
_

1
2
1
) 1 (
2
1
/ ) 1 (
*
2
2
*

Cp
M
M
V and
) / 1 (
*
2
*
Cp
2
Ma
1

,
_

+

. (9)
They represent the non-dimensional velocity and density at the edge of the boundary layers that form in
the inner region along the parabola surfaces and are functions of Cp
*
only. The formula (8) shows that
the pressure distribution along the airfoil is composed of a non-linear combination of the linear theory
solution and the nose solution. As the leading edge of the airfoil is approached (x tends to zero and 0 <
x/c < R
c
/c) the common part pressure Cp
cp
cancels the nose singularity of the outer linear pressure
coefficient solution both on the upper and lower surfaces of the airfoil. Also, in this region, V
*
is small
and tends to zero near the stagnation point. Therefore, the dominant term in the leading-edge region is
the parabola pressure coefficient Cp
*
which is strongly affected by the surface temperature distribution
around the nose and the modified Reynolds number. As x is increased beyond the leading-edge region,
< x/c < 1, the parabola density
*
and velocity V
*
tend to one, and the common part pressure Cp
cp
cancels
the parabola pressure Cp
*
. Therefore, the dominant term in the outer region is the classical subsonic
linear theory solution Cp
u,l
. In the intermediate region, R
c
/c < x/c < , the pressure changes uniformly
from Cp
*
to Cp
u,l
.
III. Numerical solution of the model parabola problem
The computed pressure coefficient distributions along the parabola surfaces for various values of
~
A
are
shown in Fig. 1 for a representative case where Re
M
= 362 and M=0.2. It can be seen that as
~
A
increases
the stagnation point (where Cp is maximal) shifts downstream along the lower surface and the point of
maximum suction (where Cp is minimal) shifts upstream and the suction deepens up to the certain state
where separation bubbles develop. In this case, the onset of stall occurs at
~
A
s
~ 1.76 where (-Cp
min
)
suddenly drops as
~
A
is further increased. Figures 2-4 demonstrate the contour lines of the axial velocity
fields around the canonic parabola before, at, and after the onset of massive separation. In Fig. 2, where
~
A
= 1.57, the boundary layers are very thin and a very short and small closed separation bubble has
already developed as predicted by Stewartson et al.
18
Figure 3 shows the growth of the short closed
separation bubble and a swelling of the boundary layer around 2.5R
c
which indicates the onset of massive
flow separation on the parabola at the certain value of
~
A
=1.76. In Fig. 4, where
~
A
=1.96, the boundary
layer is fully detached of the upper surface of the parabola. It can be seen that the bursting of the bubble
develops from the growth of the short bubble as the circulation parameter is increased from 1.76 to 1.96
for Re
M
=362. Note however that the flow state at
~
A
=1.96 is expected to be unstable and unsteady.
The computations were repeated for various modified Reynolds numbers at M=0.2. Figure 5
shows the resulting
~
A
s
as function of Re
M
for M=0.2. This line represents the universal values of the
scaled circulation parameter at stall for low-speed nearly-incompressible flows at various Re
M
.

It can be
seen that
~
A
s
increases with Re
M
. Note that the present theoretical approach is limited to Re
M
>>1/
4
.
Figure 1: The distribution of the pressure coefficient along the parabola upper and lower surfaces at
various values of
~
A
. Here Re
M
=362 and M=0.2.
Figure 2: Contours of the axial velocity field around the parabola at circulation parameter
~
A
=1.57. Here
Re
M
=362 and M=0.2.
Re
M
=362, M=0.2
-4
-3
-2
-1
0
1
2
0 1 2 3 4 5 6 7 8 9
10
x/R
c
C
p
=1.37 =1.57 =1.76 =1.86 =1.96 =2.15
Figure 3: Contours of the axial velocity field around the parabola at circulation parameter
~
A
=1.76. Here
Re
M
=362 and M=0.2.
Figure 4: Contours of the axial velocity field around the parabola at circulation parameter
~
A
=1.96. Here
Re
M
=362 and M=0.2.
Figure 5: The circulation parameter at stall,
~
A
s
, as function of Re
M
for M=0.2.
IV. Computed Results
The results of Fig. 5 can now be used to compute the stall onset angle
s
of various airfoils with a
parabolic nose at various flow conditions. For example, Fig. 6 presents the computed results for the angle
of onset of stall on the NACA0012 airfoil that are predicted from the present theoretical approach based
on the parabola solution and that are found from full direct numerical simulations of the flow around the
airfoil using Fluent. In all cases M=0.2. A good agreement, within up to of angle of attack, is found
between the present theoretical predictions and the detailed computations for the full range of Reynolds
numbers from 18,000 to 9 Million. A similar agreement between the theoretical prediction and numerical
simulations is expected for any other airfoil.
Using equation (6a), the values of the scaled circulation parameter at stall,
s A
~
, as function of the
modified Reynolds number, Re
M
, are computed for various symmetric and cambered airfoils. These
calculations are based on the experimental data of the stall angle of attack
s

at various Re as reported by
Jacobs and Sherman
7
and Selig et al.
9
and the NACA airfoils geometrical properties as defined in Abbot
and von Doenhoff.
8
Note that the range for the experimental error in all of the cases, about 5 . 0 t degree,
is not shown. The results are presented in Fig. 7 and show that all the experimental values concentrate
around the line predicted by the present theory. This demonstrates that the theoretical line prediction is
within 5 . 0 t of the experimental value which is typically equivalent to 5 . 1 t degree for the prediction of
the angle of attack of stall. It can also be seen that the theoretical line over predicts the angle of stall as
the airfoil thickness ratio increases above 15%. These results demonstrate the strong dependence of stall
on the modified Reynolds number Re
M
.
M=0.2
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
0.0E+00 1.0E+04 2.0E+04 3.0E+04 4.0E+04 5.0E+04
Re
M

s
M=0.2
0
2
4
6
8
10
12
14
16
18
0.E+00 1.E+06 2.E+06 3.E+06 4.E+06 5.E+06 6.E+06 7.E+06 8.E+06 9.E+06 1.E+07
Re

s
NACA 0012
Parabola
Figure 6: A comparison of the predicted results for the angle of onset of stall according to the present
theoretical approach and to direct numerical simulations. Here M=0.2.
M=0.2
0.0
1.0
2.0
3.0
4.0
5.0
6.0
2 2.5 3 3.5 4 4.5 5 5.5
Log(Re
M
)

s
theory
0009
0012
2412
4412
6412
4409
23012
0015
4415
6712
2414
2415
Figure 7: Comparison of experimental data with theoretical prediction for
s A
~

as function of log(Re
M
).
Figure 8a and 8b show the stall angle of attack and maximum lift coefficient as function of the
airfoils radius of curvature for 25 symmetric NACA airfoils at Re=3 Million and M=0.2 as reported by
Abbot and von Doenhoff.
8
Each data point is shown with an error range corresponding to 5 . 0 t degree.
Also shown are the theoretical predictions as computed according to equations (6a) and (6b). It can be
seen again that the experimental data for the stall angle concentrates around the theoretical line. It can
also be seen that the predicted results for C
l max
are accurate when the airfoils thickness ratio is below
15% and over estimate the experimental data for C
l max
when the thickness ratio is above 15%. These
figures, within limitations, demonstrate the strong dependence of stall on the nose radius of curvature.
Re = 3x10
6
0.00
5.00
10.00
15.00
20.00
25.00
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035
R
c
/c

s
alpha s theory
0006
0009
0012
63-006
63-009
63,1-012
63,2-015
64-006
64-009
64,1-012
64,2-015
64,3-018
65-006
65-009
65,1-012
65,2-015
65,3-018
65,4-021
66-006
66-009
66,1-012
66,2-015
66,3-018
66,4-021
747A315
Figure 8a: Comparison of experimental data and theoretical predictions for
s
as function of R
c
/c at
Re=3 millions and M=0.2.
Re = 3x10
6
0.00
0.50
1.00
1.50
2.00
2.50
0 0.005 0.01 0.015 0.02 0.025 0.03 0.035
R
c
/c
C
l max
Cl max Theor y
0006
0009
0012
63- 006
63- 009
63, 1- 012
63, 2- 015
64- 006
64- 009
64, 1- 012
64, 2- 015
64, 3- 018
65- 006
65- 009
65, 1- 012
65, 2- 015
65, 3- 018
65, 4- 021
66- 006
66- 009
66, 1- 012
66, 2- 015
66, 3- 018
66, 4- 021
747A315
t = 21% t =1 8%
t =15%
Figure 8b: Comparison of experimental data and theoretical predictions for C
l max
as function of R
c
/c at
Re=3 millions and M=0.2.
Table 1 summarizes the results of the stall angle of attack as the camber of the NACA airfoils is
varied at a fixed thickness ratio. The theory predicts according to (6a) the increase of the stall angle with
camber as is also suggested by the experiments, compare for example the results for symmetric airfoil
NACA 0012 with those for the cambered airfoils NACA 1412, NACA 2412 and NACA 23012. Note
however that for most of the families of airfoils shown below the increment is relatively small and below
0.5 degree, which is within the experimental error for the stall angle. The difference between the
experimental data and theoretical predictions is mostly within 0 . 1 t degree.

NACA

s

(theory)

s
(experiment)
Diff
0012 15.2 15.5 0.3
1412 15.3 15.7 0.4
2412 15.4 15.7 0.3
4412 15.7 15.0 -0.7
23012 16.8 16.0 -0.8
0010 12.7 13.5 0.8
1410 12.8 14.0 1.2
2410 12.9 14.2 1.3
0008 10.1 10.5 0.4
1408 10.2 11.5 1.3
2408 10.4 11.6 1.2
Table 1: Comparison of experimental data and theoretical predictions for
s
as function of camber at
Re=3 millions and M=0.2.
The effect of deflecting a split trailing-edge flap of the NACA 230XX series of airfoils is studied
in Figs. 9a,b,c. The increase of C
l max
as function of
f

is computed using equation (6a) and described in


Figs. 9a and b for various flap sizes. Also shown are the experimental results of Abbott and von
Doenhoff (Ref. 8, pp 202). It can be seen that the theory nicely predicts the increase of the maximal lift
for various flap sizes and flap deflections up to 20 degrees for the NACA 230012 (see Fig. 9a). For larger
flap deflections the theory over predicts this effect since at these conditions the flap may also suffer flow
separations. Note that for thicker airfoils, NACA 23021 and NACA 23030, the theory can even predict
the experimental data up to 40 degrees of flap deflection (see Fig. 9b). The slope dC
l max
/d
f

at small flap
deflections as function of flap chord C
f
is also computed from (6c) and shown in Fig. 9c. It can be seen
that the theoretical predictions nicely agree with the experimental results for flap sizes up to 30% of the
airfoils chord. These results demonstrate the accuracy of the theory in predicting the effect of flap size
and deflection on the stall of an airfoil.
Finally, the effect of flow compressibility at subsonic speeds on the angle of stall is studied in
Fig. 10 for a NACA 66
(109)
-210 airfoil. A nice agreement is found between the theoretical predictions
based on equation (6a) and the experimental data according to Graham.
23
Note that the experimental
point at M=0.2 is taken from Abbott and von Doenhoff (Ref. 8, pp 660) for the NACA 66-210 (which is
very similar to NACA 66
(109)
-210) . Both demonstrate the decrease of
s

with the increase of the flow


Mach number (or decrease of ). Moreover, the experimental data shows that C
l max
is nearly independent
of Mach number, as predicted by equation (6b). It should be noted however that, to the best of our
knowledge, the experimental data of stall at high subsonic speeds is quite limited and additional
experiments are needed to better clarify the influence of compressibility on stall.
NACA 23012
Change in Max Lift Vs Flap Deflection at M=0.2
0.00
0.50
1.00
1.50
2.00
2.50
0 10 20 30 40 50 60 70 80 90
Flap Deflection
f
(Degrees)
C
lmax
.10c .20c .30c .40c
.10c EXP .20c EXP .30c EXP .40c EXP
Figure 9a: Comparison of experimental data and theoretical prediction for the increase in C
l max
as
function of flap deflection for various flap sizes of a NACA 23030 at Re=3 Millions.
NACA 23030
Change in Max Lift Vs Flap Deflection at M=0.2
0.00
0.50
1.00
1.50
2.00
2.50
0 10 20 30 40 50 60 70 80 90
Flap Deflection
f
(Degrees)
C
lmax
.10c .20c .30c .40c
.10c EXP .20c EXP .30c EXP .40c EXP
Figure 9b: Comparison of experimental data and theoretical prediction for the increase in C
l max
as
function of flap deflection for various flap sizes of a NACA 23030 at Re=3 Millions.
0
0.5
1
1.5
2
2.5
0.00 0.10 0.20 0.30 0.40 0.50 0.60
C
f
dC
lmax
df
experimental
theoretical
Figure 9c: Comparison of experimental data and theoretical prediction for the slope dC
l max
/d
f
as
function of flap size at Re=3 Millions.

0.0
2.0
4.0
6.0
8.0
10.0
12.0
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90
Mach Number

s
Theory Line
66(109)-210
Figure 10: Comparison of experimental data and theoretical prediction for
s
as function of Mach
number for the NACA 66
(109)
-210 airfoil.
V. Conclusions
The leading-edge stall of two-dimensional smooth airfoils at moderate Reynolds number flows
can be studied by an asymptotic approach. The asymptotic theory is based on the work of Rusak
21
and
results in a model (simplified) problem of a uniform flow past a semi-infinite parabola with a far-field
circulatory flow that is governed by a parameter
~
A
which depends on the modified Reynolds number,
Mach number, nose radius of curvature and airfoils camber. The model parabola problem consists of a
viscous flow that is solved numerically for various values of the modified Reynolds number, Mach
number to determine the special value
~
A
s
where the nose flow first separates. Figure 5 presents the
criterion for the onset of leading-edge stall on thin airfoils at Mach number 0.2 and various Re
M
. Similar
figures can be constructed for other subsonic flow Mach numbers. The parameter
~
A
s
can be used to
determine the stall onset on thin airfoils. The predictions according to this approach show good
agreement with results from full numerical simulations of the stall on thin airfoils and with much
available experimental data. This simplified approach provides a universal criterion to determine the
stall angle of airfoils with a parabolic nose at moderate Reynolds number flows. It also explains that
increasing the airfoils thickness and/or the nose radius of curvature increase the stall angle and the
maximum lift coefficient. Increasing camber also increases the stall angle and maximum lift. Deflecting
a split trailing-edge flap decreases the stall angle but a net gain in maximum lift coefficient occurs due to
the more significant change in the zero lift angle of attack. Additionally, the theory predicts that
deflecting a leading-edge flap increases the stall angle and maximum lift coefficient. Increasing the flow
Mach number causes the stall angle to decrease and reduces the change of
max l
C
due to camber or a
trailing-edge flap.
References
1. Jones, B.M. 1934 Stalling, J. Royal Aero Society, 38, 753-770.
2. Tani, I. 1964 Low speed flows involving bubble separations, Prog. Aeronautical Sciences, 5,
70-103.
3. McCormick, B.W. 1995 Aerodynamics, Aeronautics and Flight Mechanics, 2
nd
edition, John
Wiley and Sons Inc, New York.
4. Anderson, J.D. Jr 2001 Fudamentals of Aerodynamics, 3
rd
edition, McGraw Hill, New York.
5. Liebeck, R.H. 1973 A class of airfoils designed for high lift in incompressible flow, AIAA
Journal of Aircraft, 10 (10), October.
6. Liebeck, R.H. and Ormsbee, A.I. 1970 Optimization of airfoils for maximum lift, AIAA
Journal of Aircraft, 7 (5), September-October.
7. Jacobs, E.N. and Sherman, A. 1937 Airfoil section characteristics as affected by variations
of the Reynolds number, NACA TR 586.
8. Abbott, I.H. and von Doenhoff, A.E. 1958 Theory of Wing Sections, 2
nd
edition, Dover
Publications Inc., New York.
9. Selig M.S., Lyon, C.A., Giguere, P., Ninham, C.P. and Guglielmo, J.J. 1996 Summary of
Low-Speed Airfoil Data, Volume 2, Department of Aeronautical and Astronautical
Engineering, University of Illinois at Urbana-Champaign, Urbana, Illinois.
10. Nakayama, Y. 1988 Visualized Flow, Pergamon Press, New York.
11. Hung, K., Papadakis, M. Wong, S.-H, and Wong, S.-C. 1998 Computation of pre- and post-
stall flows over single and multi-element airfoils, AIAA Paper 98-0756, 36
th
AIAA
Aerospace Sciences Meeting, Reno, NV, January.
12. Moulton, M. and Steinhoff, J. 2000 A technique for the simulation of stall with coarse-grid
CFD methods, AIAA Paper2000-0277, 38
th
AIAA Aerospace Science Meeting, Reno, NV,
January.
13. Grove, D.V., Laiosa, J.P., Woodson S.H. and Stookesberry, D.C. 2002 Computational fluid
dynamics study of an abrupt wing stall phenomena on the F/A-18E, AIAA Paper 2002-1025,
40
th
AIAA Aerospace Sciences Meeting, Reno, NV, January.
14. OMeara, M. M., and Mueller, T. J. 1987 Laminar separation bubble characteristics on an
airfoil at low Reynolds numbers, AIAA Journal, 25, No. 8, pp. 1033-1041, August.
15. Kunz, P.J. and Kroo, I. 2001 Analysis and design of airfoils for use at Ultra-low Reynolds
numbers, in: Fixed and Flapping Wing aerodynamics for micro air vehicle applications,
Editor: Muller T.J., Progress in Astronautics and Aeronautics, 195, AIAA Publication,
Reston, VA.
16. Werle, M.J. and Davis, R.T. 1972 Incompressible laminar boundary layrs on a parabola at
angle of attack: A study of the separation point, ASME J. Applied Mechanics, 39 (1), 7-12.
17. Cebeci, T., Khattab, A.K. and Stewartson, K. 1980 On nose separation, J. Fluid Mechanics,
97, 435-454.
18. Stewartson, K. Smith, F.T. and Kaups, K. 1982 Marginal separation, Studies in Applied
Mathematics, 67 (1), 45-61.
19. Ruban, A.I. 1982 Asymptotic theory of short separation bubbles at the leading edge of a thin
airfoil, Izv. Akad. Nauk SSSR, Mech. Zhid Gaza, (1), 42-52.
20. Cheng, H.K. and Smith, F.T. 1982 The influence of airfoil thickness and Reynolds number
on separation, J. of Applied Mathematics and Physics (ZAMP), 33, March, 151-180.
21. Rusak, Z. 1994 Subsonic flow around the leading edge of a thin aerofoil with a parabolic
nose, European Journal of Applied Mathematics, 5, 283-311.
22. Van Dyke, M.D. 1956 Second order subsonic airfoil theory including edge effects, NACA
Rep. 1274.
23. Graham, D.J. 1947 High-speed tests of an airfoil section cambered to have critical Mach
numbers higher than those attainable with a uniform-load mean line, NACA TN 1396.

Potrebbero piacerti anche