Sie sind auf Seite 1von 10

AERODYNAMIC FORCES

Drag The total drag acting on a body can be regarded as the sum of several components.

Total (profile, form) drag Skin-friction drag Pressure drag Nonlifting bodies

Profile (form) drag Skin-friction drag

Induced drag Pressure drag

In general, the drag is composed of a frictional component, related to viscous shearing in boundary layers, and pressure drag, which is related to the pressure differential between the fore and aft of the body. If the body also experiences lift, then there is an additional component of the pressure drag, called induced drag. The drag force is typically written in terms of the non-dimensional drag coefficient: CD = Drag force , 1 U 2 A 2

where is the density of the fluid, U is the velocity of the fluid and A is the reference area. The reference area is usually the projected area of the body in the direction of flow. For example, the reference area for a sphere would be r 2 , where 'r' is the radius. In the case of a circular cylinder, the area would be d.L, where L is the length. Sometimes, however, a different reference area is used. For example, in the

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

Lifting bodies

Total drag

case of an aerofoil, the reference area is usually the plan area, Length x Chord, regardless of the angle of attack. In the case of a thin flat plate oriented parallel to the flow, the entire drag is skin friction (boundary layer) drag. The skin friction coefficient obtained from, for example, the theory of Blasius, is in fact a drag coefficient. Drag force on a plate oriented normal to the flow In contrast to the case of a plate oriented parallel to the flow, where all the drag is due to skin friction, the drag on a plate oriented normal to the flow is all pressure drag.

Wake

In this case the drag is the result of the pressure difference between the upstream side of the plate and the downstream side. It is possible to predict the drag coefficient for this arrangement. The pressure on the upstream side of the plate will be approximately equal to the stagnation pressure of the flow, P0 = P + U2 /2, where P is the far field pressure. On the downstream side the pressure is approximately equal to P . In fact it will be less than this due to the acceleration of the fluid as it passes the edges of the plate. Hence, the difference is approximately U2 /2, and the drag coefficient is then approximately equal to 1. Experiment shows that the true value is between 1.2 and 1.5, depending on the aspect ratio of the plate. Values of the drag coefficient can be found tabulated in most text books. Finally, note that the drag coefficient for the case above is not dependent on the Reynolds number (compare with the friction coefficient on a plate oriented parallel to the flow). This is because the structure of the flow does not change as the Reynolds number is changed, provided that the Reynolds number is greater than some critical value (about 1000 in the above case). The boundary layer separates from the sharp edges of the plate, regardless of the Reynolds number, and so the size of the wake region does not change. This is typical of a body having abrupt edges, but is quite different to the case of flow past a body having gentler changes in geometry, such as a sphere or a cylinder. Here the drag coefficient is very dependent on the Reynolds number.

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

Drag coefficient on a circular cylinder In the case of a plate oriented parallel to the flow, the drag is all due to skin friction. If the plate is normal to the flow, the drag is all due to pressure. In the case of a cylinder, the drag is a combination of skin friction and pressure. The Reynolds number is defined as Re = Ud/, where d is the diameter and is the kinematic viscosity. Refer to the separate chart showing the variation of drag coefficient with Re. Re < 1

In this range the inertial effects of the flow are negligible. The flow is symmetrical fore and aft. The drag is dominated by boundary layer friction effects. It is possible to show that in this range: CD = 24/Re. 1< Re < 1000 Boundary layer separation

In this range the inertial effects begin to dominate. The flow becomes asymmetrical and the boundary layer separates from the rear face of the cylinder, producing a wake. The wake in this case consists of attached vortices. The pressure drag is now becoming more important than friction drag. The magnitude of the pressure drag depends on the extent of the rear of the body that is exposed to the wake. As Re is increased, the separation point moves forward over the surface of the cylinder, thereby increasing the size of the wake region and, as a result, the pressure drag. Since the structure of the flow changes with Re (the separation point moves), the drag coefficient is also a function of Re. Note that the dracoefficient, CD, is a reducing function of Re. However, this does not imply that the drag force itself is reducing. The drag force is indeed increasing with Re. The drag coefficient is the drag force divided by U2 /2. The observation that

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

the drag coefficient reduces as Re increases merely implies that the term U2 /2 increases faster than the drag force. 1000 < Re < 200,000 Boundary layer separation

In this range the separation point has moved to the top of the cylinder. As Re is increased further the separation point does not move. Hence, the structure of the flow does not change. This is reflected by an approximately constant drag coefficient. The wake is no longer attached to the cylinder. Instead it breaks away in the form of vortices shed first from one side of the cylinder and then from the other. 200,000 < Re < 500,000 In this range there is a sudden drop in the drag coefficient. This corresponds to the transition from laminar flow to turbulent flow in the boundary layer. When this happens the separation point moves back around behind the cylinder. The size of the wake region is now reduced and the drag force drops. This phenomenon is actively encouraged in Golf balls. The dimpled surface encourages early transition to turbulence in the boundary layer. A dimpled ball will fly about twice the distance that a smooth ball will fly. Drag coefficients for a variety of bodies can be found in most text books on fluid mechanics.

Streamlining It is obvious that pressure drag dominates the drag on a bluff body. Consequently, one might expect that by reducing the front profile of a body, and shaping the trailing surface to avoid boundary layer separation, the total drag may be reduced. This process is called streamlining.

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

t c
0.1 CD Friction 0 0 t/c 0.5 Total Pressure

The above graph illustrates the behaviour as the aspect ratio, t/c, is varied, for a given total volume. As t/c is increased, the body becomes more bluff and the pressure drag consequently rises. On the other hand, the frictional component is reducing. At some value of the aspect ratio, the total drag will be minimised. Streamlining is a common solution to the problem of drag minimisation employed by nature in the design of, for example, birds and fish. Lift The lift coefficient is defined in the same manner as the drag coefficient: CL = Lift force . 1 2 U A 2

For an aerofoil, both CD and C L depend to some extent on Re, but they are more dependent on the angle of attack. Separation at stall U Angle of attack, Increasing Separation point

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

During normal operation of the aerofoil, the boundary layer on the upper surface separates near the trailing edge. As the angle of attack is increased, the lift dramatically increases (as does the drag). The separation point creeps forward. At a particular angle of attack the separation point jumps forward to the leading edge of the foil. When this happens the lift on the foil suddenly drops. This is the stall condition. 2.0

1.6

0.020 0.016 CD 0.012

CL

1.2

0.8 0.008 0.4 0.004 0 0 4 8 12 16 20 Angle of attack

Mechanism of lift The body experiences lift because the average pressure on the lower surface is greater than the average pressure on the upper surface. The combination of the shape of an aerofoil and the angle of attack produces a flow of air over the top surface that has a higher average velocity than the air flow over the lower surface. Bernoulli's theorem then implies that the average pressure on the upper surface is less than that on the lower surface, hence a lift is produced.

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

PHigher velocity P+ Lower velocity Imagine for a moment that the flow over the top of the foil has the same speed as the flow over the bottom of the foil. This would have two consequences. Firstly, there would be no pressure differential between the top and bottom, so there would be no lift. Secondly, because the length of the path over the top surface is greater than over the bottom surface, the fluid will not travel as far along the top surface as it does along the bottom surface in the same time period, leading to the picture illustrated below. Equal velocity

Equal velocity This flow is obviously impossible since it implies infinite acceleration at the trailing edge to make the fluid swing around the sharp corner. In order to convert this hypothetical flow into the actual flow, we can add a rotating or circulating flow in the clockwise direction, as illustrated below.

Circulation

PHigher velocity P+ Lower velocity

The summation of the circulation and the original flow will produce the real flow and generate lift. The circulation is defined as = uds ,
s

where the integral is performed around any closed path, s, surrounding the body and u is the tangential velocity on this path.

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

The above arguments suggest that circulation must exist for the body to experience lift. In fact, we can quantitatively relate the lift to the circulation: Lift/unit span = U. Any body that exhibits a circulation will experience a lift force. A sphere that spins while it travels through the air, for example, will generate a circulating flow and hence will experience a lift force (a force normal to the direction of motion). This phenomenon is exploited by base-ball pitchers, who are able to make the ball follow a curved path through the air. Now imagine that a foil is moving through the air at a given angle of attack and at a constant speed. The above equation indicates that the lift felt by the foil corresponds to a given circulation, . What happens if we now accelerate the foil to a higher speed? We know that the lift increases, since Lift = CL U2 A/2. If the speed increases, then the circulation must also increase. However, if the circulation increase around the foil, then it must reduce somewhere else in the fluid body to maintain conservation of angular momentum in the whole body of fluid. This is achieved by shedding off the trailing edge of the foil a small vortex having a circulation exactly equal to the increase in circulation around the foil, but having the opposite sign. Thus, the circulation around the foil can increase while the total circulation remains unchanged. Such startup vortices have been observed experimentally. Startup vortices are shed whenever the lift on the foil changes, whether it be by changes in speed or angle of attack. +d

U+du

d Accelerate Startup vortex Wings of finite length In practice, all lifting bodies have finite length. Near the end of body, the pressure differential between the bottom and top surfaces produces a flow of air around the end, from beneath the body to the top of the body. This flow produces a vortex, the so called wing tip vortex, which then trails off into the wake.

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

Lower pressure above

Downwash Higher pressure below

Wing tip vortex The wing tip vortex and the corresponding downwash near the wing tip acts to reduce the lift on the wing and increase the drag. The additional drag is termed induced drag. The downwash reduces the effective angle of attack on the wing, which leads to the reduction in lift and increase in drag. Actual approach velocity d Downwash velocity Effective approach velocity Lift with downwash Lift without downwash d Reduced lift vector

Additional (induced) drag

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

The induced drag can be as much as 50% of the total drag. If we wish to minimise the induced drag, the wing shape should be long and thin. In this case, the disturbance at the end of the wing is less significant relative to the total drag. Gliders and other low powered aircraft are built with very long, thin wings for this reason. Why don't all aircraft have long, thin wings? Although this geometry minimises induced drag, it also makes the aircraft less manoeuvrable since the moment of inertia about an axis running along the fuselage is relatively high. For high manoeuvrability the aircraft should have short wings. The design of aircraft wings is therefore a compromise between the two effects. Finally, it is noteworthy that automobiles have a shape that is similar to a wing; an automobile is a lifting body. As a result, the forward motion of the vehicle produces lift and reduces the contact force between the road and the wheels. High speed racing vehicles are usually fitted with some form of foil to produce down-thrust in order to counteract this effect. In addition, the motor vehicle is a wing of very small aspect ratio. The induced drag therefore represents a substantial portion (around 50%) of the total drag. We can streamline the vehicle as much as possible to minimise the profile drag, but we cannot remove the induced drag.

ThermoFluids 313

Aerodynamic forces

M.B. Bush 1998

10

Das könnte Ihnen auch gefallen