You are on page 1of 43

AERODYNAMIC COEFFICIENTS

The aerodynamic characteristics of a body are more fundamentally described by the force and moment coefficients than by the actual forces and moments themselves. aerodynamic force on a body depends on: - velocity of the body through the air the density of the ambient air - size of the body - orientation of the body relative to the free-stream direction, (angle of attack) (Clearly, if we change the velocity, the aerodynamic force should change. Also, the force on a body moving at 100 feet per second through air is going to be smaller than the force on the same body moving at the same velocity through water, which is nearly a thousand times denser than air. Also, the aerodynamic force on a sphere of 1-inch diameter is going to be smaller than that for a sphere of 1-ft diameter, everything else being equal. Finally, the force on a wing will clearly depend on how much the wing is inclined to the flow. - Moreover, since friction accounts for part of the aerodynamic force, the force should depend on the ambient coefficient of viscosity. - Also important is the compressibility of the medium through which the body moves. A measure of the compressibility of a fluid is the speed of sound in the fluid the higher the compressibility, the lower the speed of sound.

Hence we have

If we want to study how L,D,M depend on these variables we have vary one and keep the others constant. With 6 unknowns it could be very time-consuming, and moreover, the large amount of wind tunnel time could be quite costly. But the amount of unknowns can be reduced using the non dimensional groups: Reynolds number Dynamic pressure

Mach number

Imagine that we have a given body at a given angle of attack in a given flow, where p, V, density, and a, are certain values. Let us call this the "green" flow. Consider another body of the same geometric shape (but not the same size) in another flow where p, V, density and a, are all different; let us call this flow the "red" flow. Dimensional analysis, tells us that even though the green flows and the red flow are two different flows, if the Reynolds number and the Mach number are the same for these two different flows, then the lift coefficient will be the same for the two geometrically similar bodies at the same angle of attack. The two flows, the green flow and the red flow, are called dynamically similar.

Variation of Cl with the angle of attack and Reynolds


The slope of this linear portion is called the lift slope and is designated by a0. For thin airfoils, a theoretical value for the lift slope is 2pi per radiant, or 0.11 per degree.
L =0

there is a finite value of Cl at zero angle of attack, and that the airfoil must be pitched down to some negative angle of attack for the lift to be zero. This angle of attack is denoted by L =0 If positively cambered airfoils have negative zero-lift angles of attack. In contrast, symmetric airfoil has L=0 =0 a negatively cambered airfoil has a positive

L=0

At the other extreme, at high angles of attack, the lift coefficient becomes nonlinear, reaches a maximum value denoted by Cl max then drops as a further increased.

This is because a separation occurs over the top surface of the airfoil and the lift decreases (sometimes precipitously). In this condition, the airfoil is said to be stalled. In contrast, over the linear portion of the lift curve, the flow is attached over most of the airfoil surface. the linear portion of the lift curve is essentially insensitive to variations in Re. By increasing Reynolds number Clmax increases

Variation of Cm with the angle of attack and Reynolds

over most of the practical range of the angle of attack the slope of the moment coefficient curve is essentially constant. This slope is positive for some airfoils (as shown here), but can be negative for other airfoils. The variation becomes nonlinear at high angle of attack, when the flow separates from the top surface of the airfoil, and at low, highly negative angles of attack, when the flow separates from the bottom surface of the airfoil.

Variation of Cd with the angle of attack and Reynolds


For a cambered airfoil, the minimum value Cd does not necessarily occur at zero angle of attack, but rather at some finite but small angle of attack. For this angle-of-attack range, the drag is due to friction drag and pressure drag. In contrast, the rapid increase in cd which occurs at higher values of alpha, is due to the increasing region of separated flow over the airfoil, which creates a large pressure drag. The friction decreases by increasing the Reynolds number. Moreover, the Reynolds number influences the extent and characteristics of the separated flow region, and hence it is no surprise that Cd at the larger values of alpha is also sensitive to the Reynolds number.

NACA AIRFOIL NOMENCLATURE

The major design feature of an airfoil is the mean camber line, which is the locus of points halfway between the upper and lower surfaces, as measured perpendicular to the mean camber line itself. The most forward and rearward points of the mean camber line are the leading and trailing edges, respectively. The straight line connecting the leading and trailing edges is the chord line of the airfoil, and the precise distance from the leading to the trailing edge measured along the chord line is simply designated the chord of the airfoil, denoted by c. The camber is the maximum distance between the mean camber line and the chord line, measured perpendicular to the chord line. The camber, the shape of the mean camber line, and, to a lesser extent, the thickness distribution of the airfoil essentially control the lift and moment characteristics of the airfoil.

First family of airfoils NACA airfoils are indicated by a series of 4 digits. The numbers in the designation mean the following: The first digit gives the maximum camber in percentage of chord. The second digit is the location of the maximum camber in tenths of chord, measured from the leading edge. The last two digits give the maximum thickness in percentage of chord. For example, the NACA 2412 airfoil has a maximum camber of 2% of the chord (or 0.02c), located at 0.4c from the leading edge. The maximum thickness is 12% of the chord (or 0.12c) Second family of airfoils The numbers mean the following: The first digit, when multiplied by 3/2, gives the design lift coefficient in tenths. The second and third digits together are a number which, when multiplied by 1/2, gives the location of maximum camber relative to the leading edge in percentage of chord. The last two digits give the maximum thickness in percentage of chord. For example, the NACA 23012 airfoil has a design lift coefficient of 0.3, the location of maximum camber at 15% of the chord (or 0.15c) from the leading edge, and a maximum thickness of 12% of the chord (or 0.12c).

THE AERODYNAMIC CENTER


The aerodynamic center is the point on a body about which the moments are independent of the angle of attack.

Differentiating with respect to angle of attack a gives

If the aerodynamic center is the point about which moments are independent of the angle of attack.

dcma . c d

=0

for a body with linear lift and moment curves, where m0 and a0 are the values, the aerodynamic center does exist as a fixed point on the airfoil.

Variation of Cl with Ma
At subsonic speeds, the "compressibility effects" associated with increasing Ma, result in a progressive increase in CI. The reason for this is that the lift is mainly due to the pressure distribution on the surface. As Ma, increases, the differences in pressure from one point to another on the surface become more pronounced. Hence, CI increases as Ma, increases. The Prandtl-Glauert rule, the first and simplest (and also the least accurate) of the several formulas for subsonic "compressibility corrections," predicts that Cl will rise inversely proportional to (1-Ma2)0.5. In the supersonic region, the dashed curve shows the theoretical supersonic variation for a thin airfoil, where CI = 4/(1-Ma2)0.5-. The oscillatory variation of Cl near Mach=1 is typical of the transonic regime, and is due to the shock wave-boundary layer interaction that is prominent for transonic Mach numbers.

Cd stays relatively constant with Ma, up to, and slightly beyond the critical Mach number (that free-stream Mach number at which sonic flow is first encountered at some location on the airfoil). The drag in the subsonic region is mainly due to friction, and the "compressibility effect" on friction in the subsonic regime is small. The flow over the airfoil in this regime is smooth and attached, with no shock waves present. As Ma increases above Ma critical, a large pocket of locally supersonic flow forms above, and sometimes also below, the airfoil. These pockets of supersonic flow are terminated at the downstream end by shock waves. The presence of these Shocks will affect the pressure distribution in such a fashion as to cause an increase in pressure drag (this drag increase is related to the loss of total pressure across the shock waves). However, the dominant effect is that the shock wave interacts with the boundary layer on the surface, causing the boundary layer to separate. Finally, in the supersonic regime, Cd gradually decreases,

Dependence of Cd with Ma

Incompressible Flow about Wings of Finite Span


For a wing of finite span, the high-pressure air beneath the wing spills out around the wing tips toward the lowpressure regions above the wing. As a consequence of the tendency of the pressures acting on the top surface near the tip of the wing to equalize with those on the bottom surface, the lift force per unit span decreases toward the tips.

Variation of lift along the span


The resultant lift force acting on a section, obtained by integrating the pressure distribution over the chord length, has a spanwise variation:

As a result of the spanwise pressure variation, the air on the upper surface flows inboard toward the root. On the lower surface, air will tend to flow outward toward the tips. The resultant flow around a wing of finite span is three dimensional, having both chordwise and spanwise velocity components.

Trailing vortices
the difference in spanwise velocity components will cause the air to roll up into a number of streamwise vortices, distributed along the span. These small vortices roll up into two large vortices just inboard of the wing tips

Visualization of tip vortices


Very high velocities and low pressures exist at the core of the wing-tip vortices. In many instances, water vapor condenses as the air is drawn into the low-pressure flow field of the tip vortices. Condensation clearly defines the tip vortices

LIFTING-LINE THEORY FOR UNSWEPT WINGS


We assume that the lift acting on an element of the wing is related to the local circulation through the Kutta-Joukowski theorem

we represent the spanwise lift distribution by a system of vortex filaments the axis of which is normal to the plane of symmetry and which passes through the aerodynamic center of the lifting surface The strength of the bound-vortex system at any spanwise location is proportional to the local lift acting at that location

Trailing vortices
the vortex theorems of Helmholtz state that a vortex filament cannot end in a fluid. Therefore, we model the lifting character of the wing by a large number of vortex filaments (infinitesimal strength filaments) that lie along the quarter chord of the wing. This is the bound-vortex system, which represents the spanwise loading distribution. When the lift changes at some spanwise location, the total strength of the boundvortex system changes proportionally. But vortex filaments cannot end in the fluid. Thus, the change is represented in our model by having some of the filaments from our bundle of filaments turn 90 degree and continue in the streamwise direction.

Trailing vortices

Lanchester's own drawing of the wing-tip vortex on a finite wing.

Downwash velocity
The strength of the trailing vortex is given by

Downwash velocity (2)


The vortex at y induces a velocity at a general point y1 on the aerodynamic centerline which is one-half the velocity that would be induced by an infinitely long vortex filament of the same strength:

Downwash velocity (3)


the resultant induced velocity at any point y1 due to the cumulative effect of all the trailing vortices is

The resultant induced velocity at y1 is in a downward direction (i.e., negative) and is called the downwash.

High-Aspect-Ratio Straight Wing


The classic theory for such wings was worked out by Prandtl during World War I and is called Prandtl's lifting line theory.

a0 =

dcl d

airfoil wing

a=

dCL d
2

lift slope per radian and e1 is a factor that depends on the geometric shape of the wing, including the aspect ratio and taper ratio.

b AR = S

Prandtl's lifting line theory does not apply to low-aspect-ratio wings. It holds for aspect ratios of about 4 or larger.

the lift slope for a finite wing decreases as the aspect ratio decreases. The angle of attack for zero lift, denoted CL=0 is the same for all the seven wings; at zero lift the induced effects theoretically disappear. At any given angle of attack larger than the value of CL becomes smaller as the aspect ratio is decreased.

Prandtl's lifting line theory, also holds for subsonic compressible flow,

where

Substituting we have

It gives a quick, but approximate correction to the lift slope; because it is derived from linear subsonic flow theory it is not recommended for use for Ma greater than 0.7. For supersonic flow over a high-aspect-ratio straight wing, the lift slope can be approximated from supersonic linear theory

Low-Aspect-Ratio Straight Wings


When applied to straight wings at AR < 4, the equations for high AR do not apply because are derived from a theoretical model which represents the finite wing with a single lifting line across the span of the wing. However, when the aspect ratio is small, the same intuition leads to some misgivings-how can a short, stubby wing be properly modeled by a single lifting line? The fact is-it cannot. Instead of a single spanwise lifting line, the low-aspect-ratio wing must be modeled by a large number of spanwise vortices, each located at a different chordwise station
Modern panel methods can quickly and accurately calculate the inviscid flow properties of low-aspect-ratio straight wings,

An approximate relation for the lift slope for low-aspect-ratio straight wings was obtained by H. B. Helmbold in Gemany in 1942 For subsonic compressible flow, is modified as follows In the case of supersonic flow over a lowaspect-ratio straight wing,

At subsonic speeds, a low-aspect-ratio wing is plagued by large induced drag, and hence subsonic aircraft (since World War I) do not have low-aspect-ratio wings. On the other hand, a low-aspect-ratio straight wing has low supersonic wave drag, and this is why such a wing was used on the F-104-the first military fighter designed for sustained Mach 2 flight. At subsonic speeds, and especially for takeoff and landing, the low-aspect-ratio wings were a major liability to the F-104.

F104 Fortunately, there are two other wing platforms that reduce wave drag without suffering nearly as large a penalty at subsonic speeds, namely, the swept wing and the delta wing.

Swept Wings
The main function of a swept wing is to reduce wave drag at transonic and supersonic speeds. Consider a straight wing and a swept wing in a flow with a free-stream velocity V. Assume that the aspect ratio is high for both wings, so that we can ignore tip effects. Let u and w be the components of V, perpendicular and parallel to the leading edge, respectively. The pressure distribution over the airfoil section oriented perpendicular to the leading edge is mainly governed by the chordwise component of velocity u; the spanwise component of velocity w has little effect on the pressure distribution. For the straight wing the chordwise velocity component u is the full V, for the swept wing the chordwise component of the velocity u is smaller than V: u = V cos

Since u for the swept wing is smaller than u for the straight wing, the difference in pressure between the top and bottom surfaces of the swept wing will be less than the difference in pressure between the top and bottom surfaces of the straight wing. Since lift is generated by these differences in pressure, the lift on the swept wing will be less than that on the straight wing. The wingspan b is the straight-line distance between the wing tips, the wing platform area is S, and the aspect ratio and the taper ratio are defined AR = b^2/S and taper ratio ct/cr.

an approximate calculation of the lift slope for a swept finite wing, Kuchemann suggests the following approach. The lift slope for an infinite swept wing should be a0 cos therefore

The subsonic compressibility effect is added by replacing

a0

with

a0 1 Ma

Supersonic Delta wings


For a swept wing moving at supersonic speeds, the aerodynamic properties depend on the location of the leading edge relative to a Mach wave emanating from the apex of the wing. The Mach angle is given by

= cos1 (1 / Ma )

If the wing leading edge is swept inside the Mach cone the component of Ma perpendicular to the leading edge is subsonic; hence, the swept wing is said to have a subsonic leading edge. For the wing in supersonic flight, there is a weak shock that emanates from the apex, but there is no shock attached elsewhere along the wing leading edge. In contrast, if the wing leading edge is swept outside the Mach cone the component of Ma, perpendicular to the leading edge is supersonic; hence the swept wing is said to have a supersonic leading edge. For this wing in supersonic flight, there will be a shock wave attached along the entire leading edge. A swept wing with a subsonic leading edge behaves somewhat as a wing at subsonic speeds, although the actual free-stream Mach number is supersonic.

Delta Wings
Swept wings that have platforms such as shown in Fig are called delta wings.

dominant aspect of this flow is the two vortices that are formed along the highly swept leading edges, and that trail downstream over the top of the wing. This vortex pattern is created by the following mechanism. The pressure on the bottom surface of the wing is higher than the pressure on the top surface.
Thus, the flow on the bottom surface in the vicinity of the leading edge tries to curl around the leading edge from the bottom to the top. If the leading edge is relatively sharp, the flow will separate along its entire length. This separated flow curls into a primary vortex above the wing just inboard of each leading edge. The stream surface which has separated at the leading edge loops above the wing and then reattaches along the primary attachment line. The primary vortex is contained within this loop. A secondary vortex is formed underneath the primary vortex, with its own separation line, and its own reattachment line. Unlike many separated flows in aerodynamics, the vortex pattern over a delta wing is a friendly flow in regard to the production of lift. The vortices are strong and generally stable. They are a source of high energy, relatively high vorticity flow, and the local static pressure in the vicinity of the vortices is small. Hence, the vortices create a lower pressure on the top surface than would exist if the vortices were not there. This increases the lift compared to what it would be without the vortices.

The difference between the experimental data and the potential flow lift is the vortex lift. The vortex lift is a major contributor to the overall lift; The lift slope is small, on the order of 0.05 per degree. The lift, however, continues to increase over a large range of angle of attack (the stalling angle of attack is about 35).

The net result is a reasonable value of CLmax=1.35. The lift curve is nonlinear, in contrast to the linear variation exhibited by conventional wings for subsonic aircraft. The vortex lift is mainly responsible for this nonlinearity. The next time you have an opportunity to watch a delta-wing aircraft take off or land, for example, the televised landing of the space shuttle, note the large angle of attack of the vehicle. Also, this is why the Concorde supersonic transport, with its low-aspect-ratio deltalike wing, lands at a high angle of attack. In fact, the angle of attack is so high that the front part of the fuselage must be mechanically drooped upon landing in order for the pilots to see the runway.

Static Aeroelasticity

Rigid flat plate mounted on a torsional spring

If the spring were very stiff or airspeed were very slow, the rotation would be rather small; however, for flexible springs or high flow velocities the rotation may twist the spring beyond its ultimate strength and lead to structural failure.

The equation of static equilibrium simply states that the sum of aerodynamic plus elastic moments about any point on the airfoil is zero. By convention, we take the point about which moments are summed as the point of spring attachment, the so-called 'elastic center' or 'elastic axis' of the airfoil. The total aerodynamic angle of attack, , is taken as the sum of some initial angle of attack, 0 (with the spring untwisted), plus an additional increment due to elastic twist of the spring e.

No changes with

For a symmetrical airfoil CL0=0

k e

If

goes to infinity

This is the divergence condition and the corresponding dynamic pressure is termed the 'divergence dynamic pressure'

You might also like