You are on page 1of 39

Lifting Airfoils in Incompressible Irrotational Flow

AA200
Lecture 2
January 18, 2012
AA200 - Applied Aerodynamics 1
AA200 - Applied Aerodynamics Lecture 2
Governing Equations
For an incompressible uid, the continuity equation takes the form
u
x
+
v
y
= 0. (1)
If the ow is irrotational ( u 0), the velocity eld can be written
as the gradient of a scalar function called the potential function (typically
represented by the symbol ). This is expressed in the following vector
identity
u = (2)
When substituting the denition of the potential, , into Eq. 1, we obtain
the governing equation for , Laplaces equation:

2
=

2

x
2
+

2

y
2
= 0 (3)
2
AA200 - Applied Aerodynamics Lecture 2
which the potential function must satisfy everywhere, including the
appropriate boundary conditions. In this way we have reduced the problem
of nding the velocity eld (2 components, (u, v)) to a single equation. In
addition, one can show that the momentum equations can be combined
with the statement of irrotationality to obtain a relation between velocity
magnitude and pressure called the Bernoulli equation
p +
1
2
(u
2
+v
2
) = p
0
. (4)
One way to obtain solutions to Laplaces equation (subject to the appropriate
boundary conditions) is to exploit its linear nature and the principle of
superposition. Because Laplaces equation is linear, if we have two solutions

1
and
2
, the linear combination
1
+
2
is also a solution, with and
arbitrary scalars. Obviously, the principle of superposition can be used to
add up an arbitrary number of elementary solutions to Laplaces equation.
3
AA200 - Applied Aerodynamics Lecture 2
This method is rather powerful and will be used in the remainder of
this lecture to construct the solution of the ow about an arbitrarily shaped
lifting airfoil.
In this lecture we will use three types of elementary solutions: free
stream, source/sink, vortex. With a large number of each of these three
kinds of solutions we will be able to construct the ow about rather general
airfoils at arbitrary angles of attack.
4
AA200 - Applied Aerodynamics Lecture 2
Free Stream Potential
The potential function for a free stream of magnitude V

, aligned with
the xaxis is given by

= V

x. (5)
Taking the gradient of this potential we see that the resulting velocity eld
is given by
u(x, y) = V

(6)
v(x, y) = 0.
That is, the velocity is uniform everywhere in the domain. The potential
function can be rotated at an arbitrary angle so that

= V

(cos x +
siny), which for small angles of attack can be approximated as

(x +y).
5
AA200 - Applied Aerodynamics Lecture 2
Source/Sink Potential
A source/sink that expels/absorbs an amount of uid volume/unit time
= Q can be constructed from the following potential

S
=
Q
2
lnr, (7)
where r =
_
x
2
+y
2
. The resulting velocity components are
u(x, y) =
Q
2
x
x
2
+y
2
(8)
v(x, y) =
Q
2
y
x
2
+y
2
.
The streamlines in these ows can be shown to be straight lines emanating
radially outwards from the location of the source/sink.
6
AA200 - Applied Aerodynamics Lecture 2
Point Vortex Potential
The potential of a point vortex with circulation can be shown to be

V
=

2
, (9)
where is dened positive if the induced circular ow is clockwise. is the
angle measured in polar coordinates from some arbitrary origin radial line,
typically = tan
1
_
y
x
_
. Taking the gradient of this function, we see that
the velocity eld of a vortex is given by
u(x, y) =

2
y
x
2
+y
2
(10)
v(x, y) =

2
x
x
2
+y
2
,
7
AA200 - Applied Aerodynamics Lecture 2
and it creates streamlines that are concentric circles centered about the
location of the point vortex.
Notice that the circulation around any countour that encloses the point
vortex is constant and equal to . Furthermore, the ow outside of the
point vortex is fully irrotational. All of the vorticity in this ow is contained
at the singular location of the point vortex. More on this later.
Notice also that the form of the potential for a point vortex is rather
similar to that of a source/sink, with the substitutions Q , x y, and
y x in the appropriate formulae for the velocity eld: the velocity elds
are perpendicular to each other (and so are the equipotential lines).
8
AA200 - Applied Aerodynamics Lecture 2
Thickness and Camber Problems
Consider uniform ow past an airfoil at an angle of attack .
y
x
V

Y
T
The airfoil surface is dened by its upper and lower surfaces located at
y = Y
u
(x) and y = Y
l
(x).
The line that joins the leading and trailing edges is aligned with the
x-axis and spans the interval [0, c].
9
AA200 - Applied Aerodynamics Lecture 2
The thickness and camber distributions are given, by denition, by
T(x) = Y
u
(x) Y
l
(x) (11)

Y (x) =
1
2
[Y
u
(x) +Y
l
(x)],
which simply yield
Y
u
(x) =

Y +
1
2
T (12)
Y
l
(x) =

Y
1
2
T.
In thin airfoil theory, we assume that u V

, and furthermore, we apply


the airfoil surface boundary conditions along the x-axis instead of at the
exact surface location.
10
AA200 - Applied Aerodynamics Lecture 2
A fundamental necessity to create lifting ows is that our solution must
admit discontinuities across the x-axis. For that purpose, we consider the
values of all the relevant variables at 0

, depending on whether we are


looking at the upper or lower surface of the airfoil.
With these assumptions, the ow tangency boundary conditions at a
solid wall reduce to:
v(x, 0
+
) V

_
d

Y
dx
+
1
2
dT
dx
_
(13)
v(x, 0

) V

_
d

Y
dx

1
2
dT
dx
_
,
which can be satised in a variety of ways. A physically meaningful approach
is to construct a potential function for the ow around the complete airfoil
11
AA200 - Applied Aerodynamics Lecture 2
with 3 main components:
=

+
T
+
C
, (14)
where

is the potential of a free stream inclined at an angle of attack


to the x-axis, and
T
and
C
are the potentials created by line distributions
of sources/sinks and vortices placed along the x-axis. They will be later
related to the thickness and camber portions of the problem.
Note that

alone satises the boundary conditions at the far eld.


Therefore,
T
and
C
must vanish at the far eld so as not to violate
the boundary condition there.
T
is chosen to satisfy the portion of the
ow tangency boundary condition associated with the thickness distribution,
while
C
will be chosen so that, together with

the camber portion of


the boundary condition is satised.
12
AA200 - Applied Aerodynamics Lecture 2
In other words, for small ,

(x +y), (15)
v
T
=

T
y
=
1
2
V

dT
dx
at y = 0

for 0 < x < c (16)


v
C
=

C
y
= V

_
d

Y
dx

_
at y = 0

for 0 < x < c. (17)


It is possible to show that, the thickness problem can be solved with a line
distribution of sources along the x-axis, while the camber problem can be
solved with a distribution of vortices along the same axis. Both distributions
will be assumed to be continuously varying with strength q(x) and (x)
respectively.
13
AA200 - Applied Aerodynamics Lecture 2
Thickness problem
The velocity eld created by a distribution of sources is given by
v
s
=
_
c
0
q(x

)
2
y
(x x

)
2
+y
2
dx

(18)
u
s
=
_
c
0
q(x

)
2
x x

(x x

)
2
+y
2
dx

.
Taking the limit of these expressions as y 0 allows us to extend the limits
of integration from to , since the integrand eectively behaves as a
delta function as y 0. With this approximation, we nd that
v
s
(x, 0

) =
1
2
q(x), (19)
14
AA200 - Applied Aerodynamics Lecture 2
that is, half of the volume ux per unit depth out of a dierential element of
source is sent directly upwards, while the other half is directed downwards.
This relationship between source strength and ow speed directly produces
the strength of a source distribution that is necessary to generate the shape
of a body of interest (i.e. satisfy the ow tangency boundary condition). In
particular, for the thickness problem, Equation 16 together with Equation 19
yield
q(x) = V

dT(x)
dx
. (20)
The contribution to the x-component of the velocity vector from this
distribution of sources is given by
u
T
(x, 0

) =
V

2
_
c
0
dT
dx
dx

x x

for 0 < x < c (21)


15
AA200 - Applied Aerodynamics Lecture 2
Note that u
T
is continuous across the x-axis, while v
T
is discontinuous:
u
T
(x, 0
+
) = u
T
(x, 0

) (22)
v
T
(x, 0
+
) = v
T
(x, 0

)
Also note that near the leading edge of an airfoil, the thickness slope
gets large. This violates the assumptions of thin airfoil theory and leads
to diculties with the evaluation of equation 21. In the 1950s Riegels
introduced a correction to thin airfoil theory that greatly improves the
accuracy. In fact, the Riegels correction produces results that work for
ellipses up to and exceeding 100% thick, while the previous theory only
works well up to 10 or 15%. The velocity on the surface of the airfoil with
Riegels modication is:
V
tangent
V

=
1
_
1 +
dy
dx
2
_
1 +
1

_
c
0
dT
dx
dx

x x

_
16
AA200 - Applied Aerodynamics Lecture 2
Camber Problem
Notice from Equation 17 that the vertical velocity component due to
the vortex distribution v
C
is continuous across the surface of the airfoil.
Therefore, we can infer that the horizontal velocity component due to
this vortex distribution must be allowed to be discontinuous across the
surface of the airfoil, since, otherwise, the total horizontal velocity would be
continuous across the surface of the airfoil, which we know not to be true
from a physical standpoint, and which would lead to problems when trying
to obtain non-zero lift from this mathematical setup.
Lets represent the solution to the camber problem via a distribution of
vortices along the x-axis, with strength per unit length (x):

C
=
_
c
0
(x

)
2
tan
1
_
y
x x

_
dx

. (23)
17
AA200 - Applied Aerodynamics Lecture 2
Following a derivation similar to the one used for the source distribution,
we nd the velocity components due to
C
to be:
u
C
(x, 0

) =
1
2
(x) for 0 < x < c (24)
v
C
(x, 0

) =
_
c
0
(x

)
2
dx

x x

for 0 < x < c


In order to satisfy the ow tangency boundary condition in Equation 17,
(x) must satisfy the following integral equation:
1
2
_
c
0
(x

)
dx

x x

= V

_

d

Y
dx
_
for 0 < x < c (25)
This integral equation is dicult to solve, although there are well-known
solutions such as that for a at plate airfoil at an angle of attack, (which
is shown in the following slides).
18
AA200 - Applied Aerodynamics Lecture 2
Assuming the solution (x) which satises the portion of the boundary
conditions for the camber problem can be found, the velocity eld can be
derived from appropriate substitution into Equations 19 and 25, and using
the Bernoulli equation, one can found the pressure distribution on the upper
and lower surfaces of the airfoil, from which forces and moments can be
computed.
19
AA200 - Applied Aerodynamics Lecture 2
Kutta Joukowski Theorem
We can compute the total force that the uid exerts on the airfoil by
simply integrating the pressure around the airfoil surface. The total force
can be decomposed into components along the x and y directions as follows:
F

x
=
_
c
0
_
p
u
dY
u
dx
p
l
dY
l
dx
_
dx (26)
F

y
=
_
c
0
(p
l
p
u
) dx.
From our expressions for u
T
and u
C
on the upper and lower surfaces of the
airfoil we see that
p
u
p
l
V

,
and therefore, the total force per unit depth in the vertical direction is given
20
AA200 - Applied Aerodynamics Lecture 2
by
F

y
V

_
c
0
(x)dx = V

, (27)
where
=
_
c
0
(x)dx
is the total circulation around the airfoil. With a certain amount of algebra,
the x component of the force can be shown to be
F

x
V

_
c
0
(x)dx = V

. (28)
The vector addition of these two components of the total force can be shown
to be perpendicular to the freestream (remember that for small angles of
attack, , tan ). This result is often known as DAlemberts paradox.
21
AA200 - Applied Aerodynamics Lecture 2
Figure 1:
Theorem 1. [Kutta-Joukowski Theorem] An isolated two-dimensional
airfoil in an incompressible inviscid ow feels a force per unit depth of
F

.
Since the lift and drag forces are dened as the forces in the directions
22
AA200 - Applied Aerodynamics Lecture 2
normal and parallel to the free stream, a direct consequence is that
L

= V

= 0,
and thus, the airfoil produces lift while creating no drag. This result is
clearly unphysical since it would amount to a perpetual motion machine,
which would violate the second law of thermodynamics. The result for the
lift force is very close to experiment (for high Reynolds numbers), while the
result for the drag force is clearly o and will be taken care of during our
lectures on viscous ow and boundary layer theory.
23
AA200 - Applied Aerodynamics Lecture 2
Symmetric Airfoil at an Angle of Attack
In order to show an example of the solution to the integral equation 25
lets assume we have a symmetric airfoil (Y
u
(x) = Y
l
(x)), such that the
camberline is a at plate aligned with the x-axis (

Y (x) = 0). The integral


equation for (x) then reduces to
_
c
0
(x

)dx

(x x

)
= 2V

, for 0 < x < c. (29)


While the integrand of this expression depends on x, the right hand side is
a constant. Because of this fact, the integral equation can be solved using
the following coordinate transformation
x =
c
2
(1 cos )
24
AA200 - Applied Aerodynamics Lecture 2
x

=
c
2
(1 cos

)
(x

) = g(

).
The integral equation is transformed into
_

0
g(

) sin

cos

cos
= 2V

.
A useful result in applied aerodynamics (which will come back when we
discuss nite wing theory) is that (look in your CRC):
_

0
cos n

cos

cos
d

=
sinn
sin
. (30)
Since the right hand side of this integral reduces to when n = 1, a
25
AA200 - Applied Aerodynamics Lecture 2
solution to the integral equation is certainly
g() sin = 2V

cos .
However, for n = 0, the integral vanishes, so that we may add an arbitrary
constant to the term sin g() while still retaining a valid solution to the
integral equation 29!!! Hmmmm . . . the implications of this constant are
quite relevant. In general,
g() =
k
sin
+ 2V

cos
sin
,
where k is the arbitrary constant. It can be shown that the total circulation
around the airfoil is given by
=
_

0
g()
c
2
sind =
kc
2
.
26
AA200 - Applied Aerodynamics Lecture 2
Since the constant k can change the value of the total circulation, , it
will change the total value of the force exerted by the uid on the airfoil
. . . you may have already guessed that something is wrong here.
Notice that the reverse transformation is given by
cos = 1
2x
c
sin = 2
_
x
c
_
1
x
c
_
,
and therefore, the vortex strength distribution is given by
(x) =
k
2
_
x
c
_
1
x
c
_
+ 2V

1
2x
c
_
x
c
_
1
x
c
_
.
27
AA200 - Applied Aerodynamics Lecture 2
If k is set to 2V

, one can show by applying LHospitals rule that the


circulation vanishes at the trailing edge point x = c, leaving the following
expression for the vortex strength distribution
(x) = 2V

_
1 (x/c)
_
x/c
.
This choice of k is the one that satises the Kutta condition which we will
discuss next.
The gure below shows the resulting circulation distribution on a
symmetric airfoil at a 3

angle of attack. Note that there is still a


singularity at the leading edge of the airfoil, but all ambiguity has been
removed by xing the value of the constant k.
28
AA200 - Applied Aerodynamics Lecture 2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
Circulation Distribution on a Symmetric Airfoil
X/C
G
a
m
m
a
Finally, the section lift and moment can be found to be
L

= V
2

c (31)
M

l.e.
=

4
V
2

c
2
, (32)
so that the center of pressure (point on the airfoil at which the lift vector
29
AA200 - Applied Aerodynamics Lecture 2
appears to be acting) is located at the quarter chord point
x
c.p.
=
M

l.e.
L

=
c
4
.
Furthermore, one can see that since the lift vector appears to be acting
at the quarter chord location (independently of angle of attack for this
symmetric airfoil...is that true for non-symmetric airfoils?) we can compute
the moment about any point (other than the leading edge) for this airfoil.
If we do this, we would nd that for
x
c
=
1
4
the moment is independent of
the angle of attack of the airfoil. This point is rather useful and is called
the aerodynamic center. As we will see later, thin airfoil theory predicts
this location of the aerodynamic center for airfoils with arbitrary camber as
well.
30
AA200 - Applied Aerodynamics Lecture 2
Kutta Condition
As we saw in the previous example, the circulation distribution on
an arbitrary airfoil is xed by both the far eld and solid wall boundary
conditions up to a constant. In order to obtain a unique solution to this
problem, we must nd some additional information that helps us determine
the proper value of the constant, k.
Observation of the ow near the trailing edge shows that we must have
one of the following alternatives:
1. The velocity at the trailing edge is innite.
2. The ow leaves the trailing edge smoothly along the direction of the
bisector of the trailing edge angle.
31
AA200 - Applied Aerodynamics Lecture 2
In a real ow, the velocity cannot be innite, since it would lead to very low
pressure (high adverse pressure gradient) and the ow would separate. The
second option is the one that occurs in reality.
Theorem 2. [Kutta Condition] The ow leaves the trailing edge of a
sharp-tailed airfoil smoothly; that is, the velocity is nite there.
The following corollaries for the Kutta-condition can also be derived:
The streamline that leaves a sharp trailing edge is an extension of the
bisector of the trailing edge angle.
Near the trailing edge, the ow speeds on the upper and lower surfaces
of the airfoil are equal at equal distances from the trailing edge.
Unless the trailing edge is cusped, the ow has a stagnation point at the
trailing edge.
32
AA200 - Applied Aerodynamics Lecture 2
The most useful of these corollaries is the second one. This leads to the
conclusion that there cannot be a jump in u
C
velocity at the trailing edge.
Since, according to Equation 25 the jump in u
C
is equal to (x), we must
have (c) = 0. This is the condition that we imposed in the previous
example leading to the elimination of the constant k.
33
AA200 - Applied Aerodynamics Lecture 2
Cambered Thin Airfoil
Lets now assume that the camberline of the airfoil is no longer at,

Y = 0. As in the previous example, we need to nd the circulation


distribution (x) that satises Equation 25, and that satises the Kutta
condition (c) = 0.
Using the same coordinate transformation as before we have
1
2V

_

0
g(

) sin

cos

cos
= s(), for 0 < < , (33)
where
d

Y
dx
= s(),
and the Kutta condition is given by
g() = 0.
34
AA200 - Applied Aerodynamics Lecture 2
One option to solve this equation is to expand g() sin into a Fourier
cosine series, but we would run into trouble evaluating the Kutta condition.
Alternatively, we can start directly with a form of the solution that always
satises the Kutta condition. One such form is given by
g() sin = 2V

_
A
0
(1 + cos ) +

n=1
A
n
sinn sin
_
. (34)
Using Equation 30, some trigonometric identities, and the fact that it can
be shown that one can expand
s() = A
0

n=1
A
n
cos n
35
AA200 - Applied Aerodynamics Lecture 2
it can be shown that the solution to the thin cambered airfoil is given by
A
0
=
1

_

0
s()d
and
A
m
=
2

_

0
s() cos m d for m = 1, 2, . . .
Note that in the Fourier series above we are expanding the camber slope
distribution

Y

(x) =
d

Y
dx
into a series that converges even if the function
being expanded contains jump discontinuities (i.e. the camberline slope is
not continuous, which would occur for airfoils with aps and slats).
Formulae for the lift and moment coecient, as well as the location of
the center of pressure can be worked out in a manner similar to the one
36
AA200 - Applied Aerodynamics Lecture 2
followed for the symmetric airfoil at an angle of attack. In summary
L

= V
2

c(A
0
+
1
2
A
1
)
M

l.e.
=

4
V
2

c
2
(A
0
+A
1

1
2
A
2
)
x
c.p.
=
c
4
_
A
0
+A
1

1
2
A
2
A
0
+
1
2
A
1
_
One can see that for this airfoil, the location of the center of pressure
changes with angle of attack (since the A
0
coecient will change with ).
However, one can show there is a point on the airfoil about which the
aerodynamic moment does not change with angle of attack. This point
is called the aerodynamic center, and, according to thin airfoil theory it is
37
AA200 - Applied Aerodynamics Lecture 2
located at the quarter chord, that is
x
a.c
=
c
4
Finally, the coecients of lift and moment can be shown to be
c
l
= 2(A
0
+
1
2
A
1
) = 2
_

1

_

0
s()(1 cos )d
_
(35)
c
mac
=

4
(A
1
A
2
) =
1
2
_

0
s()(cos cos 2)d
This airfoil theory then predicts that
c
l
= m
0
(
L0
)
38
AA200 - Applied Aerodynamics Lecture 2
where m
0
is the lift-curve slope
m
0
= 2
and
L0
is the zero-lift angle of attack

L0
=
1

_

0
s()(1 cos )d
Since the perturbation velocity is: u =
2g
V

, the local pressure dierence


is given by:
C
p
l
C
p
u
=
2g(x)
V

= 4
_
A
0
cot

2
+

1
A
n
sinn
_
39

You might also like