Vous êtes sur la page 1sur 10

Munson, B. R., Cronin, D. J.

Airfoils/Wings
The Engineering Handbook.
Ed. Richard C. Dorf
Boca Raton: CRC Press LLC, 2000























































1998 by CRC PRESS LLC
36
Airfoils/Wings
36.1 Nomenclature
36.2 Airfoil Shapes
36.3 Lift and Drag Characteristics for Airfoils
36.4 Lift and Drag of Wings
Bruce R. Munson
Iowa State University
Dennis J. Cronin
Iowa State University
A simplified sketch of a wing is shown in Fig. 36.1. An airfoil is any cross section of the wing
made by a plane parallel to the xz plane. The airfoil size and shape usually vary along the
span.
Airfoils and wings are designed to generate a lift force, L, normal to the free stream flow that is
considerably larger than the drag force, D, parallel to the free stream flow. The lift and drag are
strongly dependent on the geometry (shape, size, orientation to the flow) of the wing and the speed
at which it flies, V
o
, as well as other parameters, including the density, ; viscosity, ; and speed of
sound, a, of the air. The following sections discuss some properties of airfoils and
wings.
36.1 Nomenclature
The shape, size, and orientation of an airfoil can be given in terms of the following parameters
(Fig. 36.2): the chord length, c, the chord line that connects the leading and trailing edges; the
angle of attack, , relative to the free stream velocity, V
o
; the mean camber line that is halfway
between the upper and lower surfaces; and the thickness distribution, t, which is the distance
between the upper and lower surfaces perpendicular to the camber line.
Various classes of airfoils have been developed over the years. These include the classic
National Advisory Committee for Aeronautics four-, five-, and six-digit series airfoils (for
example, the NACA 2412 airfoil used on the Cessna 150 or the NACA 64A109 used on the Gates
Learjet [Anderson, 1991]) as well as numerous other modern airfoils [Hubin, 1992].
Performance characteristics of wings are normally given in terms of the dimensionless lift
coefficient and drag coefficient,























































1998 by CRC PRESS LLC
Figure 36.2 Airfoil geometry.
Figure 36.1 Wing geometry.
Figure 36.3 Typical lift and drag coefficients as a function of angle of attack.























































1998 by CRC PRESS LLC
C
L
=
L
q
o
S
(36:1)
C
D
=
D
q
o
S
(36:2)
where q
o
=
1
2
V
2
o
is the dynamic pressure and S is the planform area of the wing. The planform
area is the area seen by looking onto the wing from above: the span times the chord for a
rectangular wing. Typical characteristics for lift and drag coefficients as a function of the angle of
attack are shown in Fig. 36.3. An efficient wing has a large lift-to-drag ratiothat is, a large
C
L
=C
D
.
FLYING MACHINE
Orville & Wilbur Wright
Patented May 22, 1906
#821,393
An excerpt:
Our invention relates to that class of flying-machines in which the weight is sustained by the
reactions resulting when one or more aeroplanes are moved through the air edgewise at a small
angle of incidence, either by the application of mechanical power or by the utilization of the force
of gravity.
The objects of our invention are to provide means for maintaining or restoring the equilibrium or
lateral balance of the apparatus, to provide means for guiding the machine both vertically and
horizontally, and to provide a structure combining lightness, strength, convenience of construction,
and certain other advantages which hereinafter appear.
The Wrights became interested in flight after reading of successful glider flights in Germany. They built
and flew three glider bi-planes before attempting their unassisted powered flight at Kitty Hawk, N.
Carolina in 1903. The U.S. Army was interested in powered flight and eventually awarded the Wrights a
contract for the first military aircraft in 1909. (1992, DewRay Products, Inc. Used with
permission.)























































1998 by CRC PRESS LLC
36.2 Airfoil Shapes
As shown in Fig. 36.4, typical airfoil shapes have changed over the years in response to changes in
flight requirements and because of increased knowledge of flow properties. Early airplanes used
thin airfoils (maximum thickness 6 to 8% of the chord length), with only slight camber [Fig.
36.4(a)]. Subsequently, thicker (12 to 18% maximum thickness) airfoils were developed and used
successfully on a variety of low-speed aircraft [Fig. 36.4(b) and (c)].
Figure 36.4 Various airfoil shapes.
As the angle of attack is increased from small values, the lift coefficient increases nearly linearly
with . At larger angles there is a sudden decrease in lift and a large increase in drag. This
condition indicates that the wing has stalled. The airflow has separated from the upper surface, and
an area of reverse flow exists (Fig. 36.3). Stall is a manifestation of boundary layer separation.
This complex phenomenon is a result of viscous effects within a thin air layer (the boundary layer)
near the upper surface of the wing in which viscous effects are important [Schlichting, 1979].
In addition to knowing the lift and drag for an airfoil, it is often necessary to know the location
where these forces act. This location, the center of pressure, is important in determining the
moments that tend to pitch the nose of the airplane up or down. Such information is often given in
terms of a moment coefficient,
C
M
=
M
q
o
Sc
(36:3)
where M is the moment of the lift and drag forces about some specified point, often the leading
edge.
For a given geometry, the lift and drag coefficients and the center of pressure (or moment
coefficient) may depend on the flight speed and properties of the air. This dependence can be
characterized in terms of the Reynolds number based on chord length, Re = V
o
c= , and the
Mach number, Ma = V
o
=a . For modern commercial aircraft the Reynolds number is typically on
the order of millions (10
6
) . Mach numbers range from less than 1 (subsonic flight) to greater than
1 (supersonic flight).























































1998 by CRC PRESS LLC
36.3 Lift and Drag Characteristics for Airfoils
Lift and drag forces on airfoils are the result of pressure and viscous forces that the air imposes on
the airfoil surfaces. Pressure is the dominant factor that produces lift. A typical pressure
distribution is shown in Fig. 36.5. In simplistic terms, the air travels faster over the upper surface
than it does over the lower surface. Hence, from Bernoulli's principle for steady flow, the
pressure on the upper surface is lower than that on the bottom surface.
Figure 36.5 Typical pressure distribution on an airfoil.
For an unstalled airfoil, most of the drag is due to viscous forces. This skin friction drag is a
result of the shear stress distribution on the airfoil. For a stalled airfoil, pressure forces contribute
significantly to the drag.
For a two-dimensional body such as an airfoil (a wing of infinite span), the section lift, drag, and
moment coefficients (about a defined point) are based on the lift, drag, and moment per unit span,
L
0
[force=length ] , D
0
[force=length ] , and M
0
[force length=length ] , respectively. That
is,
A relatively recent development (c. 1970s) has been design and construction of laminar flow
airfoils that have smaller drag than previously obtainable [Fig. 36.4(d)]. This has resulted from
new airfoil shapes and smooth surface construction that allow the flow over most of the airfoil to
remain laminar rather than become turbulent. The performance of such airfoils can be quite
sensitive to surface roughness (e.g., insects, ice, and rain) and Reynolds number
effects.
With the advent of commercial and business jet aircraft, it became necessary to develop airfoils
that operate properly at Mach numbers close to unity. Since air, in general, accelerates as it passes
around an airfoil, the flow may be locally supersonic near portions of the upper surface, even
though the flight speed is subsonic. Such supersonic flow can cause shock waves (discontinuities
in the flow) that degrade airfoil performance. Airfoils have been developed (denoted supercritical
airfoils) to minimize the effect of the shock wave in this locally supersonic region [Fig. 36.4(e)].
Certain flow phenomena, such as shock waves, occur in supersonic flight that do not occur for
subsonic flight [Anderson, 1991]. The result is that supersonic airfoils tend to be thinner and
sharper than those for subsonic flight [see Fig. 36.4(f)].























































1998 by CRC PRESS LLC
c
l
=
L
0
q
o
c
(36:4)
c
d
=
D
0
q
o
c
(36:5)
c
m
=
M
0
q
o
c
2
(36:6)
For most airfoils, the lift coefficient is nearly linear with angle of attack up to the stall angle.
According to simple airfoil theory [and verified by experiment (Fig. 36.6)], the lift-curve slope,
dc
l
=d , is approximately equal to 2 with in radians (or dc
l
=d = 0:1096 deg
1
when is in
degrees) [Anderson, 1991].
Figure 36.6 Effect of camber on airfoil lift coefficient.
For symmetrical (no camber) airfoils, zero angle of attack produces zero lift; for airfoils with
camber, the zero-lift condition ( =
0L
) occurs at nonzero angle of attack (Fig. 36.6).
The maximum lift coefficient for an airfoil (c
l
= c
l;max
) is typically on the order of unity and
occurs at the critical angle of attack ( =
CR
) (Fig. 36.3). That is, the lift generated per unit span
is on the order of the dynamic pressure times the planform area: L
0
= c
l
q
o
c q
0
c . The drag























































1998 by CRC PRESS LLC
coefficient, on the other hand, is on the order of 0.01. Hence, the maximum lift-to-drag ratio is on
the order of 100.
As illustrated in Fig. 36.7, the airfoil geometry may be altered by using movable trailing or
leading edge flaps. Such devices can significantly improve low-speed (i.e., landing or takeoff)
performance by increasing the maximum lift coefficient, thereby reducing the required landing or
takeoff speed.
Figure 36.7 Effect of flaps on airfoil lift coefficient.
All wings have a finite span, b, with two wing tips. The flow near the tips can greatly influence the
36.4 Lift and Drag of Wings
flow characteristics over the entire wing. Hence, a wing has different lift and drag coefficients than
those for the corresponding airfoil. That is, the lift and drag coefficients for a wing are a function
of the aspect ratio, AR = b
2
=S . For a wing of rectangular planform (i.e., constant chord), the
aspect ratio is simply b=c.
Because of the pressure difference between the lower and upper surfaces of a wing, the air tends
to "leak" around the wing tips (bottom to top) and produce a swirling flowthe trailing or wing tip
vortices shown in Fig. 36.8. This swirl interacts with the flow over the entire length of the wing,
thereby affecting its lift and drag. The trailing vortices create a flow that makes it appear as though
the wing were flying at an angle of attack different from the actual angle. This effect produces
additional drag termed the induced drag.























































1998 by CRC PRESS LLC
As shown by theory and experiment [Anderson, 1991], the larger the aspect ratio is, the larger
the lift coefficient is and the smaller the drag coefficient is (Fig. 36.9). A wing with an infinite
aspect ratio would have the properties of its airfoil section. Very efficient flyers (i.e., soaring birds
and sailplanes) have long, slender (large AR) wings.
Figure 36.8 Trailing vortex
Figure 36.9 Lift coefficient as a function
For many wings the chord length decreases from the wing root (next to the aircraft body) to the
wing tip. In addition, the shape of the airfoil section may change from root to tip, as may the local
angle of attack (i.e., the wing may have some "twist" to it).
Figure 36.10 Various wing planforms.
of aspect ratio.























































1998 by CRC PRESS LLC
Although wind tunnel tests of model airfoils and wings still provide valuable (and sometimes
unexpected) information, modern computational fluid dynamic (CFD) techniques are widely used.
Techniques involving paneling methods, finite elements, boundary elements, finite differences, and
viscous-inviscid interaction are among the powerful tools currently available to the aerodynamicist
[Moran, 1984].
Defining Terms
Airfoil: The cross section of a wing, front to back.
Angle of attack: The angle between a line connecting the leading and trailing edges of an airfoil
and the free stream velocity.
Bernoulli's principle: Conservation of energy principle that states that an increase in flow speed
is accompanied by a decrease in pressure and vice-versa.
Camber: Maximum distance between the chord line and the camber line.
Center of pressure: Point of application of the lift and drag forces.
Chord: Distance between the leading and trailing edges of an airfoil.
Dynamic pressure: Pressure increase resulting from the conversion of kinetic energy into
pressure.
Flaps: Leading and trailing edge devices used to modify the geometry of an
airfoil.
Lift and drag coefficients: Lift and drag made dimensionless by dynamic pressure and wing
area.
Moment coefficient: Pitching moment made dimensionless by dynamic pressure, wing area, and
chord length.
Planform: Shape of a wing as viewed from directly above it.
Span: Distance between the tips of a wing.
Stall: Sudden decrease in lift as angle of attack is increased to the point where flow separation
occurs.
References
Anderson, J. D. 1990. Modern Compressible Flow with Historical Perspective, 2nd ed.
McGraw-Hill, New York.
Anderson, J. D. 1991. Fundamentals of Aerodynamics, 2nd ed. McGraw-Hill, New York.
Hubin, W. N. 1992. The Science of Flight: Pilot Oriented Aerodynamics. Iowa State University
Press, Ames, IA.
Moran, J. 1984. An Introduction to Theoretical and Computational Aerodynamics. John Wiley
& Sons, New York.
Schlichting, H. 1979. Boundary Layer Theory, 7th ed. McGraw-Hill, New York.
Further Information
Abbott, I. H. and van Doenhoff, A. E. 1949. Theory of Wing Sections. McGraw-Hill, New York.
Anderson, D. A., Tannehill, J. C., and Pletcher, R. H. 1984. Computational Fluid Mechanics and
Heat Transfer. Hemisphere, New York.
Anderson, J. D. 1985. Introduction to Flight, 2nd ed. McGraw-Hill, New York.
Many modern high-speed wings are swept back with a V-shaped planform; others are delta
wings with a triangular planform (Fig. 36.10). Such designs take advantage of characteristics
associated with high-speed compressible flow [Anderson, 1990].























































1998 by CRC PRESS LLC

Vous aimerez peut-être aussi