Você está na página 1de 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 eects.
1 The Vortex Lattice Method
We here describe the implementation of the vortex lattice method for two-dimensional ows past
thin airfoils. The method is even more useful for three-dimensional wings, i.e., for the ow past
wings of nite 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 nite number of discrete vortices along the chord line, with the boundary condition that
the induced vertical velocity
v =
dy
c
dx
, (1)
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 dy
c
/ dx; the magnitude of the free stream velocity is taken to be unity.
Thus, we discretize the chord line into a nite 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 v
i,j
induced at the ith control point by the jth point vortex is given by
v
i,j
=

j
2
1
x
vj
x
ci
where x
vj
is the chordwise coordinate of the jth vortex having strength
j
, and x
ci
is the chordwise
coordinate of the ith control point. The total vertical velocity at the ith control point induced by
1 THE VORTEX LATTICE METHOD 2
y
x
x x
i i+1

i
x x
v c
(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
v
i
=
N

j=1

j
2
1
x
vj
x
ci
=
N

j=1
a
i,j

j
where
a
i,j
=
1
2
_
x
vj
x
ci
_
is the inuence coecient representing the eect 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 = [v
1
v
2
. . . v
N
]
T
and
= [
1

2
. . .
N
]
T
,
and dene the matrix of inuence coecients
A =
_
_
_
_
_
_
_
_
a
1,1
a
1,2
a
1,N
a
2,1
a
2,2
a
2,N



a
N,1
a
N,2
a
N,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

j=1

j
1 THE VORTEX LATTICE METHOD 3
whence the lift coecient is
C

=

1
2
U
2
c
=
U

N
j=1

j
1
2
U
2
c
=
2
Uc
N

j=1

j
or
C

= 2
N

j=1

j
(3)
if we interpret to be normalized by the product Uc (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
m
c/4
= U
N

j=1

j
_
x
vj

1
4
_
whence the moment coecient is
C
m
c/4
=
m
c/4
1
2
U
2
c
2
=

N
j=1

j
_
x
vj

1
4
_
1
2
Uc
2
=
2
Uc
2
N

j=1

j
_
x
vj

1
4
_
or
C
m
c/4
= 2
N

j=1

j
_
x
vj

1
4
_
(4)
if we again interpret to be normalized by the product Uc.
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
(x
vj
)x
j
=
j
or
(x
vj
) =

j
x
j
. (5)
The jump in chordwise velocity across the vortex sheet is given by
u(x) = (x) .
To rst order, only the chordwise component of velocity contributes to changes in pressure, so from
the (incompressible) Bernoulli equation
p = p p

=
1
2

_
U
2
V
2

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

= U
2
u
so the change in pressure coecient across the vortex sheet is
C
p
=
U
2
u
1
2
U
2
= 2u
whence
C
p
(x) = 2(x) (6)
That is, the net lifting pressure dierence across the camber line (or, equivalently, the vortex sheet)
is simply 2(x).
2 CLASSICAL THIN-AIRFOIL THEORY 4
2 Classical Thin-Airfoil Theory
In classical thin-airfoil theory, the boundary condition of Eq. (1) is satised by a continuous dis-
tribution (x) of vorticity along the chord line. This distribution of vorticity induces the vertical
velocity
v(x) =
1
2
_
1
0
()
x
d
at any point on the chord line, so we must solve the integral equation
1
2
_
1
0
()
x
d =
dy
c
dx
(7)
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
and
x =
1 cos
2
, (8)
following which Eq. (7) becomes
1
2
_

0
() sin
cos cos
d =
dy
c
dx
. (9)
The vorticity distribution can then be represented by the innite series
() = 2
_
A
0
cot

2
+

n=1
A
n
sin n
_
. (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 sinn = 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])
_

0
cos n
cos cos
d =
sinn
sin
. (11)
Thus, substitution of the vorticity distribution of Eq. (10) into the integral Eq. (9), using the Glauert
Integral results in
A
0

n=1
A
n
cos n =
dy
c
dx
(12)
Integrating this equation from 0 to gives
A
0
=
1

_

0
dy
c
dx
d, (13)
while multiplying by cos n and integrating from 0 to gives
A
n
=
2

_

0
dy
c
dx
cos n d, for n 1. (14)
Note for future reference that the values of A
n
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
.
2 CLASSICAL THIN-AIRFOIL THEORY 5
2.1 The Flat Plate
The camber line for a at plate airfoil is simply y
c
= 0, so Eqs. (13) and (14) give
A
0
=
A
n
= 0 for n 1.
(15)
Thus, for the at plate airfoil
() = 2cot

2
= 2
_
1 + cos
sin
_
or, transforming back to x,
(x) = 2
_
1 x
x
(16)
2.2 Lift and Moment Coecients
The airfoil lift coecient is dened as
C

=

1
2
U
2
c
=
1
1
2
U
2
c
_
c
0
U(x) dx
or, if we interpret and x to be normalized by U and c, respectively,
C

= 2
_
1
0
(x) dx.
Introducing the transformation of Eq. (8) and the vorticity distribution of Eq. (10), we nd
C

= 2
_

0
_
A
0
cot

2
+

n=1
A
n
sin n
_
sin d
which can easily be evaluated to give
C

= [2A
0
+ A
1
] . (17)
Thus, the lift coecient is seen to depend only on the rst two terms of the innite series representing
the vorticity distribution. Also, as noted earlier, the only dependence on the angle of attack is
through the dependence of A
0
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 coecient about the leading edge of the airfoil is given by
C
m
le
=
m
le
1
2
U
2
c
2
=
1
1
2
U
2
c
_
c
0
U(x)xdx,
or, if we again interpret and x to be normalized by U and c, respectively,
C
m
le
= 2
_
1
0
(x)xdx.
2 CLASSICAL THIN-AIRFOIL THEORY 6
Introducing the transformation of Eq. (8) and the vorticity distribution of Eq. (10), we nd
C
m
le
=
_

0
_
A
0
cot

2
+

n=1
A
n
sin n
_
(1 cos ) sin d ,
which can easily be evaluated to give
C
m
le
=

2
_
A
0
+ A
1

A
2
2
_
. (19)
Thus, the pitching moment is seen to depend only on the rst three terms of the innite 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
C
m
x/c
= C
m
le
+C

_
x
c
0
_
.
Using the results of Eqs. (17) and (19), this relation gives
C
m
x/c
= 2A
0
_
x
c

1
4
_
+
__
x
c

1
2
_
A
1
+
1
4
A
2
_
.
This equation shows that the pitching moment will be independent of A
0
and, therefore, indepen-
dent of the angle of attack when
x
c
=
1
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 coecient, referenced
to the quarter-chord point, is then given by
C
m
c/4
=

4
(A
1
A
2
) . (20)
2.3 Circular-arc Camber Line
The simplest non-trivial camber line is a circular arc which, for small amplitudes, can be approxi-
mated as the parabola
y
c
= 4x(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
dy
c
dx
= 4 (1 2x) ,
which, when the transformation of Eq. (8) is used, becomes
y
c
= 4 cos .
For this camber line, then, Eqs. (13) and (14) give the Fourier coecients as simply
A
0
= ,
A
1
= 4 ,
A
n
= 0 for n 2.
(21)
3 COMPARISONS WITH NONLINEAR THEORY 7
The lift and moment coecients are then seen to be
C

= 2 ( + 2) ,
C
m
c/4
= .
(22)
The expression for the lift coecient 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
y
c
= 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 ow past the circular-arc camber
line using only one panel , we will nd the vortex strength to be given by
= ( + 2) .
The lift coecient 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 specic location of the control point gives us the correct angle for zero lift. This single-
panel approximation for the ow 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 coecient
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.
3 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, com-
pressible ow 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, incompress-


ible ow computations. So the only dierence between these results and those of thin-airfoil theory,
are due to the fact that thickness eects are taken into account. Note, in particular, that viscous
eects still are neglected in these computations.
We study the ow past the ve-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 coecient of C

= 0.30 [1].
Figure 2 shows the camber line shape, the camber line slope, and the total lifting pressure distribution
REFERENCES 8
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.5
0
0.5
1
1.5
2
2.5
Thin Airfoil Theory
x
C
l
= 0.30087
C
m(c/4)
= 0.012838
y
c
(x 10)
dy/dx
Delta C
p
Figure 2: Thin-airfoil solution for NACA 230 camber line at design angle of attack = 1.65

.
C
p
, 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 signicantly 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 quanties the eect of thickness ratio on the lift and quarter-chord moment coecients.
The moment coecient is seen to be nearly independent of thickness ratio, while the lift coecient
varies more strongly, but diers 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 Doenho, 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 9
(a) NACA 23003 (b) NACA 23012
-2
-1.5
-1
-0.5
0
0.5
1
1.5
0 0.2 0.4 0.6 0.8 1
P
r
e
s
s
u
r
e

C
o
e
f
f
i
c
i
e
n
t
Chordwise position, x/c
Surface Pressure, Total Enthalpy, and Entropy Change Distributions
Pressure coefficient
Total enthalpy (x100)
Total pressure (x10)
-2
-1.5
-1
-0.5
0
0.5
1
1.5
0 0.2 0.4 0.6 0.8 1
P
r
e
s
s
u
r
e

C
o
e
f
f
i
c
i
e
n
t
Chordwise position, x/c
Surface Pressure, Total Enthalpy, and Entropy Change Distributions
Pressure coefficient
Total enthalpy (x100)
Total pressure (x10)
(c) NACA 23003 (d) NACA 23012
-2
-1.5
-1
-0.5
0
0.5
1
1.5
0 0.2 0.4 0.6 0.8 1
P
r
e
s
s
u
r
e

C
o
e
f
f
i
c
i
e
n
t
Chordwise position, x/c
Lifting Surface Pressure Distribution
Lifting Pressure Coefficient
Lifting Pressure Coefficient
Thin Airfoil Theory
-2
-1.5
-1
-0.5
0
0.5
1
1.5
0 0.2 0.4 0.6 0.8 1
P
r
e
s
s
u
r
e

C
o
e
f
f
i
c
i
e
n
t
Chordwise position, x/c
Lifting Surface Pressure Distribution
Lifting Pressure Coefficient
Lifting Pressure Coefficient
Thin Airfoil Theory
(e) NACA 23003 (f) NACA 23012
Figure 3: Full inviscid solutions for ow past NACA 23003 and 23012 airfoils at design incidence
= 1.65

and M

= 0.05. (a), (b) show contours of constant pressure coecient in C


p
= 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.
REFERENCES 10
-0.05
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
F
o
r
c
e
/
M
o
m
e
n
t

C
o
e
f
f
i
c
i
e
n
t
Thickness ratio, t/c
NACA 230xx Family of Airfoils
Lift
Moment
Figure 4: Lift and quarter-chord moment coecients 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.

Você também pode gostar