You are on page 1of 10

M&AE 305

October 3, 2006

Thin Airfoil Theory


D. A. Caughey
Sibley School of Mechanical & Aerospace Engineering
Cornell University
Ithaca, New York 14853-7501
These notes provide the background needed to implement a simple vortex-lattice numerical method
to determine the properties of thin airfoils. This material is covered in Lecture, but is not in
the textbook [5]. A summary of results from the analytical theory also is provided, as well as a
comparison of the thin-airfoil results with those of a complete inviscid theory that accounts for
thickness effects.

The Vortex Lattice Method

We here describe the implementation of the vortex lattice method for two-dimensional flows past
thin airfoils. The method is even more useful for three-dimensional wings, i.e., for the flow past
wings of finite span, but that problem is not considered here. Instead, the reader is referred to
standard aerodynamics texts, e.g., [2]. In this numerical procedure to solve the thin-airfoil problem,
we place a finite number of discrete vortices along the chord line, with the boundary condition that
the induced vertical velocity
dyc
v=
,
(1)
dx
be enforced at selected control points to determine the vortex strengths. Equation (1) simply says
that the net velocity vector, comprised of components due to the free stream, at angle of attack
to the chord line, plus that induced by the point vortices, is tangent to the camber line whose slope
is dyc / dx; the magnitude of the free stream velocity is taken to be unity.
Thus, we discretize the chord line into a finite number N of segments, or panels, as illustrated in
Fig. 1 (a). On each panel we place a point vortex and a control point, as illustrated in Fig. 1 (b).
The most accurate results are obtained by locating the vortex one-quarter of the panel length, and
the control point three-quarters of the panel length, aft of the leading edge of the panel. (This
strategy can be shown to reproduce the exact results of analytical thin-airfoil theory for parabolic
camber lines using a single panel , as shown in Section 2.3.1.)
The vertical velocity vi,j induced at the ith control point by the jth point vortex is given by
vi,j =

j
1
2 xvj xci

where xvj is the chordwise coordinate of the jth vortex having strength j , and xci is the chordwise
coordinate of the ith control point. The total vertical velocity at the ith control point induced by

THE VORTEX LATTICE METHOD

xi

x i+1

(a)

(b)

Figure 1: Sketch of discretization of chord line for implementation of vortex lattice calculation.
(a) Chord line subdivided into N panels; (b) Single panel showing location of point vortex and
control point.
all the vortices representing the airfoil camber line is thus
vi =

N
X
j
1
2 xvj xci
j=1

N
X

ai,j j

j=1

where
ai,j =

2 xvj xci

is the influence coefficient representing the effect on the induced vertical velocity at the ith control
point of a vortex of unit strength located on the jth panel.
If we introduce the vector notation
v = [v1

...

vN ]

...

N ] ,

and
= [1

v2

and define the matrix of influence coefficients

a1,1 a1,2
a2,1 a2,2

A=

aN,1 aN,2

a1,N
a2,N


aN,N

then the system of equations representing the enforcement of the boundary condition of Eq. (1) at
each of the control points can be written
A = v .

(2)

Since the elements of A and v are known, Eq. (2) represents a linear system of equations that can
be solved for the N unknown values j .
The net lift on the airfoil is then given by the Kutta-Joukowsky theorem as
` = U

N
X
j=1

THE VORTEX LATTICE METHOD

whence the lift coefficient is


PN
N
U j=1 j
2 X
`
=
j
C` = 1 2 =
1
2
U c j=1
2 U c
2 U c

or

C` = 2

N
X

(3)

j=1

if we interpret to be normalized by the product U c (or, equivalently, take U = c = 1).


The pitching moment referenced to the quarter-chord point of the airfoil is similarly given by the
sum of the contributions of the individual lifting forces as

N
X
1
j x v j
mc/4 = U
4
j=1
whence the moment coefficient is
Cmc/4

mc/4
= 1 2 2 =
2 U c

PN

j=1

j xvj 41
1
2
2Uc

or
Cmc/4 = 2

N
2 X
1
= 2
j x v j
U c j=1
4

N
X

1
j x v j
4
j=1

(4)

if we again interpret to be normalized by the product U c.


Since we have lumped the entire contribution of the continuous vorticity distribution (x) over each
panel into a single point vortex, an approximation of the continuous distribution can be determined
from
(xvj )xj = j
or
(xvj ) =

j
.
xj

(5)

The jump in chordwise velocity across the vortex sheet is given by


u(x) = (x) .
To first order, only the chordwise component of velocity contributes to changes in pressure, so from
the (incompressible) Bernoulli equation

1 2
U V2
2

1
= U 2 1 (1 + u)2 v 2 = U 2 u
2

p = p p =

so the change in pressure coefficient across the vortex sheet is


Cp =

U 2 u
= 2u
1
2
2 U

whence
Cp (x) = 2(x)

(6)

That is, the net lifting pressure difference across the camber line (or, equivalently, the vortex sheet)
is simply 2(x).

CLASSICAL THIN-AIRFOIL THEORY

Classical Thin-Airfoil Theory

In classical thin-airfoil theory, the boundary condition of Eq. (1) is satisfied by a continuous distribution (x) of vorticity along the chord line. This distribution of vorticity induces the vertical
velocity
Z 1
()
1
d
v(x) =
2 0 x
at any point on the chord line, so we must solve the integral equation
Z 1
()
1
dyc
d =

(7)
2 0 x
dx
to determine the vorticity distribution for a given camber line shape at a given angle of attack .
Equation (7) is the continuous analog of Eq. (2).

The solution to Eq. (7) can be obtained by introducing the change of variables
=

1 cos
2

x=

1 cos
,
2

and

(8)

following which Eq. (7) becomes


1
2

dyc
() sin
d =
.
cos cos
dx

The vorticity distribution can then be represented by the infinite series


#
"

X
An sin n .
() = 2 A0 cot +
2 n=1

(9)

(10)

The Kutta Condition requires that the vorticity strength go to zero at the trailing edge. Since the
trailing edge is located at = , and cot /2 = 0 and sin n = 0 for all integer values of n, the
above distribution is seen to satisfy the Kutta Condition () = 0 automatically.
All the terms that need to be integrated once the vorticity distribution of Eq. (10) is substituted
into Eq. (9) can be evaluated using the Glauert Integral (see, e.g., [4])
Z
sin n
cos n
d =
.
(11)
sin
0 cos cos
Thus, substitution of the vorticity distribution of Eq. (10) into the integral Eq. (9), using the Glauert
Integral results in

X
dyc
(12)
A0
An cos n =
dx
n=1
Integrating this equation from 0 to gives

A0 =

dyc
d,
dx

while multiplying by cos n and integrating from 0 to gives


Z
2 dyc
cos n d, for n 1.
An =
0 dx

(13)

(14)

Note for future reference that the values of An for n 1 depend only on the shape of the camber
line; the only dependence on angle of attack is that shown explicitly in Eq. (13) for A 0 .

CLASSICAL THIN-AIRFOIL THEORY

2.1

The Flat Plate

The camber line for a flat plate airfoil is simply yc = 0, so Eqs. (13) and (14) give
A0 =
An = 0 for n 1.

(15)

Thus, for the flat plate airfoil


() = 2 cot

= 2
2

1 + cos
sin

or, transforming back to x,


(x) = 2

2.2

1x
x

(16)

Lift and Moment Coefficients

The airfoil lift coefficient is defined as


`
C` = 1 2 =
2 U c

1
1
2
2 U c

U (x) dx
0

or, if we interpret and x to be normalized by U and c, respectively,


Z 1
C` = 2
(x) dx .
0

Introducing the transformation of Eq. (8) and the vorticity distribution of Eq. (10), we find
#
Z "

X
An sin n sin d
C` = 2
A0 cot +
2 n=1
0
which can easily be evaluated to give
C` = [2A0 + A1 ] .

(17)

Thus, the lift coefficient is seen to depend only on the first two terms of the infinite series representing
the vorticity distribution. Also, as noted earlier, the only dependence on the angle of attack is
through the dependence of A0 shown in Eq. (13). Thus the lift curve slope C` / is given by
C`
= 2 .

(18)

Thus, the lift curve slope is the same for all thin airfoils, independent of the camber line shape, and
is equal to 2.
The moment coefficient about the leading edge of the airfoil is given by
Z c
mle
1
Cmle = 1 2 2 = 1 2
U (x)x dx ,
2 U c
2 U c 0
or, if we again interpret and x to be normalized by U and c, respectively,
Z 1
Cmle = 2
(x)x dx .
0

CLASSICAL THIN-AIRFOIL THEORY

Introducing the transformation of Eq. (8) and the vorticity distribution of Eq. (10), we find
#
Z "

X
A0 cot +
An sin n (1 cos ) sin d ,
Cmle =
2 n=1
0
which can easily be evaluated to give
Cmle

A2

A0 + A 1
=
.
2
2

(19)

Thus, the pitching moment is seen to depend only on the first three terms of the infinite series
describing the camber line.
Now, the pitching moment about any other point on the chord line, say x/c, is related to that about
the leading edge by the relation

x
0 .
Cmx/c = Cmle + C`
c
Using the results of Eqs. (17) and (19), this relation gives

x 1
1
x 1
Cmx/c = 2A0
+
A1 + A2 .

c
4
c
2
4

This equation shows that the pitching moment will be independent of A 0 and, therefore, independent of the angle of attack when
1
x
= ,
c
4
that is, when the pitching moment is measured relative to the quarter-chord point. This reference
point about which the pitching moment is independent of angle of attack is called the aerodynamic
center ; thus, thin-airfoil theory shows that the aerodynamic center is independent of the shape of the
camber line and located at the quarter chord point. The value of the moment coefficient, referenced
to the quarter-chord point, is then given by
Cmc/4 =

2.3

(A1 A2 ) .
4

(20)

Circular-arc Camber Line

The simplest non-trivial camber line is a circular arc which, for small amplitudes, can be approximated as the parabola
yc = 4 x (1 x) ,
where the parameter expresses the maximum deviation of the camber line from the chord, as a
fraction of chord length. The camber line slope is thus found to be
dyc
= 4 (1 2x) ,
dx
which, when the transformation of Eq. (8) is used, becomes
yc = 4 cos .
For this camber line, then, Eqs. (13) and (14) give the Fourier coefficients as simply
A0 = ,
A1 = 4 ,
An = 0 for n 2.

(21)

COMPARISONS WITH NONLINEAR THEORY

The lift and moment coefficients are then seen to be


C` = 2 ( + 2 ) ,
Cmc/4 = .

(22)

The expression for the lift coefficient can be used to see that the angle for zero lift is
0 = 2 .

(23)

In these results, it is seen that both the pitching moment about the aerodynamic center and the
angle for zero lift are proportional to the amplitude of the camber line. While the result here has
been shown only for the case of circular-arc camber, this is a general result (following from the
linearity of the equations we solve in the thin-airfoil approximation). That is, for a camber line of
any shape, given by, say
yc = f (x)
both the angle for zero lift and the pitching moment about the aerodynamic center will be directly
proportional to the parameter .

2.3.1

Connection to Vortex Lattice Method

Note that if we use a vortex lattice approximation to represent the flow past the circular-arc camber
line using only one panel , we will find the vortex strength to be given by
= ( + 2 ) .
The lift coefficient is therefore given by
C` = 2 = 2 ( + 2 ) ,

(24)

which agrees exactly with the result given above in Eq. (22). Note that it is the half-chord separation
between the vortex position and the control point that gives us the correct lift-curve slope of 2,
while the specific location of the control point gives us the correct angle for zero lift. This singlepanel approximation for the flow also gives us a constant moment about the quarter-chord point,
but gives the incorrect value of zero for this moment. This reproduction of the exact lift coefficient
for the simplest non-trivial camber line, when using a single vortex panel, provides the motivation
for our placement of the vortex and control point in the vortex lattice method.

Comparisons with Nonlinear Theory

In this section we compare the results of thin-airfoil theory with full non-linear, but inviscid, theory.
The non-linear results are calculated using a numerical procedure [3] to solve for the inviscid, compressible flow past complete airfoils. These calculations are, in fact, performed at very low Mach
numbers, typically M = 0.05, so the results can be interpreted as equivalent to linear, incompressible flow computations. So the only difference between these results and those of thin-airfoil theory,
are due to the fact that thickness effects are taken into account. Note, in particular, that viscous
effects still are neglected in these computations.
We study the flow past the five-digit NACA 230xx camber line at its design angle of attack = 1.65 .
This camber line has its maximum amplitude at the 15 per cent chord station, has zero curvature aft
of a point just behind the maximum camber station, and has a design lift coefficient of C ` = 0.30 [1].
Figure 2 shows the camber line shape, the camber line slope, and the total lifting pressure distribution

REFERENCES

Thin Airfoil Theory


2.5
y (x 10)
c

C = 0.30087
l

m(c/4)

dy/dx
Delta Cp

= 0.012838

1.5

0.5

0.5

0.1

0.2

0.3

0.4

0.5
x

0.6

0.7

0.8

0.9

Figure 2: Thin-airfoil solution for NACA 230 camber line at design angle of attack = 1.65 .
Cp , as functions of chordwise position, computed according to the thin-airfoil approximation.
Although this solution was computed numerically using 128 vortex panels, it can be considered to
be essentially an exact solution to the thin-airfoil problem.
The results of the full, inviscid computations are shown in Fig. 3. Figures 3 (a)-(d) show that the
actual pressure distributions vary significantly with the added contribution due to thickness, but
Figs. 3 (e) and (f) show that the net lifting pressure distributions vary only weakly with thickness.
Figure 4 quantifies the effect of thickness ratio on the lift and quarter-chord moment coefficients.
The moment coefficient is seen to be nearly independent of thickness ratio, while the lift coefficient
varies more strongly, but differs from the thin-airfoil value by only about 15% for a thickness ratio
of 12 per cent.

References
[1] Ira H. Abbott & Albert E. von Doenhoff, Theory of Wing Sections, Dover, New York, 1959.
[2] John J. Bertin, Aerodynamics for Engineers, Fourth Edition, Prentice-Hall, New York,
2002.
[3] A. Jameson & D. A. Caughey, How Many Steps are Required to Solve the Euler Equations of
Steady, Compressible Flow: In Search of a Fast Solution Algorithm, AIAA Paper 2001-2673,
AIAA 15th Computational Fluid Dynamics Conference, June 11-14, Anaheim, California.
[4] L. M. Milne-Thompson, Theoretical Aerodynamics, Dover, New York, 1958.
[5] Richard S. Shevell, Fundamentals of Flight, Second Edition, Prentice-Hall, New York, 1989.

REFERENCES

(a) NACA 23003

(b) NACA 23012

Surface Pressure, Total Enthalpy, and Entropy Change Distributions


-2

Surface Pressure, Total Enthalpy, and Entropy Change Distributions


-2

Pressure coefficient
Total enthalpy (x100)
Total pressure (x10)

-1.5

-1
Pressure Coefficient

-1
Pressure Coefficient

Pressure coefficient
Total enthalpy (x100)
Total pressure (x10)

-1.5

-0.5

-0.5

0.5

0.5

1.5

1.5
0

0.2

0.4
0.6
Chordwise position, x/c

0.8

0.2

(c) NACA 23003

0.8

Lifting Pressure Coefficient


Lifting Pressure Coefficient
Thin Airfoil Theory

-1.5

-1
Pressure Coefficient

-1
Pressure Coefficient

Lifting Surface Pressure Distribution


-2

Lifting Pressure Coefficient


Lifting Pressure Coefficient
Thin Airfoil Theory

-1.5

0.8

(d) NACA 23012

Lifting Surface Pressure Distribution


-2

0.4
0.6
Chordwise position, x/c

-0.5

-0.5

0.5

0.5

1.5

1.5
0

0.2

0.4
0.6
Chordwise position, x/c

(e) NACA 23003

0.8

0.2

0.4
0.6
Chordwise position, x/c

(f) NACA 23012

Figure 3: Full inviscid solutions for flow past NACA 23003 and 23012 airfoils at design incidence
= 1.65 and M = 0.05. (a), (b) show contours of constant pressure coefficient in Cp = 0.05
increments in the vicinity of the airfoil surface; (c), (d) show surface pressure distributions; (e), (f)
compare lifting pressure distributions with those of thin-airfoil theory for the same camber line.

10

REFERENCES

NACA 230xx Family of Airfoils


0.4

Lift
Moment

0.35

Force/Moment Coefficient

0.3
0.25
0.2
0.15
0.1
0.05
0
-0.05
0

0.02

0.04

0.06
0.08
Thickness ratio, t/c

0.1

0.12

0.14

Figure 4: Lift and quarter-chord moment coefficients for airfoils with NACA 230 camber line at
design angle of attack = 1.65 as functions of thickness ratio. Finite thickness ratio results are
computed using full, nonlinear theory; zero thickness ratio result is from thin-airfoil theory.

You might also like