Category Aircraft Flight

The longitudinal static stability of conventional aircraft

The movement of the centre of lift on a cambered wing has a destabilising effect. Figure 11.4 shows a cambered-section wing that is balanced or trimmed

Constant

moment

Aerodynamic centre

Fig. 11.2 Centre of pressure and aerodynamic centre

(a) The centre of lift or centre of pressure moves forward as the wing angle of attack increases.

(b) For an aerofoil, the situation shown in (a) above can be represented by a constant moment or couple (shown here in its true directional sense) and a lift force acting through a fixed point known as the aerodynamic centre.

N. B. The sign convention is that nose-up pitching is positive. A cambered aerofoil as shown above therefore gives a mathematically negative moment about the aerodynamic centre

at one angle of attack, by arranging the line of action of the lift to pass through the centre of gravity. If the angle of attack increases due to some upset, then the lift force will move forward, ahead of the centre of gravity, as shown. This will tend to make the front of the wing pitch upwards. The more it does so, the greater will be the upsetting moment. Such a wing is, therefore, not inherently stable on its own.

The condition for longitudinal static stability is that a positive (nose-up) change of angle of attack should produce a negative (nose-down) change in pitching moment.

where M is the pitching moment, and a is the pitch angle.

The conventional method for making an aircraft longitudinally stable, is to introduce a secondary surface, which is called a tailplane in the British con­vention, or more aptly, a horizontal stabiliser in American terminology.

To give some idea of how the tailplane works, we will consider two simple cases, one very stable, and the other highly unstable. Fig. 11.5(a) shows an aircraft trimmed for steady level flight. For simplicity we consider a case where the thrust and drag forces both pass through the centre of gravity, and thus produce no moment. We also ignore the forces and moments on the fuselage. The tail is initially producing a downward force, and hence, a nose-up pitching moment about the centre of gravity, whereas the wing lift produces a nose – down moment. The couple Mo is drawn nose up in Fig. 11.5, which is the normal mathematical convention.

If the aircraft is tipped nose-up by some disturbance as in Fig. 11.5(b), then the tail downforce and its moment will decrease, and the wing lift and its moment will increase. The moments are, therefore, no longer in balance, and

Fig. 11.5 Longitudinal static stability – a highly stable case

In this simplified example, the thrust and drag forces pass through the centre of gravity, and effects due to the fuselage are ignored.

(a) Aircraft trimmed in pitch LW x a – Mo = Lt x b

(b) If the angle of attack increases due to some upset, the tail-down force will decrease, and wing lift will increase. This will result in a nose-down pitching moment, tending to restore the aircraft to its original attitude.

N. B. Mo is shown here in the mathematically positive (rather than true) sense

there is a net nose-down moment, which will try to restore the aircraft to its original attitude. This aircraft is thus longitudinally statically stable.

Note, that in the simplified description above, we have ignored the inertia of the aircraft, and we have neglected the flexibility of the aircraft and its controls. Most importantly, we have ignored the effect of the fuselage which normally has a significant destabilising effect. It should further be noted that our simple example does not include any effects due to wing sweep. We should also have taken account of the fact that as the wing angle of attack and lift increases, so will the downwash at the tail. The increased downwash at the tail means that the tail downforce does not fall off as sharply with changing angle of attack as would otherwise have been expected. The restoring force is thus weakened.

Fig. 11.6 Longitudinally unstable arrangement – negative longitudinal dihedral necessitated by having centre of gravity much too far aft

Although the aircraft was initially trimmed, any increase in angle of attack will produce an unstable nose-up pitching moment.

(a) Aircraft trimmed with wing at 2°. Angle of attack and tail at 4°.

(b) Aircraft attitude increased by 2°. Wing now at 4° and tail at 6°. The wing lift will double, but the tail lift will only increase by 50%. The aircraft becomes untrimmed, with a nose-up pitching moment, and it will diverge from its initial attitude.

N. B. Just having the centre of gravity behind the wing aerodynamic centre does not make the aircraft unstable; it depends on how far aft the CG is

Wing downwash on the tail generally has a destabilising effect. The influence can be reduced by mounting the tail high relative to the wing.

Figure 11.6 shows a case of an aircraft which is trimmed, yet in a longitudin­ally unstable condition. The wing is initially at 2° angle of attack, and the tail is at 4° angle of attack. If we look at what happens when the angle of attack increases by 2°, due to a disturbance, then we see that the wing angle of attack will double. Since lift is directly proportional to angle of attack, it follows that the wing lift will double. In contrast, increasing the tail angle of attack by 2° from 4° to 6° represents only a 50% increase, with a corresponding 50% increase in tail lift (which is further reduced by the effects of downwash). The resulting forces will, therefore, produce a nose-up pitching moment, and the aircraft will continue to diverge from its original attitude.

The flare and touch-down

The final stages of the landing also offer the pilot a choice of techniques. Two alternatives are illustrated in Fig. 13.8. In the first the angle of attack is increased over a comparatively short period to arrest the descent, a manoeuvre known as the flare. The aircraft then flies parallel to the runway as the speed falls further and finally sinks onto the undercarriage. In the now less-common tail-wheel undercarriage, the final touch-down can either be on the main wheels only, or, with a greater amount of pilot skill, the aircraft can be brought to the angle of attack which results in all three wheels touching simultaneously, the so-called three-point landing.

Fig. 13.8 Alternative landing techniques

(a) Rapid ‘flare’ following straight ‘glide’ (b) Gradual ’round out’

(b) is easier than (a) but gives poorer obstacle clearance

An alternative method is to reduce the glide angle more progressively and to fly the aircraft along an almost circular path onto the runway. This type of approach is less demanding on the pilot, but results in slightly worse ability to clear obstacles near the threshold.

The generation of lift by a wing

In order to understand how the planform of the wing affects lift and drag, we need to look at the three-dimensional nature of the air flow near a wing.

You may remember, that we described in Chapter 1, how the wing pro­duced a circulatory effect; behaving like a vortex. A major breakthrough in the

The generation of lift by a wing

Fig. 2.1 Wing geometry understanding of aircraft aerodynamics came at the end of the nineteenth century, when the English engineer F. W. Lanchester reasoned that if a wing or lifting surface acts like a vortex, then it should possess all the general prop­erties of a vortex. Long before the Wright Brothers’ first flight, a theory of vortex behaviour had been developed which indicated that a vortex could only persist if it either terminated in a wall at each end, or formed a closed ring like a smoke ring. In Fig. 2.3 we show in very simplified form, how this requirement of forming a closed circuit is met. In the diagram we see that the circulatory effect of the wing, which is known as the wing-bound vortex, turns at its ends to form a pair of real vortices, trailing from near the wing tips. The ring is com­pleted by a so-called starting vortex downstream.

These vortices do exist in reality and we can easily detect the trailing vortices in a wind tunnel by using a wool tuft which will rotate rapidly if placed in the appropriate position behind a model. On a real aircraft they can sometimes be seen as fine lines of vapour streaming from near the wing tips, as seen in Fig. 2.4. This often occurs at airshows, particularly on damp days. They are most likely to be seen when an aircraft is pulling out of a dive, and is therefore

The generation of lift by a wing

Fig. 2.2 High aspect ratio on the powered glider version of the Europa (lowest aircraft)

(Photo courtesy of Europa Aircraft Ltd)

The generation of lift by a wing

Fig. 2.3 Simplified wing vortex system

The generation of lift by a wing

Fig. 2.4 Trailing vortices originating at the wing tips of the late-lamented TSR-2, made visible by atmospheric vapour condensation (Photo courtesy of British Aerospace)

generating a large amount of lift, so that the wing circulation and trailing vortices are strong.

In the wing trailing vortex, as in a whirlwind or whirlpool, the speed of the rotating fluid decreases with distance from the central core. From the Bernoulli relationship, we can see that, since the air speed in the centre of the vortex is high, the pressure is low. The low pressure at the centre is accompanied by a low temperature, and any water vapour in the air tends to condense and become visible in the centre of the trailing vortex lines, as in Fig. 2.4. Note that the vapour trails frequently seen behind high-flying aircraft are normally formed by condensation of the water vapour from the engine exhausts, and not from the trailing vortices. Figure 2.5 is a flow-visualisation picture showing the trailing vortices forming at the wing tips.

Boundary layer scale effect – model testing

In Fig. 3.18 we show two thin almost flat wing sections, a full-size one and a scale model, placed at zero angle of attack in a stream of air. In this situation, the position of transition from laminar to turbulent flow will be roughly the same distance from the leading edge in both cases, as illustrated.

From the diagram, you will see that the scale model will therefore have a greater proportion of laminar boundary layer, and consequently a lower drag per unit of area than for the larger one. So the drag per unit area measured on the model is not representative of full scale.

To correct for the effect of scale, the model could be placed in a stream of air moving faster than that for the larger section. This would increase the Reynolds number, and move the position of transition forward. If the speed were sufficiently high, transition could be moved to a position corresponding to that of the full-scale section.

Boundary layer scale effect - model testing

MODEL

Fig. 3.18 On a thin flat plate at zero angle of attack, the transition position would be at roughly the same distance from the leading edge for both model and full-size plates. The model would therefore have a higher proportion of laminar boundary layer

The same principle applies to all shapes, and to obtain similar flow patterns between model and full scale, it is necessary to ensure that the Reynolds num­ber in the model test is the same as for the full-size aircraft in flight.

The Wright brothers and other early experimenters were either unaware of this fact, or did not bother about it. Their simple wind-tunnel tests conducted on very small models at low speeds indicated that thin plate-like wings gave a better ratio of lift to drag than ones with a thicker aerofoil type of section. Thus, early aircraft had thin plate-like wings. It was Prandtl who spotted the error, and found that when the Reynolds number of the tests was increased by running the tunnel faster, or using larger models, thicker wing sections pro­duced a better lift-to-drag ratio than curved or flat plates.

The reason for the poor performance of thick aerofoil sections at very low Reynolds numbers (small models at low speeds), is that the flow will be laminar over most of the surface and thus will separate very easily. A thin plate with a sharp leading edge generates turbulence at the leading edge, and the resulting turbulent boundary layer is better able to stay attached. Model aircraft often perform better when equipped with means of turbulating the boundary layer, and require quite different wing section shapes from full-size aircraft, as described by Simons (1999).

Hypersonic flow

We have examined the way in which the flow changes as the speed of sound is exceeded and in Chapter 8 we will look at the way aircraft have developed to operate at speeds up to about twice the speed of sound (Mach 2). Some aircraft, however, have to operate at very much higher Mach numbers, particularly re-entering satellites and space shuttles. We find that a number of problems are associated with flight at these very high Mach numbers (up to about 27). Some of these are direct aerodynamic problems associated with the extreme speeds. Some are primarily structural and material problems caused by the high temperature induced by the flow. The SR-71 spy-plane used expansion joints in the structure, which lead to conspicuous, but not dangerous, leakage of fuel when on the ground. Other problems are due to the height at which such flight conditions are most likely to be made and are caused by the very low density encountered. For realistic flight conditions these problems begin to be felt at Mach numbers above about six, and so flight above this rather imprecise demarcation line is known as hypersonic.

What, then, are the aerodynamic problems associated with this flight regime? Initially nothing particularly dramatic occurs. All the main features of the supersonic flow are there, such as the bow shock wave and expansions. As would be expected from the increase in Mach number, the bow shock wave is more acutely swept to the free-stream direction. It is when we come to look at the details of the flow that we find the important changes that have taken place.

We have already seen that the shock wave is quite a traumatic experience for the flow passing through it. Pressure, density and temperature all increase dramatically over a very short distance. However, the basic composition of the air passing through the wave does not change in supersonic flow. It still consists of a mixture of roughly 70 per cent nitrogen, 20 per cent oxygen, 9 per cent carbon dioxide, with a few rare gases thrown in for good measure. The molecules of each of the various constituents are in their usual form with the nitrogen and oxygen molecules both being diatomic (i. e. with two atoms to each molecule). All the constituents are also electrically neutral, the electrons in each molecule exactly balancing the charge in the molecular nucleus.

As the Mach number increases and the shock wave gets stronger this situ­ation changes and the so-called real gas effects become important. The relatively simple relationships between gas properties which occur under moderate con­ditions of temperature and pressure break down. The two atoms in the gas molecules become detached from each other, a process known as dissociation and energy is released into the flow. This dissociation may also be present in the high temperature regions of the boundary layer near the surface of the vehicle.

A further problem occurs due to the fact that molecules may become elec­trically charged, or ionised. This means that electrical forces may further com­plicate the fluid motion. This may not necessarily be a bad thing, and schemes have been suggested to use this feature to control the flow or even provide a propulsion system.

Yet another complication arises when we consider flight at extreme altitude. For normal aircraft operation, the air molecules are very close together. The average distance between molecular impacts (the mean free path at sea level) is about 6.6 x 10-5 mm. At 120 km altitude, this distance increases to 7 m, a distance that is quite large when compared with the size of the vehicle travel­ling through the air. In this case we can no longer think of the air as being a continuous fluid, but must consider the action of individual molecules, and average their effect.

From the above, it will be appreciated that the theoretical prediction of such flows becomes very difficult. Experimental work under such extreme condi­tions is also an arduous and costly undertaking. For further information, the interested reader is referred to Cox and Crabtree (1965).

Thrust reversal

An early problem with turbo-jet propulsion was that the high speed aircraft that it produced tended to have a high landing speed. As there was no propeller drag to help slow them down, they needed very long runways. One solution to this problem is to fit thrust reversers in the form of a movable device to deflect the exhaust jet forwards. Thrust reversers take many forms, and may either use cold air from the compressor, or the hot exhaust gases. Two typical hot-jet

Thrust reversal

Fig. 6.31 Hot-jet thrust reversers

(a) Bucket type (b) Clam-shell type

deflector designs are shown in Fig. 6.31. The louvred cascade for the hot-jet deflector used on Concorde may be seen in Fig. 6.32.

Despite the added cost and weight penalty, thrust reversers are now popular even on small executive jets. Apart from reducing the landing run, the feature enables the aircraft to manoeuvre more easily on the ground under its own power.

A more detailed description of jet engine components is given in the well – illustrated Rolls-Royce publication The Jet Engine (1986).

The centre section

Any real swept wing with subsonic leading edges will not behave in quite the same way as it would at subsonic speed because of the fact that there must be a limit to its span. For simplicity we will first examine the simple case of a swept wing without a fuselage separating the two halves (Fig. 8.11).

In this case the flow can only be influenced at a finite distance upstream and for very thin wings at small angles of incidence the appropriate zone of influence will be approximately defined by the Mach waves at the apex of the wing (Fig. 8.11(a)). If the disturbance to the flow is bigger, because of increased wing thickness or angle of attack, then a shock wave forms at the apex (Fig. 8.11(b)) and, because of its higher propagation speed, the zone of influence of the wing will be extended slightly in the upstream direction.

We now see that for a real swept wing we will still generate wave drag due to this bow shock wave. However we have gained one important advantage and that is due to the fact that the wing now works in much the same way at both sub – and supersonic speeds over most of the span. This means that the problem of choosing a wing section which is a suitable compromise between high and low speed has been made very much easier than before.

Fig. 8.11 Influence of centre section of swept wing

For thicker wing the bow shock wave is less swept than the Mach cone (a) Thin wing at low angle of attack (b) Increased thickness or angle of attack

If we now introduce a fuselage the overall picture does not look very differ­ent, although some clever aerodynamic design at the wing fuselage junction may well prove very worthwhile – but more of that later.

The tip region

Unlike the unswept wing which we discussed earlier in this chapter, the tip region lies within the Mach cone of all the upstream wing sections. This region thus again behaves in a manner very similar to its subsonic counterpart (Chapter 2) and a trailing vortex sheet will be generated along the trailing edge of the wing and this will roll up into two large vortices which stream down­stream of the wing in a position close to the tips.

The vee-tail

A final variant of tailplane design is the vee-tail, illustrated in Fig. 10.9. The hinged trailing-edge control surfaces are moved differentially (one up and the other down) to provide a sideforce component like a rudder, and collectively (both up or down together) to provide a vertical component like a conventional tail. The claimed advantage of the vee-tail is that the number of surfaces is reduced, with consequential reductions in drag and weight. The F117A stealth fighter Fig. 6.34 uses a vee-tail which avoids the right-angles that cause large radar reflections.

Fig. 10.9 Vee-tail on this unusual light jet-propelled Marmande Microjet

Deliberately unstable aircraft

Flying an aircraft in a neutrally or slightly unstable condition is not necessarily difficult or dangerous, but it involves hard work for the pilot, who cannot take his hands off the controls, and must make continuous control adjustments. The Wright brothers’ original aircraft was unstable, which made it more responsive and controllable than many of its contemporary rivals. There are, however, other more important advantages in moving the centre of gravity aft. By mov­ing it to the neutral point, the position where the aircraft is neutrally stable, the tail has to produce no trimming force, and hence there is no trim drag. By moving it even further aft to an unstable position (negative CG margin) we can arrive at a position where the wing and the tail are both producing lift at an efficient positive angle of attack. This considerably improves the lift-to-drag ratio of the aircraft, and can dramatically improve its performance. An un­stable aircraft will also respond more quickly to control inputs, making it highly manoeuvrable.

Aviation safety regulations traditionally took a dim view of flying in an unstable condition, but for military applications, the performance advantages are considerable. With the development of increasingly reliable electronic control systems it became practical to build aircraft that could be flown in a naturally unstable condition, relying entirely on automatic systems to maintain artificial stability. Most high performance military aircraft are in any case

totally unflyable in the event of a major electrical failure, so further depend­ence on electrical systems does not significantly reduce their safety. The X-29 (Fig. 9.20), and Typhoon (Fig. 10.8) are both designed to be inherently un­stable at subsonic speeds. For civil aircraft, some reduction in stability may be tolerated, if the overall system can be shown to be capable of coping safely with failures in individual elements. This normally entails duplicate or multiple components, and rapid automatic fault diagnosis.

The influence of structural materials

The introduction of new materials has opened up a range of possibilities for the design of more efficient aircraft, and even new types of aircraft. Man-powered flight would probably not have been possible using traditional materials.

Since the First World War, aluminium alloys have been almost universally used as the primary structural material, even for supersonic aircraft capable of Mach numbers up to about 2.2, such as Concorde. However, for sustained flight at Mach numbers above about 2.5, the effects of kinetic heating render conventional aluminium alloys unsuitable. Instead titanium and steel alloys may be employed. Unfortunately, their use presents something of an economic barrier. Apart from the higher cost of these materials, the fabrication tech­niques required tend to be more expensive. It is this economic barrier, rather than any purely aerodynamic problem, that has limited the maximum Mach number to around 2.5 for all but a handful of experimental aircraft. Rare exceptions are the MiG-25 combat aircraft, which is capable of Mach 3, and the even faster specialised Lockheed SR-71 reconnaissance aircraft.

Since the 1950s, gradually increasing use has been made of fibre reinforced materials. Originally, glass fibres were used, but a major advance came with the introduction of carbon (graphite) fibres. Carbon fibres can be produced in a number of forms, and can be optimised either for high strength, or for high stiffness (high modulus). It is the high stiffness of carbon fibres that make them a particularly attractive alternative to metals in aircraft construction. Boron fibres show even better properties, but are less cost effective than carbon fibres, and have only been used in experimental or highly specialised applications.

Although fibre reinforced or composite materials can have a higher strength – to-weight or stiffness-to-weight ratio than metals, they cannot simply be used as a direct replacement. The main problem is that they do not deform plastically like metals, and cannot be joined by conventional types of bolts or rivets, since this causes local cracking. The general adoption of fibre reinforced materials

Fig. 14.6 Progress in the use of advanced structural materials

The Airbus A350XWB makes considerable use of composite materials for its primary structure. This saves considerable weight, which helps to improve fuel efficiency

(Photo courtesy of Airbus)

was, therefore, slowed down by the need to develop suitable fastenings and construction techniques. Increasing use of composites is now being made, particularly in military combat aircraft and helicopters. The Beech Starship (Fig. 4.10) was one of the first civil transport aircraft designed for large-scale production, to use composites for its primary structure.

In addition to high strength and stiffness, fibre reinforced materials have some other important special properties. By aligning the fibres in particular patterns within a structure, it is possible to control the relationship between bending and torsional stiffness. This technique is one of the methods that can be used to reduce the tendency to structural divergence of forward-swept wings, and gives us another example of the way in which the development of materials can influence aerodynamic design judgements.

The use of moulded composite structures has also made it economically practical to produce complex aerodynamically optimised shapes, even for light aircraft.

Because fibre reinforced materials are built-up, rather than being cut or bent out of solid block and sheet, they can be produced in much more complex, ‘organic’ forms, with continuous variations in thickness, curvature and stiff­ness. Such structures begin to resemble the highly efficient optimised shapes found in the bones of birds.

Composite structures were initially restricted to smaller components such as control surfaces, but more recent aircraft employ composite materials for the main structural components, as on the Airbus A350XWB shown in Fig. 14.6, and also the A400 heavy lifter, and the Boeing Dreamliner. The weight saving allows for significant improvement in fuel consumption, or enhanced payload capacity.

Further discussion of aircraft structural design is beyond the intended scope of this book, but Megson (2007) gives a good introduction.

Recommended further reading

Megson, T. H. G., Aircraft structures for engineering students, 4th edn, Butterworth – Heinemann, 2007, ISBN 9780750667395. A well-respected standard undergraduate textbook. Includes examples. Solutions manual available.

Stinton, Darrol, The anatomy of the aeroplane, 2nd edn, Blackwell Science, Oxford, 1998, ISBN 0632040297. A classic introduction to aircraft design.

Wilkinson, R., Aircraft structures and systems, 2nd edn, Mechaero Publishing, St Albans, UK, 2001, ISBN 095407341X. A good easily read introductory text with a non-mathematical approach.

Conclusion

This concludes our introduction to the subject of aircraft flight. We have tried to include all of the important basic principles, and one or two items of inter­est. Inevitably we will have omitted something important, but the references given in this book should lead you to most of the missing information.