Você está na página 1de 476

081 PRINCIPLES OF FLIGHT

G LONGHURST 1999 All Rights Reserved Worldwide

COPYRIGHT
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the prior permission of the author. This publication shall not, by way of trade or otherwise, be lent, resold, hired out or otherwise circulated without the author's prior consent. Produced and Published by the PROFESSIONAL PILOT STUDY CENTRE EDITION 1.01.00 1999 This is the first edition of this manual, and incorporates all amendments to previous editions, in whatever form they were issued, prior to July 1999. EDITION 1.01.00 COPYRIGHT 1999 G LONGHURST

The information contained in this publication is for instructional use only. Every effort has been made to ensure the validity and accuracy of the material contained herein, however no responsibility is accepted for errors or discrepancies. The texts are subject to frequent changes which are beyond our control.

G LONGHURST 1999 All Rights Reserved Worldwide

Online Documentation Help Pages

Help TO NAVIGATE THROUGH THIS MANUAL


When navigating through the manual the default style of cursor will be the hand symbol. This version of the CD-Online manual also supports a mouse incorporating a wheel/ navigation feature. When the hand tool is moved over a link on the screen it changes to a hand with a pointing finger. Clicking on this link will perform a pre-defined action such as jumping to a different position within the file or to a different document. Navigation through a manual can be done in the following ways:

G LONGHURST 1999 All Rights Reserved Worldwide

Online Documentation Help Pages

Help
The INDEX button takes you to the Index of the manual you are in, if it is available. The EMAIL button takes you to the PPSC web site, where you can e-mail the instructors with your questions, provided you have an internet connection. The PAGE button takes you to the previous and next pages in the book. The FIND button allows you to search for specific words within the manual The arrows are used to display the previous and next words whilst using the search tool.

The CONTENTS button takes you to the first page of the main Table Of Contents.

The WEB button takes you to the Click2PPSC web site.

The BACK button returns you to your previous position in the document.

The HELP button takes you to the help pages.

The EXIT takes you to the main menu.

G LONGHURST 1999 All Rights Reserved Worldwide

TABLE OF CONTENTS
Aerodynamic Principles Lift Drag Stalling Lift Augmentation Control Forces in Flight Stability High Speed Flight Limitations

G LONGHURST 1999 All Rights Reserved Worldwide

TABLE OF CONTENTS
Special Circumstances Propellers

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Aerodynamic Principles
Units Systems of Units Newton's Laws of Motion The Equation of Impulse Basic Gas Laws Airspeed Measurement Shape of an Aerofoil The Equation of Continuity Bernoullis Theorem

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

Aerodynamic Principles

Units
1. In order to define the magnitude of a particular body in terms of mass, length, time, acceleration etc., it is necessary to measure it against a system of arbitrary units. For example, one pound (lb) is a unit of mass, so the mass of a particular body may be described as being a multiple (say 10 lb), or sub-multiple (say lb) of this unit. Alternatively the mass of the body could have been measured in kilograms, since the kilogram (kg) is another arbitrary unit of mass.

Systems of Units
2. There are a number of systems of units in existence and it is essential when making calculations to maintain consistency by using only one system. Three well-known consistent systems of units are the British, the c.g.s. and the S.I. (Systeme Internationale). These are illustrated in Figure 1-1 below:

Chapter 1 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
FIGURE 1-1 Units of Measurement

BRITISH SYSTEM LENGTH TIME ACCELERATION MASS FORCE Foot Second Ft/s Pound Poundal

C.G.S. Centimetre Second C/s Gram Dyne

S. I. Metre (m) Second (s) M/s Kilogram (kg) Newton (N)

3. The S.I. system of units is the one most commonly used. In this system, one Newton is the force that produces an acceleration of 1 M/s when acting upon a mass of 1 kg.

Newton's Laws of Motion


4. The motion of bodies is usually quite complicated, involving several forces acting at the same time as well as inertia and momentum. Before considering the Laws of Motion, as described by Sir Isaac Newton, it is necessary to define force, inertia and momentum.

Chapter 1 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
5. Force is that which changes a body's state of rest or of uniform motion in a straight line. The most familiar forces are those which push or pull. These may or may not produce a change of motion, depending upon what other forces are present. Pressure acting upon the surface area of a piston exerts a force that causes the piston to move along its cylinder. If we push against the wall of a building a force is exerted but the wall does not move, this is because an equal and opposite force is exerted by the wall. Similarly, if a weight of 1 kilogram is resting upon a table there is a force (gravitational pull) acting upon the weight but, because an equal and opposite force is exerted by the table, there is no resultant motion.

Force Can Be Quantified


6. Where motion results from an applied force, the force exerted is the product of mass and acceleration, or: F= ma Where: F = Force m = mass and a = acceleration 7. Inertia is the tendency of a body to remain at rest or, if moving, to continue its motion in a straight line. Newton's first law of motion, often referred to as the law of inertia, states that every body remains in a state of rest or uniform motion in a straight line unless it is compelled to change that state by an applied force.

Chapter 1 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles Momentum


8. The product of mass and velocity is called momentum. Momentum is a vector quantity, in other words it involves motion, with direction being that of the velocity. The unit of momentum has no name, it is given in kilogram metres per second (kg m/s). Newton's second law of motion states that the rate of change of momentum of a body is proportional to the applied force and takes place in the direction in which the force acts. 9. Newton's third law of motion states that to every action there is an equal and opposite reaction. This describes the situation when a weight is resting upon a table. For a freely falling body the force of gravity (gravitational pull), measured in Newtons , acting upon it is governed by: F = mg where g is acceleration due to gravity 9.81M/s, and m is the mass of the body in kilograms. 10. If the same body is at rest upon a table it follows that, since there is no motion, there must be an equal and opposite force exerted by the table.

Motion with Constant Acceleration


11. When acceleration is uniform, that is to say velocity is increasing at a constant rate, the relationship between acceleration and velocity can be expressed by simple formulae known as the equations of motion with constant acceleration. Under these circumstances velocity increases by the same number of units each second, so the increase of velocity is the product of acceleration (a) and time (t). If the velocity at the beginning of the time interval, (the initial velocity), is given the symbol (u) and the velocity at the end of the time interval, (the final velocity), is given the symbol (v) then the velocity increase for a given period of time can be expressed by the equation:

Chapter 1 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
v = u + a.t 12. If it is required to calculate the distance travelled (s) during a period of motion with constant acceleration, this can be done using the equation: 1 -(u + v )t s = -2 13. By substitution, using the above two equations, it is possible to develop two more equations: 1 2 s = ut + -- at 2 And: v 2 = u 2 + 2as These are the equations of motion with constant acceleration.

The Equation of Impulse


14. Given that the momentum of a body is the product of its mass and its velocity it follows that, providing mass and velocity remain constant, momentum will remain constant. A change of velocity will occur if a force acts upon the body because: F = ma

Chapter 1 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
And therefore F a = --m 15. If the force acts in the direction of motion of the body for a period of time (t), the resultant acceleration will cause a velocity increase from (u) to (v). This must also cause an increase in momentum from (mu) to (mv). Combining the equations F = ma and v = u+at gives: F v = u + t --m Which transposes to: Ft = mv mu 16. The change in momentum (final momentum minus initial momentum) due to a force acting on a body is the product of that force and the time for which it acts. This change in momentum called the impulse of the force and is usually identified by the symbol J. Hence: J = Ft Or: J = mv mu

Chapter 1 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
17. This is the equation of impulse. The S.I. unit of impulse, being the product of force and time, is the Newton second (Ns). NOT, it should be noted, Newton per second (N/s).

Basic Gas Laws


18. The Gas Laws deal with the relationships between pressure, volume and temperature of a gas. They are based upon three separate experiments carried out at widely differing times in history. These experiments investigated: (a) (b) (c) The relation between volume (V) and pressure (P) at constant temperature (Boyle's Law). The relation between volume (V) and temperature (T) at constant pressure (Charles' Law). The relation between pressure (P) and temperature (T) at constant volume (Pressure Law)

Boyle's Law
19. Boyle's Law states that the volume of a fixed mass of gas is inversely proportional to the pressure, provided that the temperature remains constant. In other words, if the volume of a given mass of gas is halved its pressure will be doubled or, if its pressure is halved its volume will be doubled, providing its temperature does not change. 20. This may be expressed mathematically as: P1 V1 = P2 V2 or PV = cons tan t

Chapter 1 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

Charles' Law
21. Charles Law states that the volume of a fixed mass of gas at constant pressure expands by 1/ 273 of its volume at 0C for every 1C rise in temperature. In other words, the volume of a given mass of gas is directly proportional to its (absolute) temperature, providing its pressure does not change. 22. This may be expressed mathematically as: V2 V1 ------ = -----T1 T2 V --- = cons tan t T

or

Pressure Law
23. The pressure law is the result of experimentation during the nineteenth century by a professor called Jolly and states that the pressure of a fixed mass of gas at constant volume increases by 1/273 of its pressure at 0C for every 1C rise in temperature. In other words, the pressure of a given mass of gas is directly proportional to its temperature, providing its volume does not change. 24. This may be expressed mathematically as: P2 P1 ----- = ----T1 T2 P -- = cons tan t T

or

Chapter 1 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

The Ideal Gas Equation


25. The three equations expressing the Gas Laws can be combined into a single or Ideal Gas Equation which may be expressed mathematically as: P2 V2 P1 V1 ------------ = -----------T1 T2 or PV ------- = cons tan t T

Static Pressure
26. The static pressure of the atmosphere at any given altitude is the pressure resulting from the mass of an imaginary column of air above that altitude. In the International Standard Atmosphere (ISA) at mean sea level the static pressure of the atmospheric air is 1013.25 millibars (mb), which equates to 14.7 pounds per square inch (psi) or 29.92 inches of mercury (in. Hg). ISA mean sea level conditions also assume an air density of 1.225 kilograms per cubic metre (kg/m) and a temperature of +15C (288A). The standard notation for static pressure at any altitude is (P).

Dynamic Pressure
27. Air has density (mass per unit volume) and consequently air in motion has energy and must exert pressure upon a body in its path. Similarly, a body moving in air will have a pressure exerted upon it that is proportional to its rate of movement, or velocity (V). This pressure due to motion is known as dynamic pressure and is given the notation (q). 28. Energy due to motion is kinetic energy (K.E.) and in the S.I. system of units is measured in joules (j). From Bernoullis equation for incompressible flow the kinetic energy due to air movement may be calculated using the formula:

Chapter 1 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
1 2 KE = --V 2 29. To calculate kinetic energy in joules, density () must be in kilograms per cubic metre (kg/m) and velocity (V) in metres per second (m/s). One joule is the work done when a force of 1 newton moves the point of application of the force 1 metre in the direction of the force. 30. If a volume of moving air is brought to rest, as in an open-ended tube facing into the airstream, the kinetic energy is converted into pressure energy with negligible losses. Hence, dynamic pressure: 1 2 - V q = -2 31. It should be noted that dynamic pressure cannot be measured in isolation, since ambient atmospheric pressure (static pressure) is always present also. The sum of the two, (q+P), is known variously as total pressure, stagnation pressure or pitot pressure and is given the notation (H or Ps). Therefore, dynamic pressure: q = (q + P ) P

Chapter 1 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

Viscosity
32. Viscosity is a measure of the internal friction of a liquid or gas and determines its fluidity, or ability to flow. The more viscous a fluid, the less readily it will flow. Unlike liquids, which become less viscous with increasing temperature, air becomes more viscous as its temperature is increased. The viscosity of air is of significance when considering scale effects in wind tunnel experiments and in terms of friction effects as it flows over a surface. Changes of density do not affect the air viscosity.

Density
33. Density () is defined as mass per unit volume. The density of air varies inversely with temperature and directly with pressure. When air is compressed, a greater mass can occupy a given volume or the same mass can be contained in a smaller volume. Its mass per unit volume has increased so, by definition, its density has increased. 34. When the temperature of a given mass of air is increased it will expand, thus occupying a greater volume. Assuming that the pressure remains constant the density will decrease because the mass per unit volume has decreased. 35. Both the above statements assume that the air is perfectly dry. When air is humid, that is it contains a proportion of water vapour, it becomes less dense. This is because water vapour weighs less than air and so a given volume of air weighs less if it contains water vapour than if it were dry. Its mass per unit volume is less.

Chapter 1 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

Airspeed Measurement
Indicated Airspeed (IAS)
36. The speed displayed on the airspeed indicator (ASI) is known as indicated airspeed. It does not include corrections for instrument errors and static pressure measurement errors (pressure error), both of which are very small. The indicated airspeed will differ progressively from actual flight speed as altitude increases and, consequently, density () decreases (q = V). The notation for IAS is (VI).

Calibrated Airspeed (CAS)


37. Also known as Rectified Airspeed (RAS), this is the speed obtained by applying the appropriate instrument error and pressure error corrections to the ASI reading. The notation for CAS is (Vc).

Equivalent Airspeed (EAS)


38. The equation for IAS (dynamic pressure) is derived from Bernoullis equation, which assumes air to be incompressible. Below about 300 knots the compression that occurs when the airflow is brought to rest (as in the pitot tube) is negligible for most practical purposes, becoming increasingly significant above that speed. EAS is obtained by applying the compressibility correction to CAS. The notation for EAS is (Ve).

Chapter 1 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

True Airspeed (TAS)


39. The true airspeed is the actual flight speed relative to the surrounding atmosphere, regardless of altitude. It must, therefore, take account of air density and is obtained by applying the formula: EAS TAS = ---------- where = relative air density = ----0 40. TAS is given the notation (V). At 40,000 ft, where standard density is one-quarter sea level density, TAS will be twice EAS (0.25 = 0.5). British ASIs, in common with most others, are calibrated for ISA mean sea level density (0), where EAS = TAS. At all greater altitudes TAS will be greater than EAS by a proportional amount.

Shape of an Aerofoil
41. The terminology for the dimensions that determine the shape of an aerofoil section is shown in Figure 1-2 below.

Chapter 1 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
FIGURE 1-2 Aerofoil Section

Chord Line
42. A straight line joining the leading edge to the trailing edge of the aerofoil.

Chord (c)
43. The distance between leading and trailing edge measured along the chord line.

Thickness/Chord Ratio
44. The maximum thickness of the aerofoil section, expressed as a percentage of chord length. A typical figure is about 12 per cent. The distance of the point of maximum thickness from the leading edge, on the chord line, may also be given as a percentage of chord length. Typically it is about 30 percent.

Chapter 1 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

Mean Camber Line


45. A line joining the leading and trailing edges which is equidistant form the upper and lower surfaces along its entire length.

Camber
46. The displacement of the mean camber line from the chord line. The point of maximum camber is expressed as a percentage and is the ratio of the maximum distance between mean camber line and chord line to chord length. The amount of camber and its distribution along the chord depends largely upon the operating requirements of the aircraft. Generally speaking, the higher the operating speed of the aircraft the less the camber (i.e. the thinner the wing).

Nose Radius
47. The nose or leading edge radius is the radius of a circle joining the upper and lower surface curvatures and centred on a line tangential to the curve of the leading edge.

Angle of Attack ()
48. The angle between the chord line and the relative airflow (RAF). This may also be referred to as incidence, but must not be confused with the angle of incidence. Furthermore, it is essential to differentiate between the angle of attack and pitch angle, or attitude, of the aircraft. The latter is, of course, measured relative to the horizontal plane.

Chapter 1 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

Angle of Incidence
49. The angle between the aircraft wing chord line and the longitudinal centreline of the aircraft fuselage.

The Wing Shape


50. The shape of an aircraft wing in planform has a great influence on its aerodynamic characteristics and will be discussed in depth in later chapters. The terminology describing the dimensions that determine wing shape is listed below.

Wing Span
51. The straight-line distance measured from tip to tip. See Figure 1-3.

Chapter 1 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
FIGURE 1-3 Wing Span

Wing Area
52. The plan surface area of the wing. In a wing of rectangular planform it is the product of span x chord.

Chapter 1 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

Aspect Ratio
53. The ratio of wing span to mean chord or span to wing area.
2

Wind Loading
54. The weight per unit wing area.

Root Chord
55. The chord length at the centreline of the wing (the mid-point along the span).

Tip Chord
56. The chord length at the wing tip.

Tapered Wing
57. A wing in which the root chord is greater than the tip chord.

Taper Ratio
58. The ratio of tip chord to root chord usually expressed as a percentage.

Quarter Chord Line


59. A line joining the points of quarter chord along the length of the wing.

Chapter 1 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles

Swept Wing
60. A wing in which the quarter chord line is not parallel with the lateral axis of the aircraft. See Figure 1-4.

Sweep Angle
61. The angle between the quarter chord line and the lateral axis of the aircraft. See Figure 1-4.

Chapter 1 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
FIGURE 1-4 Sweep Angle

Mean Aerodynamic Chord


62. The chord line passing through the geometric centre of the plan area of the wing (ie. the centroid). See Figure 1-5.

Chapter 1 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
FIGURE 1-5 Mean Aerodynamic Chord

Dihedral
63. The upward inclination of the wing to the plane through the lateral axis. See Figure 1-6.

Chapter 1 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
FIGURE 1-6 Dihedral

Anhedral
64. The downward inclination of the wing to the plane through the lateral axis. See Figure 1-7.

FIGURE 1-7 Anhedral

The Equation of Continuity


65. The equation of continuity states that mass cannot be either created or destroyed. Air mass flow is a constant.

Chapter 1 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
66. Figure 1-8 illustrates the streamline flow of air through a cylinder of uniform diameter. The air mass flow is the product of the density of the air (), the cross-sectional area of the cylinder (A) and the flow velocity (V). At any point along the cylinder: Airmass flow = AV = cons tan t

FIGURE 1-8 Streamline Flow

67. Mass flow = AV = constant is the general equation of continuity, which applies to both compressible and incompressible fluids. In compressible flow theory it is convenient to assume that changes in density can be ignored at speeds below about 0.4 Mach and a simplified equation of continuity: AV = cons tan t

Chapter 1 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
68. Consider now the streamline airflow through a venturi tube as illustrated at Figure 1-9. Given that mass flow is constant at any point and is the product of AV then at point Y, since the cross-sectional area (A) is reduced, velocity (V) must increase in order to maintain the equation of continuity.

FIGURE 1-9

69. In other words, a reduction in cross-sectional area (as in a venturi tube) produces an increase in velocity and vice-versa.

Bernoullis Theorem
70. A gas in motion possesses four types of energy. Potential energy (due to height), heat energy, pressure energy and kinetic energy (due to motion). Bernoulli demonstrated that in an ideal gas in steady streamline flow the sum of the energies remains constant. At low subsonic (less than 0.4 mach) flow air can be conveniently regarded as an ideal gas (incompressible and inviscid). In these circumstances Bernoulli's Theorem can be further simplified by assuming there is no transfer of heat or work in or out of the gas and by ignoring the insignificant changes in potential energy and heat energy. 71. For practical purposes then, in streamline flow of air around an aircraft wing at low subsonic speed:

Chapter 1 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Aerodynamic Principles
pressure energy + kinetic energy = constant 72. This can be expressed as: 1 2 P + -- V = cons tan t 2 73. Where P = static pressure. In other words, static pressure + dynamic pressur = constant.

74. From this simplified Bernoulli's Theorem it is evident that an increase in velocity of gas flow results in a decrease in static pressure, and vice versa. Hence at point Y in Figure 1-9 the increase in velocity of airflow will produce a decrease in pressure.

Chapter 1 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Lift
Airflow Round an Aerofoil Two-dimensional Flow Three Dimensional Flow Wake Turbulence Lift

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Lift

Airflow Round an Aerofoil


1. As stated in the previous chapter the relationship between pressure and velocity in the airflow patterns around an object is that defined by Bernoulli. The airflow impacting on the object at a point near its leading edge will be brought to rest, or stagnate. At the stagnation point the velocity is zero and the pressure equal to the total pressure of the air stream, that is to say ambient atmospheric pressure plus dynamic pressure. As the airflow divides and passes around the object the increases of local velocity, characterised by closely spaced streamlines, produce decreases of local static pressure. Pressures in excess of ambient atmospheric pressure are conventionally referred to as positive (+) and pressures below ambient atmospheric pressure as negative (-). The type of airflow around the body will be either Steady Streamline Flow or Unsteady Flow.

Steady Streamline Flow


2. In this type of airflow the flow pattern can be represented by streamlines. Where the streamlines appear close together high local velocities, greater than the free stream velocity, exist. Where the streamlines are widely separated velocity is lower than free stream velocity. Steady streamline flow can be divided into two types. (a) Classical Linear Flow. In this flow pattern the streamlines basically follow the contours of the body, with no separation of the airflow from the surface. Figure 2-1 illustrates classical linear flow around an aerofoil.

Chapter 2 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-1

(b)

Controlled Separated Flow or Leading Edge Vortex Flow. It is possible to design an aerofoil such that the airflow close to the surface separates at the leading edge and forms a controlled vortex. The main streamline flow is then around this vortex. This has advantages with swept-wing and delta wing planforms, as will be shown in later chapters. This is shown at Figure 2-2.

FIGURE 2-2 Leading Edge Vortex Flow

Chapter 2 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Unsteady Flow
3. Unsteady Flow occurs when the airflow separates from the surface of the body and the flow parameters (eg. speed, direction, pressure), at any point, vary with time. The flow thus cannot be represented by streamlines.

Two-dimensional Flow
4. An aerofoil section has only two dimensions, from leading edge to trailing edge and from upper surface to lower surface, known as chord and its thickness. Hence, when considering airflow around it, the consideration is limited to flow in two dimensions only.

Aerodynamic Forces on Surfaces


5. Associated with the velocity changes as air flows around an aerofoil there will, as has already been explained, also be pressure changes. If the aerofoil is inclined to the airflow as shown in Figure 2-3, it will be seen from the streamlines that the velocity over the upper surface is greater than that over the lower surface. According to Bernoulli, the greater the velocity the lower the local pressure, so there is a pressure difference between the upper and lower surfaces such that a force will be acting upwards.

Chapter 2 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-3 Airflow Around an Aerofoil

6. The angle at which the aerofoil is inclined to the airflow is called the angle of attack (). The greater the angle of attack, the greater the pressure difference and therefore the greater the upward force produced. This is true up to the point at which the airflow separates from the upper surface, known as the point of stall. As the airflow approaches the leading edge of the aerofoil it is turned towards the lower pressure on the upper surface. This effect is known as upwash. As it leaves the trailing edge it returns to its original, free stream location and this is termed downwash. The upward force produced as air flows over the aerofoil is the source of lift. 7. Figure 2-4 illustrates the aerodynamic forces acting upon an aerofoil inclined at an angle of attack () to the relative airflow.

Chapter 2 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-4 Aerodynamic Forces Acting on an Aerofoil

8. The resultant of all the aerodynamic forces acting on the wing is also referred to as the Total Reaction. The lift force is the component of the resultant force acting perpendicular to the relative air flow. The component of the resultant force acting parallel to the relative airflow is known as drag.

Streamline Pattern and Pressure Distribution


9. Let us now consider the pressure distribution around a symmetrical aerofoil. A symmetrical aerofoil is one in which the chord line and the mean chord line are co-incident. Figure 2-5a shows the streamline pattern around a symmetrical aerofoil at zero degrees angle of attack and Figure 2-5b shows the pressure distribution for the same situation.

Chapter 2 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-5 Symmetrical Aerofoil at Zero Lift Angle of Attack (a) Streamline Flow (b) Pressure Distribution

10. Notice from Figure 2-5(b) how the pressure distribution around the aerofoil can be conveniently represented in vector form. The pressure at any point on the upper and lower surfaces of the aerofoil is represented by a vector at right angles to the surface and whose length is proportional to the difference between absolute pressure at that point and free stream static pressure. Conventionally, pressures higher than ambient, ie. positive, are represented by a vector plotted towards the surface and for negative values, the vector is plotted away from the surface.

Chapter 2 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
11. At the leading edge of the aerofoil, where the streamlines diverge, positive pressure exists. Where the airflow is forced to divide and flow around the aerofoil the streamlines are close together and high local velocities and negative static pressure exists. The negative pressures are the same above and below the aerofoil, so with no pressure difference between upper and lower surfaces no lift is generated. The angle of attack at which this occurs is referred to as the zero lift angle of attack, for that particular aerofoil. It should be noted that a stagnation point also occurs at the trailing edge, where the flow velocity decreases to free stream velocity. 12. Consider now the symmetrical aerofoil at a positive angle of attack as shown in Figure 2-6. The greatest local velocities occur where the streamlines are forced into the greatest curvature as shown at Figure 2-6a. Consequently the highest velocities occur over the forward part of the upper surface. Upwash is generated ahead of the aerofoil, moving the forward stagnation point under the leading edge and creating an area of decreased local velocity below the forward part of the lower surface. Behind the aerofoil downwash is generated.

Chapter 2 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-6 Symmetrical Aerofoil at Positive Angle of Attack (a) Streamline Flow (b) Pressure Distribution

13. Figure 2-6(b) illustrates the pressure distribution from which it can be seen that there is a marked pressure differential between upper and lower surfaces, creating positive lift.

Chapter 2 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
14. A practical aircraft wing would not normally be of symmetrical aerofoil section, but would have some positive camber since such a section is capable of producing lift even at very low angles of attack. 15. Figure 2-7 illustrates the pressure distribution around a conventional cambered aerofoil inclined at a small positive angle of attack.

FIGURE 2-7 Pressure Distribution Around a Cambered Aerofoil at a Small Positive Angle of Attack

Chapter 2 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
16. At point A at Figure 2-7, total pressure (pitot pressure, stagnation pressure) prevails and this is the forward stagnation point. As the air passes over the upper surface towards point B, it is moving into an area of reducing pressure and point B is where minimum pressure exists on the upper surface Beyond B, pressure is increasing until total pressure is recovered at the rear stagnation point C and thus the air travelling from B to C is moving against an adverse pressure gradient. This is most significant as the only way the air can travel against this adverse pressure gradient is by virtue of its kinetic energy and should this be insufficient the air flow will break away or separate from the wing. This concept is fully covered in Chapter 4 under Stalling. 17. Furthermore, if points A and C are stagnation points and there is negative pressure on both upper and lower surfaces, then points X on the upper surface and Y on the lower surfaces are points of static pressure. To reduce the effect of the pressure reduction on the lower surface the curvature of the lower wing surface is kept to a minimum.

Effect of Angle of Attack on Pressure Distribution


18. Figure 2-8 illustrates the pressure distribution around a conventional cambered aerofoil through the working range of angles of attack. Such an aerofoil produces lift at zero degrees angle of attack because the aerofoil over the upper surface, with its greater curvature is accelerated more than over the lower surface creating a pressure differential and thus positive lift. It follows, therefore, that the zero lift angle of attack is a negative value (for this particular aerofoil it is -4) when the decreased pressure above and below the aerofoil is equal and hence no lift is generated.

Chapter 2 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
19. As the angle of attack is progressively increased, the negative pressure above the upper surface steadily increases, whilst that below the lower surface decreases. Beyond about +8the pressure below the lower surface becomes positive. Thus, it can be seen that with increasing angle of attack, the pressure differential between upper and lower aerofoil surface increases. However, at the lower angles of attack, it is the pressure reduction on the upper surface which is largely responsible for the lift generated whereas at the higher incidence, it is both the reduced pressure on the upper surface and the increased pressure on the lower surface which contribute to lift generation.

Chapter 2 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-8 Pressure Distribution Around an Aerofoil

Chapter 2 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Flow Separation at High Angles of Attack


20. Beyond about +14 in this typical aerofoil the low-pressure area on the upper surface suddenly reduces as the airflow separates from the surface, becoming unstable and turbulent instead of streamline, significantly reducing the total lift. The contribution to the total lift produced by the increased pressure on the lower surface however, remains relatively unchanged. This occurs at the critical or stalling angle of attack. At angles of attack beyond the stall the aerofoil may be regarded as a flat plate inclined to the airflow, as shown in Figure 2-9 and the lift produced is as a result of the stagnation pressure and flow deflection below the plate, but this is more than offset by the high drag force due to the plates resistance to the airflow. Stalling is fully covered in Chapter 4.

Chapter 2 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-9 Flat Plate Effect

Centre of Pressure
21. The pressure differential between the upper and lower surfaces can be conveniently represented by a single aerodynamic force acting at a particular point on the chord line. This point is known as the centre of pressure (CP). Both the resultant aerodynamic force and hence lift, and the point through which it acts (CP) vary with angle of attack.

Chapter 2 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
22. From Figure 2-8 it can be seen that as the angle of attack increases, the magnitude of the force increases and the centre of pressure gradually moves forwards towards the leading edge until the point of stall when, with the aerofoil past its stalling angle of attack (or critical angle), the force reduces and the CP moves rapidly rearwards. With a cambered aerofoil, the centre of pressure movement over the normal operating range of angles of attack is no further forward than approximately 25 - 30% chord, measured from the leading edge. With a symmetrical aerofoil section there is virtually no movement of the CP over the working range of angles of attack, in subsonic flight. 23. The movement of the CP with angle of attack is shown for a cambered aerofoil at Figure 2-10.

Chapter 2 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-10 Centre of Pressure Movement

Chapter 2 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
24. Let us now briefly consider moments, couples and coefficients.

Moment
25. The moment of a force about any point is the product of the force and the perpendicular distance from the line of action of the force to that point. See Figure 2-11.

FIGURE 2-11 Moment of a Force

Chapter 2 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Couple
26. Two equal forces acting parallel but in opposite directions are called a couple. The moment of a couple is the product of one of the forces and the perpendicular distance between them. See Figure 2-12.

FIGURE 2-12 Moment of a Couple

Chapter 2 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Coefficients
27. When considering lift, drag and pitching moments, it is much more convenient to use their respective co-efficients, CL,CD,CM. These co-efficients are non-dimensional and independent of density, scale of the aerofoil and velocity prevailing at the time such studies are carried out. They depend on the shape of the aerofoil and vary with angle of attack.

Aerodynamic Centre
28. An aircraft pitches about the lateral axis which passes through the centre of gravity. The wing pitching moment is therefore the product of lift and the distance between CG and CP of the wing. But, as we know, the position of the CP is not fixed and moves with changes in angle of attack and therefore, calculation of the pitching moment is quite involved and complicated. 29. The pitching moment and hence its coefficient (CM) depends not only on the lift force and the position of the CP, both of which change with change in angle of attack, but also the point about which the moment is considered. 30. For example, if we take a point of reference arbitrarily towards the leading edge then the nose down pitching moment about this point (B), increases with increasing angle of attack because, although the centre of pressure movement is forward, its effect is less than that of the increasing lift force, as shown in Figure 2-13.

Chapter 2 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-13

31. Now, about a point towards the trailing edge (A) the nose up pitching moment increases progressively with increasing incidence as shown at Figure 2-14.

Chapter 2 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-14 Pitching Moment Change About Point A

32. It follows, therefore, that if about the leading edge, the nose-down pitching moment progressively increases and about the trailing edge, the nose-up pitching moment progressively increases, then there must be a point somewhere on the chord line between points A and B about which there is no change in pitching moment with changes in angles of attack. This point is the wing aerodynamic centre, shown at Figure 2-15 and is, at subsonic speeds, approximately at quarter chord (ie. 25% chord from the leading edge).

Chapter 2 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-15 Pitching Moment About Wing Aerodynamic Centre

33. This can be represented graphically at Figure 2-16 which shows curves of CM plotted against CL where, conventionally, nose-up pitching moments are referred to as positive, and nose-down, negative.

Chapter 2 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-16
Cm against C L

Chapter 2 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
34. It can be seen from Figure 2-16 that at zero lift there is a residual pitching moment present. It is negative, and, by definition, remains constant about the aerodynamic centre up to the stall (ie CLmax). By reference to Figure 2-17 which shows the pressure distribution around our cambered aerofoil section at its zero lift angle of attack, it can be seen that the resultant forces due to the pressure reduction torwards the trailing edge on the upper surface and towards the leading edge on the lower surface produce a nose-down (negative) pitching moment. 35. The pitching moment coefficient CM at the zero lift angle of attack is referred to as CMo as shown in Figure 2-16.

Chapter 2 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-17 Pressure Distribution at Zero Lift Angle of Attack

Chapter 2 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Three Dimensional Flow


36. When considering airflow around an aircraft wing the flow becomes three-dimensional. This is because there is an element of spanwise flow above and below the wing in addition to the chordwise flow already discussed.

Spanwise Flow
37. When an aircraft wing is producing lift the local static pressure on the upper surface is lower than that on the lower surface. Air will flow from an area of higher pressure to one of lower pressure. Since a wing is of finite length, this means that air will flow from the under surface, around the wingtip, to the upper surface. Consequently, a spanwise flow of air occurs from the root outwards towards the tip on the under surface, around the tip, and from the tip inwards towards the root on the upper surface. The effect is illustrated at Figure 2-18.

FIGURE 2-18 Spanwise Flow

38. The flow at any point on the trailing edge leaving the upper surface will, therefore, be moving in a different direction from that leaving the lower surface as shown at Figure 2-19.

Chapter 2 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-19 Upper and Lower Surface Flow

39. Thus the flow at the trailing edge of the wing where the upper and lower surface flows meet is of a vortex nature and these vortices are continuously shed all along the trailing edge, shown at Figure 2-20 in which the trailing edge is viewed from behind.

FIGURE 2-20 Trailing Edge Vortices

Chapter 2 Page 27

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Wing Tip Vortices


40. For a rectangular planform wing, the spanwise flow at the wing tip is strong, decreasing inwards from the tip until it is zero at the root. Consequently, the vortices along the trailing edge tend to roll-up into a concentrated larger vortex towards each tip as shown in Figure 2-21. These vortices are known as wing-top vortices and, when viewed from behind, rotate in a clockwise direction on the port wing and anticlockwise on the starboard.

Chapter 2 Page 28

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-21 Wing Tip Vortices

Chapter 2 Page 29

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Induced Downwash
41. The effect of these trailing vortices is to produce a downward airflow, or downwash which influences the whole flow over the wing with two important consequences: (a) The effective angle of attack is reduced by the modified relative airflow and thus the lift generated is also reduced. This is shown at Figure 2-22. Furthermore, the drag characteristics of the wing are adversely affected and this induced or vortex drag will be covered in detail in Chapter 3. The flow over the tailplane in a conventional aircraft design will be affected by the downwash such that its effective incidence is also reduced with important consequences in respect of longitudinal stability, which is covered fully in Chapter 8.

(b)

FIGURE 2-22 Downwash Effect on Angle of Attack

Chapter 2 Page 30

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
42. The magnitude of the downwash is determined by the vortex formation which in turn is a consequence of spanwise flow. Spanwise flow results from wing tip spillage, the magnitude of which is determined by the pressure differential between upper and lower surfaces. In other words, any parameter which increases the pressure differential will also increase downwash and its effects.

Spanwise Lift Distribution


43. The distribution of lift along the span of the wing depends upon a number of variables, one of which is the variation of chord length along the span (in other words, the wing planform). A rectangular planform (i.e. constant chord throughout the span) wing creates most of its trailing vortices at the tips, consequently downwash is greatest at the tips. A tapered wing, with the chord progressively narrowing toward the tip, produces a greater proportion of lift at the centre and the trailing vortices are greatest towards the wing root. 44. Theoretically a constant downwash condition along the span can be achieved if the lift increases from zero at the tip to a maximum at the root in an elliptical fashion as shown in Figure 2-23. Such a condition is highly desirable for the reduction of induced drag, (explained fully in Chapter 3), and one way of achieving it is with a planform in which the chord increases elliptically from tip to root. There are, however, manufacturing and structural difficulties with such a planform and it has been found that a close approximation to elliptical spanwise lift distribution is possible using a tapered wing with varying aerofoil section.

Chapter 2 Page 31

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-23 Elliptical Lift Distribution

45. A convenient way to consider lift distribution is to use the ratio of the lift coefficient at any given point on the wing span (the local lift coefficient) (Cl), to the overall wing lift coefficient (CL), and plot this against the semi-span distance. When this is done for an elliptical planform wing a constant value is obtained from root to tip since: Cl ------ = 1.0 CL 46. Figure 2-24 shows spanwise lift distribution in this format for a number of wing planforms. The elliptical planform (A) has only been used in a few cases, most notable being the Spitfire. The rectangular planform (B) is often used for light private aircraft and trainers, because of its favourable stall characteristics. Larger aircraft invariably use tapered wings, in order to limit structural weight and maintain stiffness, with a taper ratio of between 20% and 45% (C) and (D).

Chapter 2 Page 32

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-24 Spanwise Lift Distribution

Chapter 2 Page 33

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Wake Turbulence
47. The trailing vortices produced by a wing when it is creating lift extend for a considerable distance behind the aircraft. The greater the lift being produced, the stronger these vortices will be. With a large aircraft these vortices create significant turbulence in the wake of the aircraft and take several minutes to dissipate. This wake turbulence is sufficient to seriously affect the controllability of other aircraft entering it and pilots are strongly advised to maintain a specified separation when following a large aircraft, especially in close proximity to the ground (i.e. during the take-off and landing phase). The separation required will depend upon the relative sizes of the aircraft and may be several miles if the following aircraft is much smaller than the leading one. 48. In addition, the strength of the vortices is universally proportional to aircraft speed and aspect ratio. The increased angle of attack, for a given weight, associated with low speed and stronger vortices from a low aspect ratio wing will result in increased wake turbulence. Furtheremore, when trailing edge flaps are extended, extra vortices are shed from the flap tips which tend to weaken the tip vortices and hasten vortex breakdown. 49. The vortex strength will, therefore, be greatest with increased aircraft weight, reduced speed and clean configuration ie. shortly after take-off. 50. This wake turbulence is influenced by proximity of the aircraft to the ground and also wind conditions. The vortices slowly descend under downwash influence to approximately 1000ft below the aircraft until when, in ground effect, they drift outwards from the generating aircrafts track. Any prevailing cross-wind, between 5-10kts, will retain the upwind vortex on the generating aircrafts track ie. on the runway after take-off.

Chapter 2 Page 34

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Lift
51. Lift is defined as that component of the total aerodynamic force which is acting perpendicular to the direction of flight. The magnitude of the total aerodynamic force, and therefore the lift generated, is dependent upon a number of variables of which the following are the most important: (a) (b) (c) (d) (e) (f) (g) (h) Free stream velocity (V) Air density ( ) Wing area (S) Angle of attack () Wing planform and aerofoil section Surface condition (rough or smooth) Air viscosity () Compressibility of the air.

52. The last two variables, viscosity and compressibility of the air, and their effect on lift will be discussed in subsequent chapters. 53. However, the major factors are dynamic pressure (V), wing surface area (S) and the relative pressure distribution existing on the surface, ie. the coefficient of lift (CL).

Chapter 2 Page 35

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Coefficient of Lift (CL)


54. The simplified equation for calculating aerodynamic force is VS multiplied by a coefficient proportional to the change in force that occurs when angle of attack is changed. This coefficient is the lift coefficient (CL) and the equation for lift is: L = CL

V S

CL for a given aerofoil section and planform allows for varying angle of attack and other variables not included in the equation. By transposition of formula it can be seen that: lift C L = ---------------1 2 --V S 2 55. The coefficient of lift is the ratio of lift pressure to dynamic pressure.

Effect of Angle of Attack


56. It is convenient to represent lift in coefficient form (CL) and then consider the factors affecting lift in terms of CL which can then be depicted graphically. It is possible from experimentation to obtain values of CL and plot them against angle of attack for a given wing at constant airspeed and air density.

Chapter 2 Page 36

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
57. Figure 2-25 shows a graph of CL against angle of attack () for a moderately cambered aerofoil section. It will be seen that the curve is linear for the greater part, with the coefficient of lift beginning to fall off at about +14. The lift coefficient reaches a maximum value at about +15 (CL max) as the section reaches stalling angle (stall), otherwise known as the critical angle (crit). Therefore any further increase in angle of attack results in marked reduction in CL.

Chapter 2 Page 37

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-25 Lift Curve

Chapter 2 Page 38

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Effect of Camber
58. The slope of the CL/ curve is constant regardless of the camber of the aerofoil, but values of CL are greater for any given angle of attack in sections of increased camber. This is illustrated at Figure 2-26.

FIGURE 2-26 Effect of Camber on the Lift Curve

Chapter 2 Page 39

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
59. Note that the curve passing through the origin at Figure 2-26 is representative of a symmetrical wing section. At zero angle of attack such a section produces no lift since the pressure distribution on the upper and lower surfaces is identical and there is thus no pressure differential. As we know from our pressure distribution studies earlier in this chapter, the angle of attack at which the CL is zero is known as the zero-lift angle of attack. For a symmetrical aerofoil it is 0 and for a cambered section, typically between -2 and -4.

Effect of Leading Edge Radius


60. The shape of the leading edge largely determines the stall characteristics of a wing. A bulbous leading edge with a corresponding large radius results in a well-rounded peak to the CL curve whereas a small leading edge radius will encourage a leading edge stall as the airflow will be less able to negotiate the sharper corner at large angles of attack. The peak to the CL curve is much more pronounced and the small radius produces a correspondingly more abrupt stalling characteristic.

Chapter 2 Page 40

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-27 Effect of Leading Edge Radius on the Lift Curve

Chapter 2 Page 41

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Effect of Aspect Ratio


61. The effective angle of attack is reduced by induced downwash, ie. the downward component of airflow at the rear of the wind caused by trailing edge vortices. A wing of infinite span has no tip vortices, no induced downwash and therefore no reduction in the angle of attack and it is a wing of high aspect ratio which approaches this condition of infinite span. Conversely, a wing of low aspect ratio, having greater trailing edge vorticity will have a greater reduction in the effective angle of attack and thus produce less lift than a wing of high aspect ratio, with the same wing area. Figure 2-28 shows the effect of aspect ratio on the CL curve.

Chapter 2 Page 42

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-28 Effect of Aspect Ratio on the Lift Curve

Chapter 2 Page 43

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Effect of Sweepback
62.
2

If an aircrafts wings are swept back and the wing area remains the same, the aspect ratio

( span /area) must be less than its equivalent straight wing. Therefore, the effect on the CL curve for a swept wing compared to a straight wing is similar to that for a low aspect ratio wing when compared to a high aspect ratio wing . This effect is shown at Figure 2-29.

FIGURE 2-29 Effect of Sweepback on the Lift Curve

Chapter 2 Page 44

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
63. Nevertheless, this does not account for the distinct reduction in CLMAX for highly swept wings, however this is fully explained in Chapter 4, Stalling.

Effect of Surface Condition


64. Roughness of the wings surface, especially at or near the leading edge has a considerable effect particularly on CLMAX. Figure 2-33 shows the reduction in CLMAX for a roughened leading edge when compared to a relatively smooth surface. Any roughness of the wing surface beyond 25% has little effect on CLMAX or the curve gradient.

Chapter 2 Page 45

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-30 Effect of Leading Edge Roughness on the Lift Curve

Chapter 2 Page 46

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Effect of Ice and Frost


65. Ice or frost deposits on the aircraft surface will invariably have a detrimental effect upon the performance of the aeroplane. In either case the aerodynamic shape will be changed and the boundary layer performance will be altered such that turbulence and separation occur more readily than with a clean aircraft. Since the wing is responsible for the vast majority of the lift generated the formation of ice or frost on its surface may cause considerable changes to the aerodynamic characteristics of the aircraft.

Ice at the Stagnation Point


66. There are essentially two effects of large ice formations on the leading edge of the wing. In the first place the contour of the aerofoil section may be considerably changed, as shown at Figure 2-31.

FIGURE 2-31 Effect of Ice on Leading Edge

67. This will almost certainly induce severe local pressure gradients, reducing boundary layer velocity locally and possibly causing leading edge separation, with consequent loss of lift. Secondly, some forms of ice have great surface roughness that significantly increases surface friction. This reduces the boundary layer energy and increases drag. The overall effect is a decrease in the maximum lift coefficient and an increase in drag.

Chapter 2 Page 47

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
68. The effects in practical terms are that the aircraft will require more power to maintain a given airspeed, the stalling speed will be higher and the stalling angle of attack will be lower. 69. Leading edge icing is most likely to occur during flight in icing conditions and should be prevented at the onset by the correct use of anti-icing procedures. Its effect will be most noticeable at the low flight speeds associated with approach and landing, where the higher stalling speed will require a higher landing speed.

Surface Ice and Frost


70. A thin layer of ice or frost on the upper surface of the wing may not significantly change the aerodynamic contour of the aerofoil section. However, the surface roughness, especially of hard frost, can increase surface friction and reduce boundary layer energy sufficiently to promote a loss of lift by as much as 25%. There will also be an increase of drag due to the increased skin friction. 71. The loss of boundary layer energy will lead to separation and a reduction in the stalling angle of attack. The maximum lift coefficient will be reduced and stalling speed will be increased. 72. Surface coatings of frost can occur in flight, but are more commonly associated with ground formation. Application of adequate ground de-icing and anti-icing procedures in conditions where ice or frost may form on the upper surfaces of a parked aircraft are essential prior to flight. The loss of lift and increased drag due to such coatings will seriously reduce take-off performance, to the extent that the aircraft may have difficulty in becoming airborne. Even if it does, its climb performance may be degraded to the point where obstacles cannot be cleared.

Chapter 2 Page 48

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
73. In addition to the foregoing, there is always a weight penalty associated with ice accretions on the aircraft. The added weight means that airspeed must be higher or the angle of attack must be increased to produce the extra lift required. The latter choice is a dangerous one, bearing in mind that the ice will almost certainly have reduced the stalling angle of attack. It should also be borne in mind that leading edge icing will probably have rendered the angle of attack indicator inoperative and that the stall warning system does not compensate for the reduced critical angle. 74. Figure 2-32 shows the effects of ice and frost formation on the wing on the CL - curve.

FIGURE 2-32 Effect of Ice and Frost on C L

Chapter 2 Page 49

G LONGHURST 1999 All Rights Reserved Worldwide

Lift

Lift Coefficient and Speed for Constant Lift


75. In straight and level flight the lift is equal to the weight and: 1 2 - V S Lift = C L -2 76. Wing area (S) does not change and. at constant altitude, density () remains essentially constant. In order to maintain constant lift both the lift coefficient (CL) and speed (V) must be kept constant or, if one increases the other must decrease proportionately. Since varying angle of attack varies (CL), and the optimum angle of attack has been shown to be about +4, maintaining constant lift is best achieved by adjusting airspeed. The relation between the lift coefficient and speed, for constant lift, is shown in the graph at Figure 2-33.

Chapter 2 Page 50

G LONGHURST 1999 All Rights Reserved Worldwide

Lift
FIGURE 2-33 Lift Coefficient and Airspeed Relationship for Constant Lift

77. During flight the weight will progressively decrease as fuel is used and the lift must decrease accordingly, or the aircraft will climb. The ideal aerodynamic solution to this is to reduce airspeed progressively, but in commercial operations, for ease of flight planning and other reasons, it is normal to fly at constant speed and trim the aircraft to reduce angle of attack (incidence) progressively.

Chapter 2 Page 51

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Drag
Zero Lift Drag Lift Dependent Drag Total Drag Speed Stability

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Drag

1. The total drag force acting upon an aircraft in flight is the sum of all the components of the total aerodynamic force that are acting parallel and opposite to the direction of flight. Total drag is made up of those factors arising from the generation of lift (Lift Dependent Drag) and those which are present when no lift is being generated (Zero Lift Drag).

Zero Lift Drag


2. When an aircraft in flight is not generating any lift there is no component of the total aerodynamic force acting perpendicular to the flight path. Consequently all of the aerodynamic force must be acting parallel and opposite to the direction of flight. This force is known as zero lift, or parasite, drag and comprises surface friction drag, form drag and interference drag.

Profile Drag
3. Otherwise known as boundary layer drag, profile drag is the term used to describe the combined effects of boundary layer normal pressure drag (form drag) and surface friction drag.

Chapter 3 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Form Drag
4. We have already seen that when moving air is either totally or partially brought to rest on the surface of an object, a pressure greater than static pressure, that is to say total, or stagnation pressure, is acting on the surface of the body. The velocity differences between leading and trailing edge mean that there are also pressure differences, pressure at the low velocity trailing edge being greater than at the relative high velocity leading edge. This adverse pressure gradient opposes the airflow across the surface, creating pressure drag. Form drag is otherwise known as boundary layer normal pressure drag and can form a significant portion of the total drag force acting on the aircraft. 5. Consider a circular flat plate, which is placed in a wind tunnel so that the flat surface of the plate is at right angles to the flow of air. If, over the entire surface area, the air was brought completely to rest, a pressure equal to the dynamic pressure would be felt at all points. The force thus created would be equal to the dynamic pressure multiplied by the surface area of the plate, or VS, where S is the surface area. 6. The situation is complicated somewhat since the air is not brought totally to rest over the whole surface. Some of the air flows around the edges, resulting in the formation of a low pressure area behind the back of the plate. This effectively creates a suction, which tends to retard the airflow passing the plate (or the passage of the plate through the air). 7. A turbulent wake will form behind the plate, in the case of a flat plate the amount of turbulence will be considerable and the drag factor therefore will be extremely high. With a streamlined wing the amount of turbulence will be much lower and therefore the drag factor will be considerably reduced. We can therefore see that the shape as well as the frontal area will affect the amount of drag produced. It is the shape, which gives us the co-efficient of drag (CD), and the total form drag formula now reads:

Chapter 3 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
1 2 -V S Form Drag = C D -2 8. It is clear that a flat plate is a most inefficient shape to try and move through the air. If we now consider a sphere of the same diameter as a flat plate placed in the airflow, it is not hard to visualise that this shape will produce considerably less drag than the flat plate. The nature of the airflow around these two objects is illustrated at Figure 3-1.

FIGURE 3-1 Airflow Around a Flat Plate and Sphere

9. It is seen at Figure 3-1 that the very large turbulent wake created by the flat plate has now been replaced by a much smaller one as the air flows more smoothly around the surface of the sphere and the suction drag created behind the sphere is thus reduced. The total amount of form drag 1 2 - V S ) , however the co-efficient of created by the sphere is calculated in exactly the same way ( C D -2 drag for the sphere is very much smaller than that for the flat plate.

Chapter 3 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
10. Now consider a streamlined aerofoil (ignoring any lift which it may generate), as illustrated at Figure 3-2. This shape has a very low co-efficient of drag, since the air can follow the surface of the shape almost to the trailing edge before it separates from the surface of the aerofoil and becomes turbulent. The turbulent wake produced by a streamlined shape is therefore very small. It is not possible to entirely eliminate the turbulent wake, but within limits the streamlined body can be extended with a consequent reduction in the co-efficient of drag value.

FIGURE 3-2 Airflow Around a Streamlined Aerofoil

11. Clearly it would be impractical to extend the length of the aerofoil beyond certain sensible limits since the increased weight would outweigh the improvement in the co-efficient of drag. In any event, beyond a certain length, the effect of friction between the air and the aerofoil surface prevents any further reduction in the drag factor. The amount of streamlining of a body is expressed as a fineness ratio, which is its length divided by its maximum thickness, as shown at Figure 3-3.

Chapter 3 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-3 Fineness Ratio

Coefficient of Drag (CD)


12. As with lift, drag is an aerodynamic force and may be considered as a coefficient and it is this coefficient which is the major factor in this drag formula CDVS. By transposition it can be seen drag that:- C D = ----------------1 2 --V S 2 13. The coefficient of drag is the ratio of drag pressure to dynamic pressure.

14. Figure 3-4 shows a graph of CD against angle of attack (). From this it can be seen that at low angles of attack the drag coefficient is low and it changes only slightly with small changes of angle of attack. As angle of attack increases however, drag increases and at the upper end of the range even small changes in angle of attack produce a significant increase in drag. At the stall a large increase in drag occurs.

Chapter 3 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-4 Drag Curve

Chapter 3 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-5 Lift Curve

Chapter 3 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

The Lift/Drag Ratio


15. Clearly an aerofoil is at its most efficient when it is generating the greatest possible lift for the least possible drag. From Figure 3-5 it will be seen that maximum lift is generated at an angle of attack of about 15,ie. the stalling angle. From Figure 3-4 it is seen that the least drag occurs at an angle of attack of about -2. Neither of these angles is practical for normal flight and neither is satisfactory, as the ratio of lift to drag is low in each case. What is needed is an operating angle of attack at which the lift force is high for a low drag force. In other words, a high ratio of lift to drag. 16. By combining Figure 3-4 and Figure 3-5 values of CL and CD for each angle of attack can be obtained and the ratio CL/CD calculated for each angle. A graph of CL/CD ratio against angle of attack can then be plotted, from which the best lift/drag ratio angle of attack is evident. Such a graph is illustrated at Figure 3-6, from which it can be seen that the best lift/drag ratio occurs typically at about +4. At this angle the ratio of lift to drag is likely to be between 12:1 and 25:1, depending upon the aerofoil section used and is referred to as the optimum angle of attack.

Chapter 3 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-6 Lift/Drag Curve

Chapter 3 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
17. At the zero lift angle of attack the lift/drag ratio is zero, but increases significantly for small increases in angle up to the optimum angle of attack of about +4. Beyond this the lift/drag ratio decreases steadily since, although lift is increasing, drag is increasing at a greater rate. At the stall the lift/drag ratio falls off markedly.

Surface Friction Drag


18. Modern aircraft have skin surfaces, which appear to be very smooth and polished. A closer inspection under a magnifying glass would reveal an irregular pitted surface, with the irregularities having massive dimensions when compared to the individual molecules of air flowing over the surface. It is not surprising then that the air immediately in contact with the aircraft surface is brought virtually to rest. This impedes the flow of air layer by layer until the point is reached where the air is flowing freely at free-stream speed. The total depth of air, which is flowing at less than 99% of free-stream velocity, is known as the boundary layer. The force required to overcome the shearing friction within the boundary layer is known as surface friction drag and is determined by the surface area of the aircraft, the viscosity of the air and the rate of change of velocity through the boundary layer, as illustrated at Figure 3-7.

Chapter 3 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-7 Boundary Layer

19. Within the boundary layer, and certainly within the free-stream air, it is hoped that the airflow will be laminar from the leading edge almost to the trailing edge. At a point known as the transition point, the smooth flow breaks down into a turbulent flow, which creates a much thicker boundary layer, see Figure 3-8.

Chapter 3 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-8 Transition Point

20. The effect of surface friction is, not surprisingly, far more marked in the turbulent part of the boundary layer than in the laminar part. It is therefore desirable to hang on to the laminar flow for a long as possible. 21. One of the main factors affecting the position of the transition point is the pressure distribution of the upper surface of the wing. Transition tends to occur at the minimum pressure point on the top surface, and this tends to occur at the point of maximum thickness of the wing itself. Therefore, by designing a wing where the maximum thickness occurs well back from the leading edge the laminar flow is increased, and the surface friction consequently reduced. A conventional wing and a laminar flow wing are shown at Figure 3-9.

Chapter 3 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-9 Aerofoil Sections

Interference Drag
22. Where two boundary layers meet, as at the junction between wing and fuselage, turbulent flow will ensue, leading to an increased pressure difference between leading and trailing surface areas and resulting pressure drag. The effect can be largely reduced in subsonic flight by adequate fairing at the junctions.

Profile Drag and Airspeed


23. Any form of profile drag will increase with increasing airspeed, as shown in the graph at Figure 3-10 and is in fact, proportional to the square of the speed.

Chapter 3 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-10 Profile Drag Speed Curve

Lift Dependent Drag


24. When an aircraft is generating lift additional drag is produced. This comprises induced (vortex) drag plus increases in the components that make up zero lift drag.

Chapter 3 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Induced Drag
25. As has been stated, the tip and trailing edge vortices create a downwash that angles the relative airflow to the direction of flight. The effect of the vortices is to direct the airflow downwards from the trailing edge. The angle between the airflow as it would be without induced drag, and the actual airflow, is termed the downwash angle. See Figure 3-11.

FIGURE 3-11 Downwashed Airflow

26. This induced downwash flow aft of the trailing edge influences the flow over the whole wing, (see Chapter 2-41), such that the effective angle of attack is reduced. As a result, the lift generated is also reduced and can only be restored by increasing the angle of attack. This increase in angle of attack will tilt rearwards the total reaction vector and thus the component parallel to the direction of flight is increased. This increase in drag, due to the wing vortices is induced (vortex) drag and is shown in Figure 3-12.

Chapter 3 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-12 Induced (Vortex) Drag

Chapter 3 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
27. Figure 3-12(a) represents a section of a two dimensional wing, ie. one of infinite span which is producing lift, but has no trailing edge vortices. 28. Figure 3-12(b) represents the same section but whose wing is of a finite span and thus having trailing edge vortices and hence induced downwash flow. The reduced effective angle of attack results in a reduction in the lift generated as shown. 29. In order to restore the lift to its two dimensional value (ie. value without downwash as at Figure 3-12(a), the angle of attack must be increased and the subsequent inclination of the total reaction vector causes an increase in the component parallel to the direction flight (ie. induced drag).

Effect of Airspeed
30. The lift generated by a wing can be increased by increasing the angle of attack for a given airspeed or by increasing the airspeed for a given angle of attack. The faster you fly, the lower the angle of attack necessary for the lift required and, the lower the angle of attack the less the downwash, thus reducing the induced drag, see Figure 3-13.

Chapter 3 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-13 Effect of Speed on Induced Drag

Chapter 3 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
31. Suppose, whilst maintaining level flight, the airspeed were to be doubled. The dynamic pressure producing lift (V) would be quadrupled. In order to maintain level flight the angle of attack would have to be reduced, thereby inclining the lift vector forward and reducing induced drag. Hence, induced drag decreases with increased airspeed being inversely proportional to the square of the speed, as shown in the graph at Figure 3-14.

FIGURE 3-14 Induced Drag Speed Curve

Chapter 3 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Effect of Aspect Ratio


32. Aspect ratio is defined as the ratio of the overall wing span to the mean chord. Since the magnitude of the induced drag of a wing depends upon the magnitude of the tip vortices, anything that can be done to reduce these vortices must reduce induced drag. The longer and narrower a wing the less the proportion of airflow around the tips to that over the remainder of the wing and therefore the less the influence of the tip vortices. In other words, less downwash so less induced drag than for a wing of the same area, but lower aspect ratio. This is illustrated at Figure 3-15.

FIGURE 3-15 Effect of Aspect Ratio on Induced Drag

Chapter 3 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
33. The greater the aspect ratio, then, the lower the induced drag of the wing. Taken to its ultimate conclusion, a wing of infinite span would have no induced drag at all. Clearly this is not feasible in the practical sense, although it can be proved in a wind tunnel experiment with a wing extending the full width of the tunnel. For aircraft in which low drag at moderate airspeed is a fundamental requirement, high aspect ratio wings are essential. Examples are sailplanes, long range patrol aircraft and medium speed transports.

Effect of Planform
34. Since induced drag decreases with increasing airspeed the need for high aspect ratio wings is less important for aircraft designed to operate at high subsonic or supersonic speeds. Indeed, for these a low aspect ratio is important because thin aerofoil sections are necessary, demanding a short wingspan for structural reasons. Concorde, for example, has an aspect ratio of less than 1:1, whereas a high performance sailplane may have an aspect ratio of 45:1 or greater. Clearly, with aircraft in which normal operations demand a low aspect ratio, high induced drag at the low speeds of take-off and landing has to be accepted. Similarly, training aircraft that benefit from the favourable stall characteristics of a rectangular planform wing suffer from greater induced drag than would be the case with a tapered wing. An elliptical planform gives the least induced drag for any given aspect ratio. Sweep back has a similar effect on lift distribution to decreasing taper ratio. A large sweep back angle tends to increase induced drag.

Effect of Lift and Weight


35. The downwash angle, and therefore the induced drag, depends upon the pressure differential between upper and lower wing surfaces. Consequently, an increase in lift coefficient (CL), whether due to increased weight or manoeuvre, must result in greater induced drag at a given speed. Induced drag varies as CL, and therefore as W at any given speed.

Chapter 3 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Induced Drag Coefficient (CDi)


36. For an elliptical wing planform the coefficient of induced drag is given as: ( CL ) C Di = ----------A 37. For any other planform a correction factor k is necessary, although a straight tapered wing with a taper ratio of 2:1 approximates very closely to an elliptical wing. For planforms whose lift loading is not elliptical k > 1. For conventional low speed wings the value of k is usually about 1.1 to 1.3. 38. The above equation for induced drag coefficient shows that: (a) (b) (c) There is no induced drag on a wing of infinite span (two-dimensional flow only). The greater the aspect ratio, the less the induced drag. There is no induced drag on any wing at zero lift.
2

39. Total drag is the sum of profile drag and induced drag and if the coefficients of total, profile and induced drag can be represented as CD, CDP=and CDi=respectively, then CD = CDP=+=CDi from Paragraph 3-23 above , for a wing with elliptical loading

Chapter 3 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
CL C Di = -------A However, for any other load distribution k CL 2 C Di = ----------A k CL 2 C D = C Dp + ----------A k ------A
2

where k is the induced drag factor.

Therefore

and now

can, for a given wing, be replaced by a constant K, so that, CD = CDp=+=K CL

40. Assuming that CDp= is constant, which is valid over the normal operating range of angles of attack, a graph of CD against CL as at Figure 3-16, enables the factor K to be found and hence, for a given wing, the induced drag factor k can be determined.

Chapter 3 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-16

CD - CL Curve

Induced Angle of Attack


41. The effect of downwash is to reduce the effective angle of attack. From the induced drag coefficient equation it is possible to derive an equation for the induced angle of attack (I): CL i = 18.24 -----A

Chapter 3 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
42. From the foregoing, if the influence of induced drag upon the graph of lift coefficient (CL) against angle of attack () is considered, it will be seen from Figure 3-17 that the greater the induced drag the shallower the CL - curve. Curve A in Figure 3-17 represents a wing of infinite span, in which there is no downwash (no induced drag) and therefore no reduction of the effective angle of attack. Curve B represents a wing of finite span in which there is downwash (induced drag) and a consequent reduction of the effective angle of attack. Curve C also represents a finite wing having greater induced drag than wing B. As angle of attack increases, downwash increases and the effective angle of attack progressively decreases. Hence the lift generated is less, at a given actual angle of attack, the greater the induced drag of the wing. 43. The implication of this is that a wing of low aspect ratio, and therefore higher induced drag, will have a higher stalling angle of attack than one of high aspect ratio and low induced drag. The theoretical effect illustrated at Figure 3-17 is not as pronounced in practice.

Chapter 3 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-17 Effect of Aspect Ratio on Stalling Angle of Attack

44. When considering the influence of induced drag upon the relationship between lift and drag coefficients, the lift/drag ratio, it is seen from the graph of CL against CD at Figure 3-18 that as induced drag increases, the slope of the CL CD curve decreases, indicating reduced lift/drag ratio.

Chapter 3 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-18 Effect of Aspect Ratio on CL - CD (Aeroplane Polar)

Chapter 3 Page 27

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Ground Effect
45. During flight in close proximity to the ground the three-dimensional airflow pattern is changed, since the vertical component of airflow is restricted, or eliminated altogether. This modifies the upwash, downwash and tip vortices and therefore has a significant effect upon lift and drag. 46. Assuming the same lift coefficient is maintained, as a wing enters ground effect the upwash, downwash and tip vortices are all reduced, as shown in Figure 3-19. The wing is now behaving as though its aspect ratio had been increased and, as a consequence, its induced drag coefficient (CDi) and induced angle of attack (I) are both less for the same lift coefficient (CL).

Chapter 3 Page 28

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-19 Ground Effect

Chapter 3 Page 29

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
47. For ground effect to be significant the wing must be at a height considerably less than the span of the wing. When height is equal to half-span the reduction in induced drag coefficient is less than 10%, but at a height equal to quarter-span this value rises to almost 25%. This is illustrated at Figure 3-20.

FIGURE 3-20 Effect of Height on Induced Drag Coefficient

Chapter 3 Page 30

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
48. Because the tip, or trailing vortices are reduced when a wing is in ground effect the spanwise distribution of lift is altered and the induced angle of attack (I) is reduced. Consequently a lower angle of attack (incidence) is necessary to produce the same lift coefficient. In the graph of CL against at Figure 3-21 it will be seen that the slope of the curve is increased when the wing is in ground effect. A lower angle of attack is needed for a given lift coefficient, or greater lift is generated for a given angle of attack. It should be noted that, when in ground effect, the stalling angle of attack is reduced.

Chapter 3 Page 31

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-21 Influence of Ground Effect on the Lift Curve

49. It will be appreciated that, for normal flight operations, ground effect is only of significance during landing and take-off. The general effects are an increase in lift if a constant pitch attitude is maintained and an increase in speed if the same power is maintained.

Chapter 3 Page 32

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
50. Furthermore, the reduced downwash affecting the tailplane will result in a greater contribution to longitudinal stability and, as the download of the tailplane reduces, there will be a nose down pitching moment change. However, if the aircraft is of a high tail design, then any changes in downwash will not affect the tailplane and hence have no effect on longitudinal stability or trim. 51. Pressure changes at the pressure source associated with changes in upwash and downwash due to ground effect, will usually cause an increase in static pressure sensed and therefore a reduction in indicated airspeed and altitude.

Effect on Landing
52. During the final stages of the landing approach there will be a tendency for the aircraft to float beyond the intended touchdown point if it is brought into ground effect at a constant angle of attack. The float distance may be considerable if speed is at all excessive.

Effect on Take-Off
53. The main danger from ground effect during take-off arises from the fact that, if constant angle of attack is maintained, a loss of lift may be experienced as the aircraft leaves ground effect. It is essential to delay rotation until the recommended airspeed is attained. Otherwise, in marginal conditions such as high all-up weight, high ambient air temperature and/or low ambient density, it may prove impossible to climb out of ground effect.

Chapter 3 Page 33

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Wing Tip Design


54. Various design devices are used in an attempt to reduce drag due to tip vortices. Perhaps the simplest is that employed on some rectangular planform wings on light aircraft, where the wing tip is deliberately cut off or given an upward or downward bend, as shown in Figure 3-22. The object in both cases is to encourage separation of the tip airflow, thereby reducing the tip vortex. A reduction in drag has resulted from these devices.

Chapter 3 Page 34

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-22 Effect of Wing Tip Design

Chapter 3 Page 35

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

End Plates and Tip Tanks


55. It was thought that the installation of end plates on the wing tips might prevent the formation of trailing vortices, but it was found in practice that plates large enough to achieve this usually had detrimental effects upon the handling characteristics of the aircraft. Whilst end plates do not destroy the trailing vortices they do modify them in a beneficial way. External fuel tanks, or other stores, mounted on the wing tips can have a small end-plate effect, although their primary advantage is the relief of wing bending stresses in flight.

Winglets
56. At the wing tip there is a significant sidewash in the form of outwash below the lower surface and inwash over the upper surface. By mounting a winglet vertically at the wing tip, advantage can be taken of this sidewash in that the winglet will generate a lifting force, a component of which acts in the direction of flight, opposing drag. The main tip vortex now forms at the tip of the winglet, above the main wing, so that downwash from the main wing is reduced. Thus, induced drag is reduced due to the reduction of the trailing vortices. The effect is illustrated at Figure 3-23. Winglets can be mounted either above or below the wing tip, or both. However, winglets below the tip are unusual because by mounting them upwards takes full advantage of upwash.

Chapter 3 Page 36

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-23 Effect of Winglets

Chapter 3 Page 37

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Wing Span Loading


57. With a tapered wing most of the lift is being produced at the root, and therefore the downwash at the inboard end is significantly greater than at the outboard end of the wing. This means that the effective angle of attack is greater at the tip than at the root, as illustrated at Figure 324. This is undesirable for a number of reasons, not least of which is the fact that it will cause the tip to stall before the root. Additionally, it means that the local wing loading (lift per unit area) tends to be greater toward the tip. This can be compensated for by mechanically adjusting the effective angle of attack where necessary, either by twisting the wing or by altering the camber (aerofoil section) toward the tip.

FIGURE 3-24 Effective Angle of Attack Spanwise

Chapter 3 Page 38

G LONGHURST 1999 All Rights Reserved Worldwide

Drag

Wash-Out
58. The wing is attached to the fuselage of the aeroplane at a particular angle of incidence. If the angle is maintained throughout the span of a tapered wing, the situation described in the preceding paragraph will exist. If, however, the angle of incidence is progressively decreased toward the tip a constant effective angle of attack can be maintained. Decreasing the angle of incidence from root to tip is known as wash-out. The reverse is known as wash-in.

Change of Camber
59. The same effect as that produced by wash-out can be achieved by a progressive reduction of camber from root to tip.

Total Drag
60. Of the two types of drag to which an aircraft is subjected, profile drag increases with increasing airspeed, whilst induced drag decreases with increasing airspeed. By plotting both on a graph, it is possible to establish a speed at which the sum of the two is a minimum. This is the speed for minimum drag V MD .See Figure 3-25.

Chapter 3 Page 39

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-25 Effect of Speed on Drag

Drag and Pressure Altitude


61. It should be appreciated that for a given EAS, the dynamic pressure will be the same at all altitudes and therefore a plot of drag versus EAS will apply at all altitudes. See Figure 3-26.

Chapter 3 Page 40

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-26 Effect of Altitude on Drag - EAS Curve

62.

1 However, for a given EAS, the TAS with increasing altitude increases by the factor ------- , where

is the relative density, and therefore, if drag is now plotted against TAS, the curve is mirrored progessively to the right with increasing altitude. See Figure 3-27.

Chapter 3 Page 41

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-27 Effet of Altitude on Drag - TAS Curve

Speed Stability
63. Stability, which is fully explained in Chapter 8, is the study of the aircrafts response following a disturbance and its tendency thereafter to return, or otherwise, to its pre-disturbed condition. When considering speed stability, refer.ence is made to the drag curve shown at Figure 3-27.

Chapter 3 Page 42

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
Effect of Speed Changes on Drag

64. Consider an aircraft flying at speed A. In straight and level flight thrust equals drag and so the thrust required can be indicated by the horizontal line T1. If now there was an un-demanded speed increase from A to B, with the thrust remaining as it was, the drag exceeds the thrust and so the speed will reduce to A. Conversely, if the speed reduces to C, with the thrust remaining constant, the thrust exceeds the drag and so the speed will increase. 65. This then is a speed stable situation, which will always exist on the front side of the drag curve, ie. at speeds in excess of VMD.

Chapter 3 Page 43

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
66. Now consider the same situation on the back side of the drag curve, in other words where the aircraft is flying at a speed which is lower than VMD. Initially the speed is D and the thrust T2. 67. If the speed increases to E the drag decreases, the thrust exceeds the drag and so the speed increases further. This is bad enough, but now consider a drop in speed from D to F. The drag has increased, the thrust is constant and so the speed decreases further; the drag is increased, the thrust is insufficient and so on. This then is definitely a speed unstable situation. 68. The effect of increasing profile drag, such as the use of undercarriage or speed brake, will not alter the induced drag curve, but will steepen the profile drag curve. Thus the point of intersection, and hence VMD, reduces (although the total drag increases). Thus the speed range where the aircraft is speed stable will be greater. 69. The speed-instability situation below VMD is particularly important with modern jet transport aircraft. This is because of the high weights involved, the high angles of attack at the low airspeeds involved in the approach phase and the slow reaction times of jet engines, when compared with piston engines. 70. When dealing with induced drag it was seen that an increase in weight resulted in an increase in induced drag. Since the intersection of the profile and induced drag lines gives VMD we can therefore see, at Figure 3-28, that VMD increases as the weight increases and speed stability correspondingly reduces.

Chapter 3 Page 44

G LONGHURST 1999 All Rights Reserved Worldwide

Drag
FIGURE 3-28 Effect of Weight on Speed Stability

71. Using the same sort of logic it can be seen that the higher the aspect ratio the lower the VMD and speed stability therefore increases.

Chapter 3 Page 45

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Stalling
Stalling Speed Initial Stall in the Spanwise Direction Stall Warning

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Stalling

1. The stalling characteristics for large transport aeroplanes are laid down in JAR 25. The following are the relevant extracts: During the approach to the stall the longitudinal control pull force should increase continuously as speed is reduced from the trimmed speed to approximately 1.1Vs. At speeds below 1.1Vs some reduction in longitudinal control pull force will be acceptable provided that it is not sudden or excessive. In level wing stalls the bank angle may exceed twenty degrees occasionally, provided that lateral control is effective during recovery. For stalls from a thirty degree banked turn with an entry of one knot per second, the maximum bank angle which occurs during recovery should not exceed approximately sixty degrees in the original direction of the turn, or thirty degrees in the opposite direction. For dynamic stalls (that is those with a faster entry) the maximum bank angle which occurs during the recovery should not exceed approximately ninety degrees in the original direction of the turn, or sixty degrees in the opposite direction. Deep Stall Penetration. Where the results of wind tunnel tests reveal a risk of a catastrophic phenomenon (eg. superstall, a condition at angles beyond the stalling incidence from which it proves difficult or impossible to recovery the aeroplane), studies should be made to show that adequate recovery control is available at and sufficiently beyond the stalling incidence to avoid such as phenomenon.

Chapter 4 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
It must be possible to produce and to correct roll and yaw by unreversed use of aileron and rudder controls up to the time the aeroplane is stalled. No abnormal nose-up pitching may occur. The longitudinal control force must be positive up to and throughout the stall. In addition it must be possible to promptly prevent stalling and to recover from a stall by normal use of the controls. For level wing stalls, the roll occurring between the stall and the completion of recovery may not exceed approximately twenty degrees. For turning flight stalls, the action of the aeroplane after the stall may not be so violent or extreme as to make it difficult, with normal piloting skills, to effect a prompt recovery and to regain control of the aeroplane.

2. When considering surface friction drag in the previous chapter the boundary layer was discussed and illustrated. It was seen that the boundary layer airflow is normally laminar over most of the wing surface, becoming turbulent at a point known as the transition point.

Boundary Layer Separation


3. As air flows over the upper surface of the wing, beyond the point of minimum pressure, its velocity decreases and its pressure increases, so that the air is flowing against an adverse pressure gradient. It can do this provided it has sufficient kinetic energy due to velocity. However, if the velocity of the airflow falls too far the airflow close to the surface of the wing will cease altogether. 4. The point on the surface where airflow ceases is known as the separation point. Beyond this point the adverse pressure gradient will cause the airflow adjacent to the surface to reverse in direction, becoming eddying and turbulent with the mean velocity of the airflow in the opposite direction to the free stream airflow. Aft of the separation point little or no lift is generated.

Chapter 4 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Trailing Edge Separation


5. At low angles of attack virtually no separation occurs on a normal cambered subsonic aerofoil section forward of the trailing edge. The airflow is laminar over most of the surface and remains attached in the form of a turbulent boundary layer near the trailing edge, as illustrated at Figure 4-1.

FIGURE 4-1 Boundary Layer

Chapter 4 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
6. As the angle of attack is increased the adverse pressure gradient on the upper surface increases and boundary layer separation begins to occur near the trailing edge. Further increase in the angle of attack will cause the separation point to move forward towards the leading edge, reducing the amount of lift generated. Eventually an angle of attack will be reached at which flow over the upper surface of the wing breaks down completely. The lift coefficient ceases to rise and starts to fall if angle of attack is increased further. This is known as the critical, or stalling, angle of attack and is illustrated in the lift curve at Figure 4-2. The drag coefficient increases with increasing angle of attack, rising steeply at the stalling angle, as illustrated at Figure 4-3. The lift/drag ratio falls steeply at the stalling angle of attack.

Chapter 4 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-2 Lift Curve

Chapter 4 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-3 Drag Curve

Chapter 4 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Leading Edge Separation


7. On thin wings with sharp leading edges, as used in high-speed aircraft, a form of laminar flow separation may occur above the leading edge of the aerofoil. A stationary vortex, called a bubble, forms between the separated layer and the wing surface, modifying the airflow pattern as shown at Figure 4-4. As angle of attack is increased the bubble will eventually burst (i.e. the vortex collapses) and the wing stalls abruptly. In some case the bubble formed will extend over most of the chord length, in which case the stall when the bubble bursts is more gradual.

FIGURE 4-4 Leading Edge Separation

Centre of Pressure Movement


8. In Chapter 2 it was shown that the negative pressure above the upper surface of the wing progressively increases as the angle of attack is increased, up to the stalling angle. Examination of the diagrams illustrating this will show that the centre of pressure moves progressively forward as this occurs. At the point of stall the centre of pressure moves rapidly aft.

Chapter 4 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Pitching Moments
9. The effect of the centre of pressure movement on the wing is to produce an increasing pitchup moment up to the stall, followed by a marked pitch down moment as the CP moves rapidly aft at the stall. The pitching moments on the aeroplane as a whole will depend upon the configuration of the wing and tail surfaces. It is clearly desirable that there should be a nose-down pitching moment at the stall, since this will automatically reduce the angle of attack and assist in recovery from the stall. Loss of downwash over the tailplane (horizontal stabiliser) at the stall may achieve this in aeroplanes where the nose-down pitching moment of the wing section is negated by other factors.

Pre-Stall Buffet
10. As the disturbed wake, arising from the separated flow behind the wing, strikes the tail surfaces it causes aerodynamic buffeting on the control surfaces (elevator and rudder) which can usually be felt at the control column and rudder pedals, giving warning of the incipient stall. The effectiveness of this as a warning will depend to a large extent upon the location of the tail surfaces with respect to the turbulent wake. The effectiveness of the controls may well decrease as separation occurs. This is often particularly true of the ailerons, situated as they are at the trailing edge of the wing.

Chapter 4 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Stalling Speed
11. In straight and level flight the weight of the aircraft is balanced by the lift generated and Lift = 1 2 - V S . Given constant density and wing area the variables are, as we have seen, C L and V. C L -2 If airspeed (V) is reduced, by throttling back the engine, C L will have to be increased by increasing the angle of attack, if level flight is to be maintained at the reduced speed. This practice has its limits since, if pursued, an IAS will be reached where increasing the angle of attack will mean that the stalling angle is reached. The speed at which this occurs is the stalling speed. 12. The speed corresponding to a given angle of attack, or rather lift coefficient, can be derived from the lift formula: 1 2 -V S Lift = CL -2 13. Therefore: Lift V 2 = ----------------1 - S C L -2

Chapter 4 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
14. Therefore: V = Lift ----------------1 - S C L -2

15. Since the stall occurs when the coefficient of lift is at its maximum value (CLmax), the formula for the stalling speed (Vs) is: Vs = Lift -------------------------1 C Lmax --S 2

16. The only variables in this formula are speed and lift, so stalling speed is proportional to the square root of lift: V s Lift

Basic Stalling Speed


17. This is defined as the minimum steady flight speed at which the aircraft is controllable. It is the speed below which a clean aircraft (flaps and landing gear retracted) of given weight, with engines throttled back, can no longer maintain straight and level flight and it is usually quoted for a number of different aircraft weights.

Chapter 4 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
18. In straight and level flight lift = weight, so by substitution the formula for the basic stalling speed (VB) will be: VB = Weight -------------------------1 -S C Lmax -2

19. The actual stalling speed will differ from the basic stalling speed if the conditions by which it is defined are not met. The factors which will cause variation from VB are: (a) (b) (c) (d) (e) change in weight manoeuvre (load factor) power position of centre of gravity configuration (changes in CLmax)

Weight and the Stall


20. Any change of weight will require a different value of lift for straight and level flight, an increase in weight requiring an increase in lift. It is clear therefore that the higher the weight of the aircraft, the higher the stalling speed. 21. The formula for calculating the increased stalling speed for a given increase in weight is derived from the basic stalling speed formula as follows:

Chapter 4 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
W -------------------------1 C Lmax -- S 2 is V and for an

VB =

22.

Let us assume that the basic stalling speed at an aircraft weight of W 2

S1

increased weight W

is VS . The ratio of the two stalling speeds at the two different weights is:2

W2 -------------------------1 C - S Lmax 2 V S2 --------- = -----------------------------V S1 W1 -------------------------1 C Lmax -- S 2 23. Since the two denominators on the right hand side of the equation are the same: V S2 --------- = V S1 24. Therefore: W2 ------W1

Chapter 4 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
W2 V S2 = V S1 ------W1

Example
25. Given a stalling speed of 120 kt for an aircraft at 150,000 kg, the approximate stalling speed at 180,000 kg is: Vs2 = 120 1.2 = 131 kt

Load Factor and the Stall


26. During a manoeuvre, such as a banked turn or the pullout from a dive, the load factor on the wing is increased and the effect is the same as an increase in weight. During a level turn not only must the vertical component of lift balance the aircraft weight, but also the horizontal component of lift must provide the centripetal force required to maintain the turn. The result of this is that the lift produced in a level turn has to be greater than the lift produced in straight and level flight for the same weight. Provided that the speed is kept constant, the only way that this extra lift can be achieved is by an increase in angle of attack. This increase in angle of attack puts the aircraft wing nearer to the stalling angle. The net result of having to produce more lift from the wings is that the aircraft's weight appears to be increased, hence the expression G loading. The increase in stalling speed is calculated by taking the normal stalling speed in level flight for the aircraft's weight and multiplying it by the square root of the load factor (n). 27. The load factor (n) is given by the ratio:

Chapter 4 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
L ---W 28. If we refer to the manoeuvre stall speed as VM, the ratio: L -------------------------1 - S C Lmax 2 VM ------- = -----------------------------VB W -------------------------1 C Lmax -- S 2 29. Again, the denominators on the right hand side of the equation are equal so: VM ------- = VB 30. L ---W

The relationship L/W is the load factor (n), so: VM ------- = VB n

31.

Or V M = VB n

Chapter 4 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-5 Forces in a Steady Co-ordinatd Turn

Chapter 4 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
32. Figure 4-5 illustrates the forces acting upon an aircraft in a banked turn, at bank angle . From this it can be seen that: W1 W cos = ------- = ---L L 33. Therefore: 1 L - = n ----------- = ---cos W 34. Therefore, by substitution: 1 VM = V B ----------cos

Example
35. be: In a 60 banked turn the increase in stalling speed (VM) over the basic stalling speed (VB) will

1 V M = V B -----0.5

Chapter 4 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
V M = VB 2 36. Therefore: V M = V B 1.4 37. The manoeuvre stalling speed in a 60 banked level turn is 1.4 times the basic stalling speed.

Power and the Stall


38. Stall speeds in pilot's notes and other aircraft performance data are invariably given for the power off or the all power units in idle situation. Leaving power on will result in a lower stalling speed for a given weight and configuration, but may also lead to a more violent wing drop at the stall, especially with single engine propeller driven aircraft. 39. The decrease in stalling speed with power on is due to the vertical component of thrust, which is significant at the high angles of attack involved, and which obviously applies to either pure jet or propeller driven aircraft. With propeller driven aircraft the effect of the propeller slipstream over the wing will also delay the stall. The slipstream may, however, prevent the wing behind the power units from stalling (the inboard sections), thereby leaving the tips to stall first despite the best efforts of the aeroplane designer. 40. It should be noted that recovery from a power-on stall is likely to involve less height loss since power unit response will be faster.

Chapter 4 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Centre of Gravity and the Stall


41. The position of the centre of gravity affects both the stall speed and the ease with which the stall recovery is achieved. 42. With the centre of gravity well forward of the centre of pressure, a considerable download is required at the tailplane to maintain level flight. The lift produced by the wing is therefore required to support not only the aircraft weight but also the download produced by the tail. For a given weight, the further forward the centre of gravity the greater the angle of attack required at any given airspeed, and therefore the higher the speed at which the stalling angle is reached. 43. Notwithstanding the higher stall speed with the forward C of G, the further forward the C of G the easier the stall recovery, since the weight of the aircraft is helping the pilot to lower the nose of the aircraft as an essential part of the recovery procedure. In other words, if the C of G is to the rear of the permitted envelope, stall recovery may well be impossible.

Altitude and the Stall


44. It is important to appreciate that the wing (and therefore the aircraft) stalls at a given angle of attack rather than at a given airspeed. The stalling angle is normally in the region of 15, although this varies, depending on the type of wing considered, and the stall occurs at a particular lift coefficient ( C L ) . Given that: 1 2 L = CLmax --V S 2

Chapter 4 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
45. We know that CLmax and S are constant at the stalling angle, and we also know that indicated

1 2 airspeed is given by dynamic pressure -- V . It therefore follows that any aircraft will stall at a 2 constant IAS, regardless of altitude. This of course assumes constant weight and aircraft configuration, and straight and level flight. The statement also ignores the very small variations in the difference between equivalent and indicated airspeed as altitude is changed.

Initial Stall in the Spanwise Direction


46. So far we have considered the stall and its effects in the chordwise direction, but it is important to remember that airflow over a wing is three-dimensional and there will also be a spanwise effect when a wing stalls.

Tip Stall
47. As mentioned earlier, at the point of the stall the centre of pressure moves rapidly aft on the wing, which gives a nose-down pitch. This reduces the incidence of the wing, which will therefore once again produce lift, a desirable self-correcting characteristic.

Chapter 4 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
48. A wing does not normally stall over its entire length simultaneously, the stall commencing from one part of the wing. Having commenced at one point on the wing the stall then spreads over the remainder of the wing. The main factor governing the point at which the stall commences is the shape of the wing. It is undesirable that a wing stalls from its tip first as this may lead to control difficulties. Any tendency to drop a wing at the stall may well lead to spinning, especially if the stall recovery procedure is badly executed. Once the wingtip is dropping, the downward motion increases the effective angle of attack, deepening the stall on that wing, whilst the effective angle of attack of the upgoing wing is decreased, inhibiting the stall. This worsens the situation by increasing the rate of roll. Clearly, tip stall is an undesirable characteristic. 49. Further advantages of having a wing stall first from its root instead of the tip, are that aileron control can be maintained up to the point of full stall, and that the separated flow from the wing root may cause buffet over the tail which provides a stalling warning. 50. The principal techniques for ensuring that the stall occurs initially at the wing root are discussed below: 51. The wing may be twisted so that the tip is at a smaller angle of incidence than the root, this will ensure that the root reaches its stalling angle before the tip. This decreasing angle of incidence from root to tip is termed wash-out. 52. The same effect as that produced by wash-out can be achieved by a progressive reduction of camber from root to tip. 53. A stall inducer, stall strip or root spoiler, may be fitted at the inboard end of the wing, as shown at Figure 4-6. The strip has very little effect on the airflow at normal angles of attack, however at high angles it spoils the airflow over the wing behind it, causing buffet on the tailplane and ensuring this section of the wing stalls first.

Chapter 4 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-6 Stall Strip, Stall Inducer or Root Spoiler

54. Wing platform has a significant effect upon stall characteristics. Figure 4-7 shows various platforms and the stall progression associated with each.

Chapter 4 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-7 Effect of Planform on Stall Pattern

55. Wing tip shape will also influence tip stall characteristics. Rounded tips, as seen on many WW II fighters, or chamfered tips as seen on the Islander and Dornier 228, stabilise the wingtip vortices at high angles of attack (incidence).

Chapter 4 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Use of Ailerons
56. The downgoing aileron increases the camber and therefore steepens the slope of the C L curve. It therefore produces a small reduction of the stalling angle of attack at that part of the wing to which it is attached, usually near the tip. Consequently, the use of aileron when flying at a high angle of attack close to the stall may initiate tip stall on that side. In addition, when tip stall occurs, the downgoing aileron loses effectiveness which will probably worsen the situation.

Swept Wings
57. From Figure 4-7 it can be seen that swept wings are particularly prone to tip stall, the greater the sweepback the greater the tendency. 58. The primary purpose of sweepback is to increase the value of the critical Mach number for a given aircraft. The effects of compressibility at speeds approaching the critical Mach number on the performance of an aircraft are discussed at length in Chapter 11 dealing with high-speed flight. 59. The basic problem is that, if the air over any part of the aircraft structure (usually the wing) reaches the speed of sound there is an excessive rise in drag due to compressibility effects. The actual free air speed at which the accelerated air over the wing goes supersonic is known as the critical Mach number. Figure 4-8 shows a straight unswept wing travelling at a true airspeed corresponding to Mach 0.8. Note that the air that is accelerated over the top surface of the wing is travelling at Mach 0.85.

Chapter 4 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-8 Unswept wing at MO.8

Chapter 4 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
60. Figure 4-9 shows a swept wing, again travelling through the air at Mach 0.8. As before, the acceleration of air over the top surface means that the speed of Mach 0.85 is achieved on the top surface in the direction of the airflow. However, only the component of the relative airflow across the wing that is perpendicular to the leading edge can be considered as creating lift. Consequently, only the vector speed of this chordwise component is significant when considering critical Mach number.

FIGURE 4-9 Swept Wind at MO.8

Chapter 4 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
61. The wing shown at Figure 4-9 is travelling at the same speed through the air as the wing at Figure 4-8. The accelerated air at right angles to the leading edge of the swept wing is travelling slower than the equivalent air across the straight wing, by 11% of the local speed of sound. Another way of looking at this is that the aerofoil section along the chord line (shown dashed at Figure 4-9 is thinner than its equivalent on an unswept wing. Hence the acceleration of airflow over the upper surface is less and the local velocity is lower. 62. At 35,000 ft (standard atmosphere) the local speed of sound is 575 kts and so 11% of this is 63 kt. Clearly the ability to fly 63 kts faster before reaching criticality is a big reason for using a swept wing. You don't have to spend too long looking at modern jet transport aircraft to realise that the manufacturers agree with this philosophy. 63. As already mentioned, the amount of lift generated by the wing depends upon the flow perpendicular to the leading edge. In view of this and all other things being equal, less lift will be produced by a swept wing than by a straight wing of equal area. To regain this lost lift, the angle of attack has to be increased when a swept wing is used. The lift curves (co-efficient of lift against angle of attack) are shown for both types of wing at Figure 4-10.

Chapter 4 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-10 Effect of Sweepback on the Lift Curve

Chapter 4 Page 27

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
64. It is readily seen at Figure 4-10 that the higher the angle of attack, the bigger the difference in the amount of lift generated. High angles of attack are associated with low speeds and the effect is most evident on the approach to land in swept wing jet aircraft, where a very high nose-up attitude is necessary. 65. Returning once more to Figure 4-9, the total airflow over the swept wing is shown broken down into a chordwise vector and a spanwise vector. The flow of air outwards along the wing causes the boundary layer on a swept wing to drift outwards towards the wing tip. This results in an undesirably thick boundary layer in the region of the wing tip. 66. We know that the retardation of the air by the boundary layer is one of the major causes of the stall and perhaps now we can now perhaps see an extremely undesirable situation evolving. A thick boundary layer will encourage the stall. The boundary layer is thicker at the wing tip and therefore the tip is likely to stall before the root. The undesirability of tip stall has already been discussed. Quite obviously then, something needs to be done to prevent this migration of the boundary layer outwards to the wing tip. 67. One simple way of preventing spanwise flow is to put a physical barrier in the way, as shown at Figure 4-11. This device is known as a wing fence.

Chapter 4 Page 28

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-11 Wing Fence

68. Spanwise flow can also be reduced by means of a saw tooth leading edge as shown at Figure 4-12, or a notched leading edge as shown at Figure 4-13. Both of these devices produce a very marked vortex that inhibits the spanwise flow.

Chapter 4 Page 29

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-12 Sawtooth Leading Edge

FIGURE 4-13 Notched Leading Edge

Chapter 4 Page 30

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
69. Because tip stall is reduced, the pitch up tendency at the stall will also be reduced, however there are other benefits to be derived from the saw tooth leading edge. Since the thickness/chord ratio of the tip area is reduced, the critical Mach number will be increased. Furthermore, the centre of pressure of the extended wing is further forward than it would otherwise be and therefore the mean centre of pressure of the whole wing is further forward. Consequently, when the tip eventually stalls, the forward movement of the centre of pressure is less marked, giving a less pronounced pitch up moment. 70. By causing a vortex in the wing section behind the notched leading edge, the magnitude of the vortex over the tip area is reduced, and with it the magnitude of the tip stall. 71. Sometimes the notched leading edge is used in conjunction with the saw tooth leading edge extension. This intensifies the inboard vortex behind the devices and thus a stronger restraining effect upon the boundary layer outflow. 72. Having suggested that the production of a vortex will inhibit the spanwise flow, it should come as no surprise that the final means of arresting this flow which we consider is the use of vortex generators. These are quite simply small metal plates which are mounted edge on to the airflow and which produce vortices behind them. The vortices are small, thus incurring small drag penalties, but they act by bringing down the fast-moving air above the boundary layer. This adds kinetic energy to the boundary layer, enabling it to better cope with the adverse pressure gradient. This is shown at Figure 4-14.

Chapter 4 Page 31

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-14 Vortex Generators

73. Despite the use of some or all of these devices it is inevitable that, with a significantly swept wing, the tips will stall first, and that the stall will then spread inwards towards the roots. As this is happening the total lift generated by the wing is reducing, the centre of pressure is moving inwards and, because of the sweepback, the centre of pressure moves forward. This is shown at Figure 4-15.

Chapter 4 Page 32

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-15 Effect of Tip Stall on Centre of Pressure

74. With a straight wing aircraft the nose tends to drop at the stall, and this is obviously desirable. With a swept wing aircraft the nose tends to pitch up at the stall, because of the forward movement of the centre of pressure, and this is obviously undesirable, since it may lead to a deep stall condition. 75. When a swept wing is subjected to increased loading, as in a turn, the structural nature of the wing produces not only an upwards bend but a twist as well. This is caused by the sweepback placing most of the outboard structure of the wing behind the main structural member, and the result of this is to reduce the angle of attack of the tip. As the angle of attack is reduced, the amount of lift being generated by the tip is also reduced and the centre of pressure of the wing moves inwards towards the root. As with the stall, movement of the centre of pressure inwards also means that it moves forward, and the overall result of this is that in a turn the nose of a swept wing aircraft tends to pitch up.

Chapter 4 Page 33

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Stall Warning
76. It is required that adequate advance warning of the stall is available (the margin) in order that early corrective action may be taken. In light aircraft warning of approach to the stall is usually achieved aerodynamically in the form of buffet. Where this is absent, or could be confused with turbulence, the stall warning is artificially induced. In heavy aircraft with powered controls it is normal to incorporate a stick shaker. This device shakes the stick in order to simulate the effect of the turbulent airflow over the elevators which would normally be felt at the stick as pre-stall buffet, but which is masked by the electro-hydraulically powered control system. Stick shakers are operated from a detector that senses incidence (angle of attack) and rate of change of incidence. Having gone this far it is normal to incorporate also a stick pusher. If the pilot doesn't respond to the stick shaker, the stick pusher automatically ensures that the correct incipient stall recovery procedure is initiated. 77. Following a stall warning the recovery action is to reduce incidence to below warning activation value and at the same time apply maximum power to minimise height loss. If the stall has occurred the recovery action is the same, but if wing drop occurs, lateral correction should not be applied until after the wings are unstalled since the application of aileron before the wings are unstalled will probably aggravate the wing drop. It should be appreciated that recovery at the warning stage will involve less height loss.

Chapter 4 Page 34

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Angle of Attack Sensing


78. Angle of attack sensors are installed to activate the aircraft stall warning device. The simplest of these is the Flapper Switch, or leading edge stall warning vane, and is fitted at the leading edge of the wing. It is located at, or just below, the stagnation point at normal flight angles of attack. It comprises a small vane protruding forward and connected to an electrical switch. As angle of attack increases, the stagnation point moves lower on the leading edge until it is below the flapper vane. The direction of the airflow relative to the vane has therefore changed from downward to upward and this deflects the vane upward, closing the electrical switch and completes a circuit to activate the stall warning in the cockpit. 79. Where the angle of attack sensor is required to activate a stick shaker or stick pusher, as well as a stall warning, a more precise type of sensor device is used. In most cases this consists of an aerodynamic vane protruding from the side of the fuselage near the nose, as shown at Figure 4-16. The vane is attached to the rotor of a synchro, which is free to rotate. The synchro and vane are initially aligned against reference marks on the fuselage representing the aircraft pitch attitude when the aircraft is clean with the wings at the most efficient incidence. The output from the synchro is transmitted to an electrical circuit controlling the stall warning systems. The circuit usually includes an input from the flap position monitoring system, to compensate for the pitch attitude change that occurs when flaps are deployed. The circuit is de-energised when the aircraft is on the ground.

Chapter 4 Page 35

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-16 Angle of Attack Vane

Chapter 4 Page 36

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Stick Shakers
80. The vane-operated synchro output circuit is programmed to detect not only incidence, but also rate of change of incidence. It feeds a signal to the stick shaker motor when either parameter indicates an approach to the stall. The motor drives an unbalanced ring attached to the control column. This causes the column to vibrate at a frequency similar to that caused by aerodynamic buffet. In many aircraft an audible warning accompanies the stick shaking, in case the aircraft is on autopilot at the time. The recovery procedure is to reduce incidence, preventing the stall from developing.

Stick Pushers
81. This somewhat controversial device automatically takes the recovery procedure described above, by pushing the control column forward if the stalling incidence is reached. Stick pushers are sometimes fitted in aircraft that are prone to what is known as the deep stall, or super-stall condition. Broadly speaking, these aircraft have swept back wings, high-speed wing sections and the T tail configuration. 82. The airflow following a stall in a conventional aircraft is illustrated at Figure 4-17. It can be seen that, although the air has broken away in a random manner from the upper surface of the wing, the tailplane and the elevators are still in undisturbed air. Consequently the tailplane will produce the lift required for stall recovery, especially with the elevator fully down.

Chapter 4 Page 37

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-17 Stall Conventional Aircraft

83. Figure 4-18 shows the airflow in the stall for a swept wing T tail aircraft. In this case the turbulent air completely covers the tailplane and elevators, making them virtually ineffective.

FIGURE 4-18 Deep or Super Stall

84. Additionally, the swept wings progressively stalling from tips to root causes a nose up pitching movement, which the ineffective back end is quite incapable of combating. 85. The problem is further aggravated by the fact that the aircraft will by now have acquired a significant downward vertical velocity, which will progressively increase the angle of attack beyond the stalling angle.

Chapter 4 Page 38

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
86. Finally, it is often the case with T tail aircraft that the engines are mounted on the rear fuselage. Now the chances are that the turbulent air entering the jet intakes will have caused a flameout. 87. Remember that the stick pusher, which is activated by an angle of attack sensor, is there to prevent the stall occurring in the first place.

Speed Margin
88. The margin between the stall speed and the limiting Mach number, known as the speed margin, is relatively small at the high altitudes that give optimum efficiency in terms of jet operation. Remember that as altitude increases the air temperature decreases, the local speed of sound decreases, and therefore the TAS corresponding to a given Mach number decreases. Conversely, since the stall speed in terms of IAS remains constant with change of altitude, as the aircraft climbs the TAS at which the stall occurs will increase. This situation is illustrated at Figure 4-19.

Chapter 4 Page 39

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-19 Effect of Altitude on the Speed Margin

89. Bear in mind that jet streams and associated clear air turbulence occur at or around the tropopause. It should then be obvious that a great deal of care must be taken to keep the cruising jet aircraft speed stable, and to prevent unchecked oscillatory motions about any of the aircraft axes.

Chapter 4 Page 40

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Canard Configuration
90. In the canard configured aircraft the stabiliser is mounted forward of the mainplanes (it is a tail first aircraft). Unlike the conventional configuration, in which only the mainplanes (wings) provide positive lift, in the canard both sets of surfaces provide positive lift, as shown at Figure 4-20. The stabiliser is at a slightly greater angle of incidence than the wing so, when the aircraft pitches up the canard foreplane produces a restoring nose-down pitching moment. Because the foreplane will reach the stalling angle of attack before the mainplane it has been claimed that the canard configured aircraft is unstallable, since the stalled foreplane will cause the nose to drop, automatically initiating stall recovery. Unfortunately this does not hold good during violent manoeuvres, when both planes may stall. With no effective control surfaces the aircraft is then uncontrollable and stall recovery may be impossible.

Chapter 4 Page 41

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-20 Canard Configuration

Chapter 4 Page 42

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling

Stall Recovery
91. Recovery from the stall invariably requires the angle of attack to be reduced, possibly to well below the stalling angle, in order to restore lift. During this process it is likely that a considerable amount of height will be lost.

Climbing and Descending Turns


92. In a turn the wing on the outside of the turn is travelling faster than the wing on the inside. In a climbing turn, although both wings are climbing at the same rate, the outer wing covers a greater horizontal distance than the inner wing, giving it a greater effective angle of attack. This is illustrated at Figure 4-21. Hence, in a climbing turn, the outer wing will reach the stalling angle of attack before the inner wing.

Chapter 4 Page 43

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-21 Climbing Turn

93. In a descending turn the wing on the outside of the turn travels faster, but the wing on the inside of the turn has the greater effective angle of attack and is therefore closer to the stalling angle. This is illustrated at Figure 4-22. In the event of a stall in a descending turn the aircraft is liable to roll further into the turn, which could then rapidly develop into a spin. 94. It should be appreciated that recovery from a stall where wing-drop has occurred will require more height for recovery since extra time is required to level the wings after reducing incidence. This will be more pronounced during recovery from a stall in a descending or climbing turn.

Chapter 4 Page 44

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-22 Descending Turn

The Spin
95. If a wing drop occurs at the stall, a spin may develop. The angle of attack of the dropping wing is increased well beyond the stall, due to the downward vertical velocity. Conversely the upgoing wing will experience a reducing angle of attack, and will be less deeply stalled, or even unstalled. Because of the almost non-existent lift experienced by the downgoing wing it will continue to drop, and any attempt to raise it using the ailerons will aggravate the situation by further increasing the angle of attack at the outboard end of the downgoing wing. 96. It does not, however, occur in isolation. At the same time as roll is occurring, yaw is also occurring due to the increased drag on the downgoing wing. This increased rolling and yawing motion is known as autorotation, often referred to as the incipient spin.

Chapter 4 Page 45

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
97. Once the rolling and yawing motions become self-sustaining, the aircraft is spinning. This is shown at Figure 4-23.

FIGURE 4-23 Autorotation Leading to the Spin

Chapter 4 Page 46

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
98. Recovery from the spin is aimed at reducing the rolling moment in the direction of the spin and/or increasing the anti-spin yawing moment. Of the two, the yawing moment is the more important, but the rudder (the normal yaw control) is not the only control by which the pilot may induce yaw. 99. The time-honoured sequence of control movements for spin recovery is idle power, the application of full opposite rudder, control column forward until the spin stops, maintaining the ailerons neutral. In certain aircraft the pilots handling notes may call for the use of aileron during spin recovery.

Ice and the Stall


100. Ice formation on the wings of an aircraft can have a dramatic effect on the airflow around the wing, its lift-producing capability, and its stalling angle of attack. This is illustrated at Figure 4-24.

Chapter 4 Page 47

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
FIGURE 4-24 Effect of Ice on the Lift Curve

101. Formation of ice on the leading edge of the wing can cause significant changes to the local aerofoil section, resulting in large changes in the local airflow and pressure gradients. Some forms of ice have extremely rough surfaces, which cause high surface friction and a marked reduction in boundary layer energy. The sum of these effects is a large increase in drag and a large decrease in the maximum lift coefficient ( C Lmax ) . This, plus the added weight due to the ice, will mean that greater power is required for a given speed and that the stalling speed is higher for a given weight.

Chapter 4 Page 48

G LONGHURST 1999 All Rights Reserved Worldwide

Stalling
102. The effect of a layer of frost on the upper surface of the wing (where it usually forms) is to significantly increase the surface roughness of the wing, increasing drag due to skin friction and reducing the boundary layer energy. Hence C Lmax is reduced, although not usually to the same extent as with ice formation. Nevertheless, the reduction may well bring C Lmax down to a value similar to that required for take-off. 103. The effect of ice and frost formations on take-off performance is of vital significance. The stalling angle of attack may be reduced by up to 25% and the stalling speed may be increased by a similar amount. To exacerbate the problem, the angle of attack sensor will continue to operate normally so the stall will occur before the stall warning can activate - in other words, there will be no stall warning. In addition, after recovery, speed must be increased to a value higher than normal and this will involve additional height loss. At the very time when the pilot is obliged to operate the aircraft close to the stalling speed, at high angles of attack, and close to the ground, ice or frost will narrow the safety margins to a fatal degree. The moral is clear, NEVER attempt to take off with ice or frost formation on the aircraft wing surfaces. 104. Finally, if ice forms on the wing surfaces it will almost certainly also form on the horizontal stabiliser surfaces, modifying the aerodynamic effect of the stabiliser and increasing the possibility of stabiliser stall. In this event the stall characteristics of the aeroplane may well be significantly altered, producing abnormal behaviour at the stall. 105. Airframe ice formation must always be considered when flying through cloud or rain at a temperature below 0C. The temperature range favourable for ice accretion in thunderstorms is from 0C to -45C, where water droplets can exist in a supercooled state. Below about -30C however, a large part of the free water content of the atmosphere normally consists of ice particles or crystals and snowflakes and chances of severe icing at these low temperatures are greatly reduced.

Chapter 4 Page 49

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Lift Augmentation
Trailing Edge Flaps Leading Edge Devices Boundary Layer Control Drag Augmentation

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Lift Augmentation

1. High lift devices are fitted to aircraft in order to reduce the take-off and landing distances required. They accomplish this by enabling the aircraft to achieve the required amount of lift at a lower airspeed, thereby reducing rotation and take-off safety speeds, and also approach speeds. In other words the co-efficient of lift is increased, thereby enabling a lower V in the lift formula, Lift = CLVS. The types of high lift devices considered in this chapter are trailing edge flaps, leading edge devices and boundary layer control devices. 2. In modern jet transport aircraft, with wing sections and planforms designed for high speed, high altitude flight, lift augmentation is essential at the relatively low speeds of approach and landing, take-off and climb.

Trailing Edge Flaps


3. It has already been shown that the lift coefficient of a wing is dependent, among other things, upon its camber. If the camber is increased the lift coefficient will increase, and this can be achieved in flight by means of a hinged flap that deflects downwards from the trailing edge of the wing. A type of flap that works in this fashion is illustrated at Figure 5-1 and is known as the Plain or Camber Flap.

Chapter 5 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

The Plain Flap


4. This type of flap functions purely by increasing the camber; there is no change in surface area of the wing when it is lowered. It produces a nose-down pitching moment due to aft movement of the centre of pressure and the increase in camber when it is lowered typically produces a 50% increase in lift at a given IAS. 5. The maximum lift coefficient that can be obtained with this type of flap is limited because the increased camber encourages boundary later separation over the upper surface of the wing, which reduces the stalling angle of attack (critical angle), when the flap is extended, to about 12. This can be a significant advantage during the landing approach, since less pitch up is required for a given value of CL, improving visibility ahead. 6. The plain flap considerably increases drag when it is fully lowered, which is desirable during approach and landing, but clearly not so during take-off and climb. This type of flap is best suited to aircraft that operate in the lower speed range. If the use of flap is required during take-off, it is usually limited to small deflections only.

FIGURE 5-1 Plain Flap

Chapter 5 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

The Split Flap


7. A split flap is illustrated at Figure 5-2. In this case the flap hinges down from the underside of the trailing edge of the wing and the effect is to increase wing camber much as does the plain flap. The increased camber occurs in the case of the split flap because the mean chord line of the aerofoil changes. 8. The split flap produces up to a 60% increase in lift, but at the expense of a large increase in drag. Its use is therefore mainly limited to approach and landing. As with the plain flap it causes a nose-down pitching moment, but the stalling angle of attack is only marginally reduced, to about 14. The split flap was popular with aircraft designers at the time of the second world war and for a short time afterwards. Its value was overtaken by the increase in aircraft size and the consequent need for flaps suitable for use during take-off.

FIGURE 5-2 Split Flap

Chapter 5 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
9. It is the loss of energy in the boundary layer that leads to separation when camber is increased by the use of plain or split flaps. If the boundary layer energy can be restored it follows that separation can be delayed and the wing, with flap extended, will be capable of maintaining lift at much greater angles of attack. In other words, the stalling angle will be increased. One way of achieving this is with the slotted flap.

The Slotted Flap


10. A slotted flap is shown at Figure 5-3. A small gap exists between the trailing edge of the wing and the leading edge of the flap. The effect of this is to allow air to flow from the high-pressure area beneath the wing to the low-pressure area above the wing, through the venturi passage produced by the gap. The venturi accelerates the air, and the additional kinetic energy thus produced entices the air to flow further back along the upper surface of the flap before separating. The result is that a slotted flap increases the maximum lift coefficient by about 65% and increases the stalling angle of attack by a degree or so. The drag produced when it is deployed is much less than with the previous two types, and so it is suitable for limited use during take-off as well as landing.

FIGURE 5-3 Slotted Flap

Chapter 5 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
11. Double (and even triple) slotted flaps may be used to permit slow, steep approaches at a reasonable aircraft attitude. A double slotted flap is shown at Figure 5-4. The considerable rearward movement of the centre of pressure when this type of flap is extended produces a marked nose-down pitching moment. This is usually countered with associated leading edge devices. A 70% increase of CLmax and a stalling angle of about 18 is achieved with a double-slotted trailing edge flap.

FIGURE 5-4 Double Slotted Flap

The Fowler Flap


12. A Fowler (extending) flap is illustrated at Figure 5-5. As well as moving downwards the Fowler flap moves rearwards and increases the wing area as well as the camber, producing an increase in CLmax of the order of 90%. The stalling angle of attack is virtually unchanged from that of the wing with flap retracted. Extending the Fowler flap produces a nose-down pitching moment.

Chapter 5 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
FIGURE 5-5 Fowler Flap

Chapter 5 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
13. The effects of the foregoing types of trailing edge flap on the Lift Coefficient (CL) Angle of Attack () graph is shown at Figure 5-6.

FIGURE 5-6 Effect of Trailing Edge Flaps on the Lift Curve

14. Figure 5-7 shows the variation in Coefficient of Lift (CL) with increasing trailing edge flap deflection.

Chapter 5 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
FIGURE 5-7 Increment in Coefficient of Lift with Trailing Edge Flap

It can be seen that the incremental increase in (CL) is initially linear with flap deflection but then the increase reduces quite noticeably as the maximum flap angle is approached. 15. Figure 5-8 shows the variation in Coefficient of drag (CD) with increasing trailing edge flap deflection.

Chapter 5 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
FIGURE 5-8 Increment of Coefficient of Drag with Trailing Edge Flap

Now we see that at the lower flap angles, the incremental increase in (CD) is small but it increases markedly as the flap angle is further increased. 16. Figure 5-9 is a graph showing the influence of these trailing edge flaps on the relationship between the Lift Coefficient (CL) and the Drag Coefficient (CD).

Chapter 5 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
FIGURE 5-9 Effect of Trailing Edge Flaps on C L CD Graph

Chapter 5 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Flap Asymmetry
17. Since the extension of trailing edge flaps produces such a marked increase in both lift and drag it is clear that, should the flaps deploy asymmetrically, a very large rolling moment would result. Under these circumstances the aeroplane would almost certainly become uncontrollable, because the flaps retract fairly slowly and, in the meantime, the rolling moment provided by maximum aileron deflection would most probably be insufficient to counter that of the asymmetric flaps. In view of this, many modern aircraft incorporate comparative sensing in the flap control system, which will not allow the flaps to deploy unless they are doing so symmetrically.

Take-Off
18. The use of trailing edge flaps for lift augmentation during take-off must by limited to the setting recommended in the aircraft operating manual. This will invariably call for an extension which is only a small proportion of the maximum flap extension available, and which gives the best lift/drag ratio other than with a clean wing, reducing the stalling speed and hence take-off speeds. This will, in turn, give the shortest possible take-off run. Deploying the flaps further than the recommended setting will result in a longer take-off run (albeit possibly shorter than with no flap), because of the large increase in drag. The drag penalty also means that use of flap for take-off will reduce the climb gradient (angle of climb making obstacle clearance after take-off more difficult). 19. When flaps are lowered the lift is increased, but so too is the drag. At those angles of attack which give the best lift drag ratio with a clean wing (flaps retracted) the drag increase when flaps are lowered is proportionately greater than the lift increase. Hence, lowering flap worsens the lift/drag ratio.

Chapter 5 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
20. From Figure 5-7 and Figure 5-8, the rate of increase of coefficient of lift (CL) is greatest during the first third of flap extension, whereas the rate of drag increase is least in that extension range. In other words, at low flap angles the lift/drag ratio is still favourable, whereas at high angles it is adverse.

Landing
21. For landing a large flap angle gives the required increase in lift, facilitating a lower nose attitude for a low speed of approach and the high drag permits a steeper approach without speed becoming excessive (the flap is also acting as an airbrake). This results in an improved view of the runway at a reduced approach speed. The high drag also has the advantage that it gives deceleration during the float period between round out and touchdown and subsequent reduction in the landing run.

Change in Pitching Moment


22. All trailing edge flaps produce an increased nose-down pitching moment when extended, due to the change in pressure distribution. The resulting increased pressure differential between upper and lower surfaces also increases upwash and downwash. The increased downwash may affect the airflow over the tailplane and, therefore, its longitudinal stabilising effect. Generally this gives an increased nose-up pitching moment. The amount by which the tailplane is affected will depend upon its size and position. Whether the trim change upon extending trailing edge flap is nose-up or nosedown will depend upon which is the dominant effect, the changed pressure distribution on the wing or the downwash on the tail. Leading edge devices tend to reduce the nose-down pitching moment of the trailing edge flaps.

Chapter 5 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Leading Edge Devices


23. At high incidence and low airspeed the boundary layer may separate from the leading edge, causing the wing to lose lift, and possibly stall. This is particularly so in the case of the thin, supercritical, aerofoil sections used on most modern transport aircraft. 24. To prevent leading edge separation, and to increase the stalling angle of attack, leading edge lift augmentation devices are often used. They may simply increase the wing camber, as in the case of Krueger flaps and leading edge droop, or they may re-energise the boundary layer as in the case of leading edge slats. 25. Leading edge devices tend to move the centre of pressure forward, consequently they tend to counter the nose-down pitching moment of trailing-edge flaps, by producing a nose-up moment.

The Krueger Flap


26. This is a hinged flap attached to the leading edge of the wing, as shown at Figure 5-10 in its retracted and extended positions. At small deflections it reduces lift, but when fully deflected it produces an increase in lift of about 50% and increases the stalling angle of attack to as much as 25. At full deflection it gives a nose-up pitching moment, but it should be noted that the pitching moment might vary considerably during extension.

Chapter 5 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
FIGURE 5-10 Krueger Flap

Variable Camber Flap (Drooped Leading Edge)


27. A drooped leading edge system is illustrated at Figure 5-11 in its retracted and extended positions. It acts in basically the same manner as the Krueger flap and achieves much the same improvements in lift and stalling angle. However, it has the advantage of progressively increasing lift as it is deployed, avoiding the trim changes inherent with the Krueger system. The major disadvantage with the drooped leading edge is the complexity of the engineering required to operate it.

FIGURE 5-11 Leading Edge Droop

Chapter 5 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Leading Edge Slats


28. Slats are lift augmentation devices in the form of small leading edge aerofoils. When deployed a slot is formed between the slat and the leading edge of the wing, through which air flows from the high-pressure zone beneath the leading edge. This re-energises the boundary layer over the upper surface, helping to prevent leading edge separation and delaying the point at which the airflow separates from the wing at high angles of attack. A leading edge slat is shown at Figure 5-12.

FIGURE 5-12 Leading Edge Slat

29. It is normal for slats to extend along the entire length of the wing. The effect of the slat on the co-efficient of lift is quite significant, being shown at Figure 5-13. Leading edge slats significantly increase the stalling angle of attack, hence their use on high performance fighter aircraft and where high incidence angles are required at low speeds (e.g. rotation for take-off). Slats may be deployed manually, or automatically as pressure reduces behind the leading edge with increased angle of attack.

Chapter 5 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
FIGURE 5-13 Effect of Slats on the Lift Curve

Chapter 5 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
30. In addition to the effect of the slot (already described) the high camber which occurs when the slat is extended is responsible for flattening the peak of the low-pressure envelope above the wing. This is illustrated at Figure 5-14. Because the adverse pressure gradient behind the point of lowest pressure is reduced with the slats open, separation of the smooth airflow from the upper surface of the wing is delayed, almost to the trailing edge.

FIGURE 5-14 Effect of Slats on the Pressure Distribution

31. In some relatively low speed aircraft the slats are permanently in the open position. It may be easier to think of this as a wing with slots cut through it just aft of the leading edge. With high-speed aircraft the slats lie tucked up against the leading edge of the wing during cruise, and are extended by the pilot for take-off and landing. Alternatively the position of the slats may be automatically controlled, being adjusted as and when the trailing edge flap settings are altered. Finally, control of the slats may be under the control of the pilot in normal circumstances, but may extend automatically as the stall sensor senses a dangerously high angle of attack, which is approaching the stalling angle of the clean wing. This latter method is often employed in aircraft designed to operate at low altitude and to climb away steeply, such as crop-dusters.

Chapter 5 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Lift/Drag Ratio
32. With trailing edge flaps it was seen that the increase in lift achieved is always outweighed by the increase in drag, hence any use of trailing edge flap reduces the lift/drag ratio as compared with the clean wing. Leading edge devices, and especially leading edge slots, incur a relatively small drag penalty for the increase in lift achieved and consequently the lift/drag ratio is improved at high incidence over that of the clean wing and thus the gradient of the CL -CD curve does not reduce as quickly as that for trailing edge devices.

Slat Asymmetry
33. With the type of slat that is selected for deployment by the pilot for take-off and landing it is clear that any asymmetry of deployment would induce a large rolling moment, which would be highly undesirable. In the case of this type of slat there is normally in-built control protection to prevent asymmetric deployment of the slats. 34. The automatically-deploying slat however, extending automatically when the wing reaches a high angle of attack, is capable of deploying asymmetrically if one wings incidence approaches criticality.

Boundary Layer Control


35. Increasing the lift generated by a wing of given section always involves some form of boundary layer control, either by altering the camber of the wing or by adding energy to the boundary layer. If the boundary layer can be made to remain attached to the wing surface for as long as possible, by keeping it laminar, then the lift coefficient will be increased and surface friction and form drag will be reduced.

Chapter 5 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
36. We have seen how slots at the trailing and leading edges can achieve this to some extent, but their effectiveness is largely dependent upon adequate airspeed and they invariably incur a drag penalty. There are alternative methods of boundary layer control which have proved effective, particularly on high-speed aircraft in which the weight and drag penalties of conventional flaps may be unacceptable.

Control by Suction
37. By applying suction to a porous or slotted area of the upper surface of the wing, at the point where the boundary layer is thickening due to the adverse pressure gradient, it is possible to draw off the sub-layer, leaving a thinner boundary layer. This layer has greater velocity, since the greatly retarded sub-layer has been removed. As a rule, distributing the suction over an area of porous panels has proved more effective than the use of slotted panels. The source of the suction is usually an engine-driven vacuum pump.

Control by Blowing
38. An alternative to adhesion of the boundary layer by suction is to eject high velocity air from small jets in the same direction as the boundary layer flow. This has the effect of speeding up the retarded sub-layer and re-energising the whole of the boundary layer. If this is combined with the use of trailing edge flaps it is possible to obtain a very high maximum lift coefficient (CLmax). Figure 515 shows a blown flap, which is capable of raising the CLmax from 1.5 for a plain aerofoil to about 5. If this is applied to the lift formula it will be seen that the lift increase, all other conditions remaining the same, is remarkable.

Chapter 5 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
FIGURE 5-15 Blown Flap

39. In addition to boundary layer control by blowing over the trailing edge flaps, it can be applied at the same time to the leading edge to increase lift even further. The practical limit for this type of lift augmentation is set by the power requirements for blowing or suction. Theoretically, at least, it is possible to reduce the landing speed of a high-speed swept wing aircraft by as much as 40% using these methods.

Vortex Generators
40. Vortices flowing chordwise across the upper wing surface can be induced either by small metal vanes projecting normal to the wing surface, or by small jets of air blowing normal to the surface. In either case the vortices carry the high velocity air near the upper part of the boundary layer to the slow-moving air close to the wing surface. This transfers energy to the sub-layer, maintaining boundary layer attachment towards the rear of the wing. There is a drag penalty associated with the vane-type generators, but this can be kept to a minimum with careful design. An advantage of the blown vortex generators is that they can be switched off when not required, so that the drag penalty is only incurred during the lift-augmented phases of flight.

Chapter 5 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Drag Augmentation
41. Under certain flight conditions it is often desirable to be able to increase the drag of the aircraft without, at the same time, increasing lift. The device designed to achieve this is the speed brake, or flight spoiler. On high performance sailplanes this usually takes the form of slotted plates which extend normal to the wing surfaces, creating drag to reduce the lift/drag ratio and steepen the glide angle. Similar devices are often fitted to the fuselages of fast military jet aircraft to prevent excessive speed in a dive and to control speed during manoeuvring. 42. There are also circumstances where it is an advantage to be able to destroy some of the lift generated by the wings, either on both wings simultaneously or only on one wing at a time. This is achieved by means of spoilers that normally lie flush with the wing upper surface, but are hinged at the forward edge to deploy progressively, as shown at Figure 5-16.

FIGURE 5-16 Spoiler Deployment

43.

Spoilers of this type are fitted to most large transport aircraft and serve a variety of functions.

Chapter 5 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Flight Spoilers
44. In order to establish the aircraft in the descent spoilers on both wings are deployed simultaneously and symmetrically. This destroys some of the lift being generated and creates drag, decreasing the lift/ drag ratio. Thus the aircraft is able to commence the descent in a controlled manner without the need to reduce speed for the deployment of flaps.

Ground Spoilers
45. After touchdown the aircraft is still moving at high speed and could conceivably become airborne again in the event of a gust. To "kill" the lift and hold the aircraft on the ground additional spoilers may be deployed.

Roll Spoilers
46. During the high-speed phases of flight the usefulness of ailerons for roll control is largely offset by their tendency to induce wing twisting, especially in swept wing aircraft. Because of this, spoilers are often used in conjunction with, or instead of, ailerons during high-speed flight. Under these circumstances they are deployed asymmetrically and in proportion to the demands from the pilots controls. The roll control system is necessarily complex in such aircraft, since it must be capable of selecting, for example, spoilers only at high speed and ailerons only at intermediate speeds. At low speed a combination of the two may be necessary to achieve adequate roll rates during landing and take-off. This is known as spoiler-mixer control.

Chapter 5 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
47. Figure 5-17 shows the positions of the wing-mounted spoilers on a Boeing 747. When flight spoilers are selected as speedbrakes the two inboard spoilers on each wing are deployed together with the inner two of the four outboard spoilers. Ground spoiler selection extends all the spoiler panels. Roll control uses all four outboard spoiler panels and the outer of the two inboard panels on either wing.

Chapter 5 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation
FIGURE 5-17 Spoilers -B747

Chapter 5 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Influence on the CL - Graph


48. The purpose of the spoiler is to reduce the lift/drag ratio by reducing lift and increasing drag. The effect on the graph of lift coefficient (CL) against angle of attack () of deploying spoilers is to reduce CL and therefore make the slope of the graph shallower, as shown at Figure 5-18. The pure speed brake is a drag-inducing device only, so it has little effect on the relation between CL and angle of attack.

FIGURE 5-18 Effect of Spoiler Deployment on the Lift Curve

Chapter 5 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

Lift Augmentation

Influence on the CL CD Graph


49. The spoiler reduces lift and increases drag, whilst the pure speed brake increases drag without necessarily reducing lift. In both cases the slope of the curve of CL against CD will be less steep when the device is deployed.

Chapter 5 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Control
Aircraft Axes and Planes of Rotation Interaction in Different Planes Aerodynamic Balance Trimming Pitch Control Yaw Control Roll Control Flutter Divergence Adverse Yaw Powered Controls

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Control

Aircraft Axes and Planes of Rotation


1. Figure 6-1 illustrates the three axes of rotation about which an aircraft moves in flight. It will be noted that they intersect at the aircraft centre of gravity (CG). This is always the case.

Chapter 6 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-1 Axes of Rotation

Chapter 6 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Control
2. The roll axis is the aircrafts longitudinal axis running horizontally fore to aft. It is so called because it is the axis about which the aircraft moves in roll, (left wing up, right wing down or vice versa) either in response to the pilot's roll controls or due to an external disturbance. When the aircraft moves about the roll axis it is rotating in the rolling plane, an imaginary vertical flat plane aligned with the aircraft's lateral axis. 3. The pitch axis is the aircrafts lateral axis running horizontally from side to side, intersecting the roll axis at the aircraft centre of gravity. It is the axis about which the aircraft moves in pitch (nose up or nose down) in response to the pilot's pitch controls or due to an external disturbance. When the aircraft moves about the pitch axis it is moving in the pitching plane, an imaginary vertical flat plane coincident with the aircrafts longitudinal axis. 4. The yaw axis is the aircrafts vertical, or normal, axis. It is so called because it is normal, or perpendicular, to the two horizontal axes and intersects both at the aircraft centre of gravity. It is the axis about which the aircraft moves in yaw in response to the pilot's yaw controls or due to an external disturbance. When the aircraft moves about the yaw axis (nose left or right) it is moving in the yawing plane, an imaginary horizontal flat plane coincident with the aircrafts roll and pitch axes. 5. The purpose of the flight controls is to enable an aircraft to be rotated about its three axes. The elevators achieve pitch control about the lateral axis. Control in roll about the longitudinal axis is by means of the ailerons and control in yaw about the normal axis by means of the rudder. These are known as the primary flight controls. They function by producing an aerodynamic force at a point distant from the aircraft CG, so as to create a moment about the required axis. 6. The flight control produces an aerodynamic force by changing the camber of the lifting surface to which it is attached, and/or by changing its angle of attack.

Chapter 6 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Camber Change
7. Controls usually take the form of hinged, flap type aerofoil sections mounted on the trailing edges of the wing, the tailplane and the fin. When they are moved they alter the effective camber of the section to which they are attached and therefore alter the amount of lift being generated. This type of control is illustrated at Figure 6-2. Within reason, controls are positioned as far as possible from the axis of rotation about which they are effective, so that they create the largest moment for the least amount of force.

FIGURE 6-2 Flap Type Control Surface

Angle of Attack Change


8. Changing the camber of the wing, tailplane or rudder by deflection of a hinged control surface also changes its effective angle of attack, which further changes the overall lift coefficient.

Chapter 6 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Interaction in Different Planes


9. It has been shown in the chapter dealing with stability that movement about one of the primary axes can induce a secondary movement about another axis. For example, if the rudder is moved to yaw the aircraft the wing on the outside of the turn is moving slightly faster than that on the inside. It therefore develops more lift and so the wing on the outside of the turn rises. Hence, the secondary effect of applying rudder is a roll into the turn. 10. If aileron is applied to roll the aircraft the resulting bank angle causes it to sideslip. The aerodynamic force acting behind the centre of gravity on the keel surfaces (fin, rudder, etc.) causes the aircraft to yaw towards the lower wing. The secondary effect of aileron is, therefore, yaw.

Aerodynamic Balance
11. However, when a control surface is deflected the airflow acting over it will try to return the control to the neutral position. The total force trying to return the control surface to the neutral position is the product of the lift force on the control surface and the perpendicular distance of the centre of pressure of the control surface to the hinge line. This is called the hinge moment and is illustrated at Figure 6-3.

Chapter 6 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-3 Hinge Moment

12. The magnitude of the lift force generated by any control surface will vary directly as the square of the EAS. The pilot is required to provide the force to overcome the hinge moment and deflect the control surface (in a manual system). At all but the lowest airspeeds he or she could do with some form of assistance. This assistance is supplied in the form of control balance devices. 13. Control balancing is achieved either by reducing the hinge moment, or by setting up a force that acts against the hinge moment.

Nose Balance or Inset Hinge


14. This is perhaps the simplest form of aerodynamic balance. The hinge is set back towards the CP of the control surface so that, when it is deflected, air strikes the surface forward of the hinge and reduces the force needed to move the control by partially balancing the aerodynamic force aft of the hinge.

Chapter 6 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Control
15. With this type of balance, care must be taken in the design to ensure that the centre of pressure is not too near the hinge line. When a control surface is operated its centre of pressure moves forward. If the margin between the centre of pressure and the inset hinge is too small there is a possibility that the CP will move forward of the inset hinge, reversing the direction of the hinge moment and is known as overbalance. An inset hinge is illustrated at Figure 6-4.

FIGURE 6-4 Inset Hinge

Horn Balance
16. An example of a horn balance is shown at Figure 6-5. Although it is shown here on a rudder, horn balances can equally well be used on ailerons or elevators. In this system a portion of the control surface acts ahead of the hinge line, and therefore produces a moment in opposition to the hinge moment.

Chapter 6 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-5 Horn Balance

Internal Balance
17. A projection of the control surface in the form of a balance panel, often referred to as a beak, is connected by a flexible diaphragm within a sealed chamber to a fixed structure (eg. spar). See Figure 6-6. Control surface movement produces a pressure differential between upper and lower surfaces and these upper and lower surface pressure changes are fed to the chamber to provide a partial balancing moment. Internal balance is therefore achieved with no increase in exterior drag.

Chapter 6 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-6 Internal Balance

Balance Tab
18. Like the inset hinge and the horn balance the balance tab serves the purpose of reducing the stick forces involved in moving a primary control surface, at a given airspeed. A balance tab is shown at Figure 6-7. As the primary control surface moves one way, so the balance tab moves the other. Since the tab is a considerable distance from the hinge line of the primary control, the moment produced by it is large. The balance tab imposes a small penalty in terms of drag, and diminishes slightly the effectiveness of the primary control surface to which it is fitted.

Chapter 6 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-7 Balance Tab

Chapter 6 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Anti-balance Tab
19. In some aircraft, far from requiring assistance in moving a control surface against the aerodynamic loads, the hinge moment is too small. This results in very low control column loads, a lack of feel and the possibility of over-stressing the airframe due to excessive deflection of the control surface. This often occurs because of the hinge being too close to the centre of pressure of the control surface. In order to improve the situation an anti-balance tab is fitted which operates in the same direction as the control surface. An anti-balance tab is illustrated at Figure 6-8.

Chapter 6 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-8 Anti-Balance Tab

Spring Tab
20. With many aircraft aerodynamic balancing is not considered necessary at low airspeeds, but is progressively required as airspeed increases, and with it the aerodynamic loads. The spring tab system may then be fitted to deal with this situation. A spring balance tab is shown at Figure 6-9.

Chapter 6 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-9 Spring Tab

Chapter 6 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Control
21. The movement of the control column is transmitted to a lever pivoted on the main control surface but not directly operating it. Operation of this surface is through springs, and with low aerodynamic loads the movement of this pivot arm is transmitted to the main control surface through the springs; consequently there is no alteration in the geometry between the primary control surface and the balance tab. When the aerodynamic loads increase at high speed, in order to transmit the control column movement via the pivot arm to the control surface, the spring becomes compressed. This upsets the geometry of the system and brings into operation the balance tab on the trailing edge, which moves in the opposite direction to the primary control surface, thus assisting the pilot by reducing the stick forces involved.

Servo Tab
22. When manual controls are used to operate very large control surfaces the loads involved, even with balance tab assistance, may be unacceptable. Under these circumstances servo tabs are used to operate the control surfaces. A servo tab is a small aerofoil section, once again attached to the trailing edge of the main control surface. The servo tab is operated directly by the control column, with no direct connection between the control column and the main control surfaces. As with the balance tab and the trim tab (described later), the servo tab moves in the opposite direction to the primary control surface. A servo tab is illustrated at Figure 6-10.

Chapter 6 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-10 Servo Tab

Chapter 6 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Trimming
23. There is sometimes confusion between balance tabs and trim tabs. Balance tabs serve to aerodynamically balance the forces acting on a control surface in order to reduce the force required to move it. Trim tabs serve to eliminate the force required to maintain a control surface at a particular deflection. 24. For aircraft in flight to be in equilibrium the moments about each of the three axes of the aircraft must balance. If they do not balance then an additional force must be supplied by deflection of the controls to keep the aircraft in equilibrium. The physical effort necessary to maintain this control surface deflection would place a great strain on the pilot, so to overcome this problem trim tabs are provided. 25. Consider the situation that exists in the case of an aircraft that tends to continuously fly nose up. In order to prevent this the elevator must be deflected downwards and maintained in this position. In order to achieve this the trim tab attached to the trailing edge of the elevator is deflected upwards as shown at Figure 6-11.

Chapter 6 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-11 Trim Tab

Chapter 6 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Control
26. The force required to maintain the elevator in the desired position is FP, and the work involved is FP x distance DP (the length of the arm from elevator hinge line to the elevator centre of pressure). 27. The action of the trim tab is to provide a much smaller force FT, but over a much larger arm DT (from the elevator hinge line to the trimmer centre of pressure). The work involved in holding the elevator out of the neutral position is thus done by the trim tab and requires no force from the pilots controls. 28. The trim tab slightly reduces the effectiveness of the primary control surface, since the primary control is already part way to full scale deflection with the aircraft in balanced flight. Also a trim drag penalty is suffered, hence the introduction of the variable incidence tailplane for high-speed flight, discussed later in this chapter. 29. Trim tabs may be controlled manually by trim wheels, often located on the centre console, or by trimming electric motors, the control for which is normally located on the control column. As with any control surface, the effect of any trimmer is proportional to the square of the airspeed. Whilst small movements of the trim tab will suffice at high speed, significantly larger trim movements are required at lower airspeeds, especially when significant changes of aircraft configuration, such as the lowering of flaps and undercarriage, are involved. On many light aircraft fixed trim tabs are fitted, which of course can only be adjusted on the ground.

Chapter 6 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Fixed Tabs
30. Fixed tabs are attached to the trailing edge of a control surface and can be set at a fixed angle when the aircraft is on the ground. In flight, under conditions of no force on the pilot's controls the position of the tab determines the trailing position of the primary control surface. Fixed tabs may be as simple as a small rectangle of aluminium attached to the control surface and manually bent to the required angle, which is found by trial and error. Some light aircraft still use a very early form of fixed tab comprising strips of cord stuck to one side of the trailing edge of the primary surface.

Stabiliser Trim
31. As IAS is increased the angle of attack required to maintain level flight decreases and so the elevator deflection necessary decreases. Hence less elevator trim is required for zero stick force as indicated airspeed increases. 32. From this it follows that, at the low speed of take-off, significant elevator up deflection is necessary to first initiate the take-off rotation and subsequently to maintain the necessary angle of attack until airspeed increases. In order to keep the pilots control force required within acceptable limits the stabiliser/elevator will need to be trimmed accordingly. 33. The further forward the CG of the aircraft the greater the elevator deflection necessary to overcome the inherent longitudinal stability of the aircraft, the further aft the CG the less the longitudinal stability and the more difficult the aircraft becomes to control in pitch. In the first case the pilot may run out of elevator up movement and be unable to maintain pitch attitude, in the second the aircraft may be too tender and impossible to control in pitch.

Chapter 6 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Control
34. It is clear that the CG position for take-off must be within well-defined limits and the elevator/stabiliser trim setting must be pre-set so as to give minimum stick force at the deflection required for take-off and initial climb. After all, the pilot has enough on his hands at this critical stage of flight and does not want to be fiddling with the trim wheel. 35. Consequently the take-off longitudinal trim setting for a given CG position is usually stated in the operating manual for the aircraft.

Pitch Control
36. Control of the aircraft about the lateral axis, or pitch control, is achieved with the elevator, which is deflected by moving the control column (stick) forward or backward. Forward movement of the stick deflects the elevator down, increasing the positive camber of the tailplane and creating an upward aerodynamic force. This produces a moment about the lateral axis of the aircraft through the CG, which pitches the nose down. In other words, the increase in tail lift coefficient as the elevators are moved downwards results in a reduction in the pitching moment coefficient (ie. a nose down pitching moment change) and hence a reduction in incidence. The aircraft lift coefficient is correspondingly reduced. Rearward movement of the stick deflects the elevator up, with opposite results. 37. The degree of control resulting from movement of the control column will depend upon the area of the control surface, the amount by which it is deflected, the IAS of the aircraft and the distance between the control surface and the aircraft CG (the moment arm). 38. The larger the surface area of the elevator the greater the aerodynamic force developed for a given deflection angle at a given airspeed.

Chapter 6 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Control
39. The further forward the CG the more stable the aircraft is longitudinally and the greater the elevator deflection necessary to overcome that stability. As airspeed is reduced an increasing amount of up elevator is needed to produce the same lift coefficient. If the CG were too far forward the pilot would run out of available elevator deflection at the low speeds required during take-off and landing approach. 40. Furthermore, as explained under stability in Chapter 8, the stabilising effect of the tailplane is greater during manoeuvre than in steady level flight, ie. manoeuvre stability is greater than static stability in level flight. Therefore, as in level flight, as the CG is moved forward, the greater the elevator deflection required to overcome the increased stability, however for a given CG position, the elevator deflection required is greater during manoeuvre than in steady level flight. 41. Consequently, the effectiveness of the elevators, or any other control surface for that matter, is mainly dependent upon the extent of deflection and the airspeed. The higher the airspeed the less the deflection necessary for a given attitude change. Control surface deflection required is inversely proportional to the square of EAS. 42. In the case of the elevators the effect of downwashed airflow from the wings must also be considered. As aircraft angle of attack changes so does the downwash angle. If the downwash strikes the tailplane surface, as it does in most aircraft with a fuselage mounted tailplane, a change in downwash angle alters the effective angle of attack of the tailplane and thus the lift it generates. This in turn changes the pitching moment produced by the tailplane and will require elevator deflection to maintain a given attitude. 43. There is an alternative type of flight control, in which the angle of attack, or incidence, of the whole lifting surface is changed, with no change in camber. Figure 6-12 illustrates an example in which the angle of attack of the tailplane is altered for trim changes, but which retains an elevator (camber change) for pitch control. This is referred to as a variable incidence tailplane.

Chapter 6 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-12 Variable Incidence Tailplane

44. Figure 6-13 illustrates an example in which there is no elevator and where the tailplane incidence is varied to provide pitch and trim control. This is known as a stabilator.

FIGURE 6-13 Stabilator

Chapter 6 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Yaw Control
45. Control about the normal, or yaw axis is achieved through movement of the rudder. The rudder is hinged to the rear spar of the fin and is connected to the pilots rudder pedals. Pushing on the right pedal deflects the rudder to the right, which alters the camber and effective angle of attack of the fin such that an aerodynamic force to the left is produced. This creates a yawing moment about the aircrafts normal axis through the CG, which yaws the nose to the right. Pushing on the left rudder pedal has the opposite effect. 46. The effectiveness of the rudder increases with increasing airspeed. At low speeds, such as during the landing approach, a large deflection may be necessary to produce a given amount of yaw. At higher speed only a small deflection is necessary to achieve the same effect. 47. In aircraft designed for high subsonic or transonic flight, but which must of course be capable of controlled flight at low speeds, a pedal/rudder ratio changer is employed. As airspeed increases the maximum obtainable rudder deflection angle is automatically decreased by the ratio changer. The deflection decrease may be progressive, as illustrated in the graph at Figure 6-14, or it may be in steps, associated with flap retraction for example.

Chapter 6 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-14 Maximum Rudder Deflection Available Against Speed

ff

Chapter 6 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Control
48. In many aircraft, although by no means all, the ratio changer affects extent of rudder pedal movement so that decreasing rudder deflection possible is matched by corresponding decreasing pedal movement available. 49. In addition to the yawing moment due to rudder deflection there is also a yawing moment due to engine thrust. The direct yawing moments are due to the force acting normal to the propeller plane or the jet engine inlet. The further forward the engines are located the more destabilising the yawing moments due to thrust. 50. The indirect yawing moments due to thrust arise from the propeller-induced velocities and airflow at the vertical tail, which can produce significant directional trim changes. The indirect or induced moments in jet-powered aircraft are negligible. 51. In multi-engine aircraft there is a strong yawing moment in the event of asymmetrical power due to failure of one engine. The further the engines are located outboard of the CG the greater the yawing moment developed. In propeller-powered aircraft there is little choice other than to mount the engines on the wings, and the diameter of the propeller disc inevitably means that they must be well clear of the fuselage. 52. The unbalanced thrust and drag in the event of asymmetrical power produces a yawing moment that must be opposed by deflection of the rudder. The side force at the vertical tail must produce a moment equal and opposite to the yawing moment due to the unbalanced thrust. 53. The yawing effect of asymmetric power will be greatest at low airspeed and high power, such as would be the case in the event of loss of an engine shortly after take-off. The rudder must therefore be capable of maintaining directional control at a predetermined minimum speed with a critical engine out. Figure 6-15 shows in graphical form the determination of minimum directional control speed.

Chapter 6 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-15 Minimum Directional Control Speed

ff

54. The horizontal line on the graph shows the yawing moment coefficient (Cn) obtained from maximum rudder deflection at any speed. The curved line shows the yawing moment coefficient due to asymmetric thrust with one engine out and the remaining engine (or engines) at maximum power. Clearly, when Cn due to asymmetric power exceeds Cn due to maximum rudder deflection, the rudder is incapable of maintaining directional control. Minimum directional control speed is therefore at the point where the two curves intersect.

Chapter 6 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

Control
55. The minimum control speed in the air, in the take-off configuration, is known as VMCA. It is the minimum speed at which it is possible to suffer a critical engine failure and maintain control of the aeroplane within defined limits. 56. The minimum control speed on the ground is known as VMCG. It is the minimum speed on the ground at which it is possible to suffer a critical engine failure on take-off and maintain control of the aeroplane with the use of primary aerodynamic controls alone within defined limits.

Roll Control
57. Control about the longitudinal, or roll axis of the aircraft is achieved in most aircraft through the ailerons. The ailerons are hinged to the rear spar of the wing and are deflected by rotation of the control wheel (mounted on the control column) or by sideways movement of the control column (stick). 58. Suppose the control wheel is rotated to the right. The aileron on the right wing moves up to create a downward force, whilst the aileron on the left wing moves down to create an upward force. The result is a moment about the longitudinal axis, through the CG, rolling the aircraft to the right. Rotating the control wheel to the left would produce the opposite effect. 59. As with the elevators and the rudder, the amount of deflection is directly proportional to the amount of movement of the pilots control. The greater the aileron deflection the greater the rate of roll that will result, the rate of roll being the time taken to accelerate through a given angle of roll. The rolling motion is opposed by aerodynamic damping in roll, described in Chapter 8 on Stability, so that a steady rate of roll is quickly established for any given aileron deflection.

Chapter 6 Page 27

G LONGHURST 1999 All Rights Reserved Worldwide

Control
60. Clearly the rate of roll achieved for any given aileron deflection will depend upon the surface area of the ailerons, the airspeed and the distance from the ailerons to the centreline, or longitudinal axis, of the aircraft. Aileron size is limited by a number of factors, the torsional stiffness of the wings and the drag effects of the ailerons being the principal considerations. At low airspeeds then, it is desirable for the ailerons to be mounted as far outboard as possible, to achieve the maximum rolling moment. 61. However, at high airspeeds the aerodynamic force generated by aileron deflection can easily be sufficient to cause the wing supporting it to twist. Remember, the force generated by the aileron increases as the square of the EAS. The effect of wing twist is to minimise or even reverse the effect of the ailerons and this is illustrated at Figure 6-16 and is fully explained under Flutter.

FIGURE 6-16 Effect of Wing Twist

Chapter 6 Page 28

G LONGHURST 1999 All Rights Reserved Worldwide

Control
62. One way to overcome the problem of wing twist is to fit two sets of ailerons. Conventional ailerons are fitted towards the outboard end of the trailing edge in the normal manner, for low speed flight. A second set of smaller ailerons is fitted towards the inboard end of the wings, for high-speed flight. At high speeds the outboard ailerons are locked into the neutral position and the inboard ailerons are employed. The high speed offsets the small size of these surfaces in producing the necessary control forces, whilst the thicker and more rigid wing structure at the inboard end resists the aileron imposed forces which are trying to induce twist. 63. Spoilers may be used as well as, or instead of, ailerons during the high speed phases of the flight. The effect of opening the spoiler (by a very small amount) is to cause a loss of lift, and therefore roll in the direction of the raised spoiler. The spoiler does not introduce any significant twisting force. Roll spoilers are fully covered under Drag Augmentation in Chapter 5.

Flutter
64. Figure 6-17 illustrates how a wing may twist in torsion and/or bend in flexure.

Chapter 6 Page 29

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-17 Wing Twist and Bend

Chapter 6 Page 30

G LONGHURST 1999 All Rights Reserved Worldwide

Control
65. The torsional or elastic axis is the line about which the wing will twist if a force is applied to the wing, other than on the line of the axis itself. A force applied on the line will not cause the wing to twist. However, the wing may bend or flex under this force. Thus the torsional axis is an important feature of the wing structure and is the axis about which the wing may twist in torsion or bend in flexure. 66. Another form of vibration is caused by a control surface itself vibrating in the airflow, caused by incorrect balancing or slackness in the control runs. Torsional and flexural vibrations are, in themselves, fairly harmless and are usually damped out by the rigidity of the airframe structure. However, when acted upon by an external force, the airflow, further reactions will occur which may lead to structural failure. 67. Flutter may cause structural failure. It is a violent vibration caused by the interaction of the mass of aerofoil surfaces and aerodynamic loads. Three forms of flutter affect the wing: (a) (b) (c) torsional flexural flutter torsional aileron flutter flexural aileron flutter.

68. Torsional Flexural Flutter occurs as a result of the wing flexing and bending when aerodynamic loads are applied. Figure 6-18 shows the sequence of events.

Chapter 6 Page 31

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-18 Torsional Flexural Flutter

Chapter 6 Page 32

G LONGHURST 1999 All Rights Reserved Worldwide

Control
(a) The wing is in steady horizontal flight with the torsional axis ahead of the centre of gravity. Lift L is balanced by the reaction R caused by the bending of the wing due to the weight of the aeroplane. A disturbance causes the incidence to be momentarily increased, increasing lift. L is now greater than R and the wing flexes upwards. The C of G will lag behind the torsional axis due to inertia and this causes a further increase in incidence, further increasing lift. The wing reaches its aeroelastic bending limit due to inherent stiffness and this brings the torsional axis to rest, but inertia causes the C of G to travel further, decreasing incidence. L is now less than R and the wing starts to descend. The wing stops moving down when it again reaches its aerolastic bending limit, bringing the torsional axis to rest, but inertia causes the C of G to travel further, increasing the incidence. L is now greater than R and the cycle repeats.

(b)

(c)

(d)

The full cycle is shown at Figure 6-19.

Chapter 6 Page 33

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-19 Torsional Flexural Flutter Cycle

69. Torsional flexural flutter can be prevented in the design, either by ensuring that the wing is sufficiently stiff so that the critical flutter speed is far in excess of the maximum speed limit, or by ensuring that the centre of gravity of the wing is on, or ahead of, the torsional axis. 70. Torsional Aileron Flutter is caused by the wing twisting because of aerodynamic loads produced by aileron deflection. The sequence is shown at Figure 6-20, which shows half a cycle.

Chapter 6 Page 34

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-20 Torsional Aileron Flutter

Chapter 6 Page 35

G LONGHURST 1999 All Rights Reserved Worldwide

Control
(a) (b) The aileron is deflected down, with a resulting increased lift force on the aileron hinge. The wing twists about the torsional axis. The trailing edge rises, taking the aileron with it. Because the aileron C of G is behind the hinge, the aileron lags behind this upward movement, generating an even greater upwards force from the aileron thus increasing the twisting moment. The torsional reaction of the wing arrests the twisting motion but the aerodynamic loads on the aileron, together with its upward momentum and the stretch of the control runs, cause it to continue upwards, now placing a downwards load on the wing trailing edge. The energy stored in the twisted wing and the aileron aerodynamic load cause the wing to twist in the opposite direction, and the cycle is repeated.

(c)

(d)

71. This type of flutter can be prevented either by mass-balancing the ailerons (see next section), or by making the controls irreversible. Aircraft with fully powered controls and no manual reversion do not require mass-balancing. 72. Flexural Aileron Flutter is similar to the previous phenomenon but is caused by the movement of the aileron lagging behind wing upwards and downwards movement as it flexes, This tends to amplify the oscillation. The sequence is shown at Figure 6-21.

Chapter 6 Page 36

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-21 Flexural Aileron Flutter

Chapter 6 Page 37

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-22 Mass Balancing

73. This type of flutter can be prevented by mass-balancing of the aileron as shown in Figure 6-22. The positioning of the weight is important. The nearer the wingtip the smaller is the weight required. Sometimes the weight is distributed along the whole length of the aileron in the form of a leading-edge spar, thus improving the stiffness of the aileron and preventing a concentrated weight starting torsional vibration in the aileron itself. 74. It should be appreciated that tail surfaces may also experience flutter and, unless fullypowered, will also require to be mass-balanced. Tailplanes and elevators tend to be stronger in torsion than wings because of their smaller dimensions, but large fins as found on large transport aeroplanes may be less torsionally strong than the remainder of the empennage.

Divergence
75. Lack of torsional rigidity in a wing can, in extreme cases, cause divergence, as shown at Figure 6-23.

Chapter 6 Page 38

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-23 Divergence

Chapter 6 Page 39

G LONGHURST 1999 All Rights Reserved Worldwide

Control
76. If the incidence of a wing is increased momentarily by a disturbance, lift will increase, moving the centre of pressure forward. If the torsional axis is behind the centre of pressure the lift increase and forward movement will increase the moment which is twisting the wing. Conversely, if the disturbance decrease the incidence, the reduction in lift and aft movement of the centre of pressure will further reduce the incidence. In both cases the twisting action is resisted by the torsional stiffness of the wing. However, because lift increases with the square of the speed and torsional rigidity remains constant, there is a critical speed (the divergence speed), beyond which the aerodynamic moment will overcome the resistance of the wings torsional rigidity. In such a situation the wing will twist until structural failure occurs. This may be avoided by making the wing sufficiently stiff in torsion so that the divergence speed is well beyond the aircraft speed limit, or by designing the wing such that its torsional axis is ahead of the aerodynamic axis, thus preventing divergence at any speed.

Adverse Yaw
77. When an aircraft is rolled the downgoing wing experiences an increased effective angle of attack, whilst the upgoing wing experiences a decreased angle of attack. The inclination of the lift vector consequently produces a forward component on the downgoing wing and a rearward component on the upgoing one. Thus there are yaw moments acting about the normal axis in opposition to the yaw required from the roll direction. This is known as adverse yaw. 78. In order to achieve a rate of roll one of the ailerons must be lowered and the other one raised. The induced drag created by the lowered aileron is much greater than that caused by the raised one which produces a yawing moment in the opposite direction to the intended turn. This effect is called adverse aileron yaw, and is most marked when aileron deflections are large, which usually occurs at low airspeeds. 79. There are a number of devices designed to minimise or overcome adverse aileron yaw.

Chapter 6 Page 40

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Frise Ailerons
80. With this type of aileron it is arranged that the forward part of the up-going aileron will protrude into the airflow below the wing as shown at Figure 6-24. In this way drag is generated at the wing which is on the inside of the turn to match the drag of the downgoing aileron on the outside of the turn, thus preventing yaw. It is not the most efficient solution however, since it adds profile drag to the upgoing aileron to balance the induced drag of the downgoing aileron, rather than reducing the drag of the downgoing aileron.

FIGURE 6-24 Frise Ailerons

Chapter 6 Page 41

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Differential Aileron Deflection


81. Differential ailerons are designed to minimise adverse aileron yaw by matching the drag generated by the upgoing and downgoing ailerons by differential deflection. It is the downgoing aileron which produces the greater drag per degree of deflection, so the amount by which it is deflected for a given control wheel movement is less than that of the upgoing aileron. Hence, both ailerons generate equal drag and adverse yaw is averted. Compared to Frise ailerons the total drag during aileron deflection is said to be less. The principle is illustrated at Figure 6-25.

FIGURE 6-25 Differential Ailerons

Rudder/Aileron Coupling
82. In some aircraft the rudder is coupled to the ailerons by gearing or spring so that deflection of the ailerons also moves the rudder a proportionate amount to oppose adverse yaw.

Chapter 6 Page 42

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Roll Spoilers
83. The use of spoilers rather than ailerons in high-speed flight has already been mentioned. These devices are also very useful in overcoming the problem of adverse yaw at low speeds, when its effects are most noticeable and hazardous. The spoilers operate in conjunction with the ailerons. The wing on which the aileron goes down experiences an increase of lift and on this wing the spoiler remains retracted flush with the wing surface. The wing on which the aileron goes up suffers a loss of lift and an increase in drag by the extension of the spoiler, so this wing drops in the required direction of roll and the drag induces yaw in the required direction also. 84. low. Another advantage of spoilers is that the control forces needed to operate them are relatively

Slot/Aileron Coupling
85. A method that has been used with great success on light aircraft required to execute sharp manoeuvres at low airspeed and altitude is to place slots that open automatically, or in conjunction with aileron down deflection, at the leading edge of the wing immediately ahead of the ailerons. This system is illustrated at Figure 6-26.

Chapter 6 Page 43

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-26 Slot Aileron Coupling

Effect of Propeller Slipstream


86. Propeller rotation imparts a rotation to the slipstream in the same direction and which in turn produces an asymmetric flow over the fin. This induces a side aerodynamic force causing the aircraft to yaw which may, depending on the direction of propeller rotation, reduce the adverse aileron yaw effect.

Powered Controls
87. The force needed to move the control surfaces of a large aircraft flying at high speeds makes it virtually impossible for satisfactory control to be exercised by manual controls. It is therefore necessary that the primary control surfaces be power operated. 88. Powered controls may be divided into two categories, power-assisted controls and fully power operated controls.

Chapter 6 Page 44

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Power Assisted Controls


89. With power-assisted controls, the force needed for movement of the control surface is provided partly by the physical force produced by the pilot and partly by the power system. Here then the pilot will have feel which is provided by the control surface loads. Should a fault or power failure occur in the control system, a disconnect system will be available, and control will continue to be maintained by manual means alone, although control loading will be relatively high. Trim control for a power-assisted control is provided in the same way as that for manually operated flying controls.

Fully Powered Controls


90. Where fully power operated controls are installed, power systems are provided which, while independent of each other, operate in parallel and provide all the force necessary for operation of control surfaces. Movement of the pilots controls is transmitted to actuators, which provide the force necessary to move the control surfaces. 91. Safeguards against faults or power failure may be provided by manual reversion, or by more than one system, each with its individual hydraulic circuit. Thus conventional trailing edge tabs will not be included and trim would be obtained by altering the zero positions of the artificial feel mechanism. Occasionally, where necessary, balance tabs are fitted for the maintenance of servo and hinge loads. An example of a power operated control system is shown at Figure 6-27.

Chapter 6 Page 45

G LONGHURST 1999 All Rights Reserved Worldwide

Control
FIGURE 6-27 Powered Flying Control System

92. Power operated controls are irreversible, which means that there is no feedback of aerodynamic forces from the control surface. Consequently, the pilot has no feel through the controls for the aerodynamic loading on the control surfaces. 93. Feel is provided by artificial methods and, in fact, the artificial feel rarely has any direct relationship to the forces working on the control surface. Feel can be provided by a spring that exerts a constant load for a given control position, so that the more the control is moved the greater the spring force to be overcome. The disadvantage of this system is that the resistance to control movement is the same irrespective of airspeed. A proportionate force at low airspeed would be inadequate at high speed, alternatively a proportionate force at high airspeed would be too great at low speed.

Chapter 6 Page 46

G LONGHURST 1999 All Rights Reserved Worldwide

Control
94. A much more satisfactory arrangement is to provide loading which varies in direct relationship to airspeed. This loading is a force proportional to dynamic pressure (q). The arrangement is therefore commonly referred to as q feel. Pitot and static pressure is sensed and applied to the pilots controls to produce a force resisting movement proportional to airspeed and therefore representative of the force the pilot would feel with manual controls. In its simplest form a q feel system comprises a large piston with pitot pressure applied to one side and static pressure to the other. The overall force acting on the piston is thus due to dynamic pressure. The piston is connected to the pilots controls as shown at Figure 6-28.

FIGURE 6-28 Simple q Feel

Chapter 6 Page 47

G LONGHURST 1999 All Rights Reserved Worldwide

Control
95. The bulkiness of the simple q feel system can be overcome by the utilisation of a hydraulically enhanced system in which the pitot and static pressures are fed to either side of a diaghragm attached to a hydraulic servo-valve. The servo-valve provides a metered hydraulic pressure, which is an amplified value of dynamic pressure, to a small jack which opposes control column movement. An example of a hydraulic q feel system is shown at Figure 6-29.

FIGURE 6-29 Hydraulic q Feel

Chapter 6 Page 48

G LONGHURST 1999 All Rights Reserved Worldwide

Control

Fly-By-Wire
96. In many present-generation aircraft signals from the pilot's primary controls are transmitted electrically to actuators which move the control surfaces. This is known as fly-by-wire. The system lends itself to the incorporation of sophisticated electronic processing which alters the response to control inputs by the pilots to avoid stalling, over-rapid or excess control surface movement, or unstable flight regimes. Fly-by-wire not only has the capability to improve aircraft performance, efficiency and safety, it can also incorporate co-ordination of control surface movements too complex for a pilot to achieve unaided.

Chapter 6 Page 49

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Forces in Flight
Forces in Level Flight Forces in a Steady Climb Forces in a Steady Descent Forces in a Steady Glide Forces in a Steady Co-ordinated Turn Asymmetric Thrust

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight

Forces in Flight

Forces in Level Flight


1. The four forces acting on an aircraft in straight and level flight are lift, weight, thrust and drag. Lift acts through the centre of pressure and weight through the centre of gravity. For simplicity, thrust and drag forces are considered as acting parallel to the longitudinal axis, and their displacement from this axis depends on the design of the aircraft, high wing or low wing, the position of the engine(s), and so on. 2. Providing that the centre of gravity and the centre of pressure are not coincident a force couple will be set up by the lift and the weight forces, and this will result in a pitching moment, as shown at Figure 7-1.

Chapter 7 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-1 Lift/Weight Pitching Moment

3. The magnitude of the pitching moment will depend on the magnitude of lift and weight forces, but also on the distance between the centre of gravity and the centre of pressure. 4. The position of the CG will depend on the way in which the aircraft is loaded, and on the way in which fuel is transferred/consumed in flight. The position of the centre of pressure depends on the angle of attack, with the CP moving slowly forward as the angle increases, and then rapidly backwards at the stalling angle.

Chapter 7 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
5. The couple between the thrust and drag forces may compliment the lift/weight pitching moment as shown at Figure 7-2, or oppose it as shown at Figure 7-3.

FIGURE 7-2 Lift/Weight and Thrust/Drag Pitching Moments - Complimentary

Chapter 7 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-3 Lift/Weight and Thrust/Drag Pitching Moments - In Opposition

6. With the couples arranged as shown at Figure 7-2 it would obviously be necessary for the tailplane to produce negative lift, in order to maintain level flight. 7. A change in any one of the forces will disturb the trim of the aircraft, causing it to pitch either nose up or nose down. For example, if the thrust force is reduced by throttling back the engine, it can be seen from Figure 7-3 that the lift/weight couple is now greater than that of the thrust/drag causing the aircraft to pitch nose down. To maintain balanced level flight therefore, the elevators would have to be raised to produce a down-load at the tailplane.

Chapter 7 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
8. One important consideration in respect of the lift/weight force couple is the way in which the aircraft behaves in the event that the thrust/drag couple is destroyed, in other words during engine failure with a single engine aircraft. With the centre of gravity forward of the centre of pressure, the natural tendency will be for the nose to drop when the thrust/drag couple is removed as the engine fails, and this is desirable.

Forces in a Steady Climb


9. During a climb the aircraft gains potential energy by virtue of increased altitude. This is achieved either by increasing the thrust above that required for level flight at a given speed, or by using the aircraft's kinetic energy (not increasing power and accepting a loss of airspeed) or a combination of the two. 10. During the climb the lift continues to act at right angles to the flight path and the weight vertically downwards, however the two are now no longer directly opposed, see Figure 7-4. The weight must now be resolved into two components, that which is supported by lift, and that acting in the opposite direction to the flight path, i.e. in the same direction as drag. Therefore, for equilibrium, resolving the forces along the flight path, T = D + W Sin = and resolving at right angles to the flight path, L = W cos 11. Consequently the lift required for the same weighted aircraft is less than was needed for level flight, since the lift supports only that component of the weight given by W Cos , where is the angle of climb. However, the thrust available now has to match not only the drag, but also the component of the weight given by W Sin at Figure 7-4.

Chapter 7 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-4 Forces in a Steady Climb

Chapter 7 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
12. From Figure 7-4 it can be seen that: Thrust Drag Sin = --------------------------------Weight 13. Referring now to the triangle of forces diagram at the bottom of Figure 7-4: Rate of Climb Rate of Climb Sin = --------------------------------- = --------------------------------V Climb Speed Therefore: Therefore: and Rate of Climb Thrust Drag --------------------------------- = ---------------------------------V Weight Thrust Drag Rate of Climb = V --------------------------------Weight

Power Available - Power Required Rate of Climb = ----------------------------------------------------------------------------------Weight Therefore: Excess Power Rate of Climb = ---------------------------------Weight

14. This proves that, for a specified weight and at one optimum TAS the greater the margin between power available and power required the greater would be the rate of climb. The best rate of climb will be achieved when the power margin is greatest.

Chapter 7 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
15. From the above equation it can be seen that the rate of climb varies inversely with aircraft weight. The best rate of climb for a given power margin will be less if aircraft weight is increased, and vice versa. Reference to the triangle of forces diagram at Figure 7-4 will show that a reduction in best rate of climb involves an increase in TAS for best rate of climb. 16. Similarly as altitude is increased the IAS needed to achieve the optimum TAS will decrease, due to decreasing air density. 17. The graph at Figure 7-5 shows the curves of power available against power required for a piston-engined aircraft.

Chapter 7 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-5 Best Rate of Climb - Piston Aircraft

18. The power available curve shows up the inefficiencies of the propeller at low airspeed (ineffective angle of attack) and at high airspeed (compressibility losses and high drag losses). 19. The power required curve shows a disproportionate requirement for power at low speeds in order to overcome the high induced drag factors. 20. Clearly, power available must exceed power required in order to maintain a steady climb.

Chapter 7 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
21. The graph at Figure 7-6 shows the curves of power available against power required for a jet engined aircraft. The big difference here is that the power available curve is more or less constant, since the thrust available from a jet engine at a given altitude remains fairly constant, regardless of speed.

FIGURE 7-6 Best Rate of Climb - Jet Aircraft

Chapter 7 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
22. The effect of altitude on the best rate of climb speed is shown for a piston-engined aircraft at Figure 7-7 and for a jet engined aircraft at Figure 7-8. In both cases the power available diminishes with increasing altitude (the powerplants become less efficient in low-density air). Equally the power required increases with altitude, and so eventually an altitude will be attained where any further climb is impossible. This is the absolute ceiling. The service ceiling is the altitude at which the rate of climb has dropped to 100 feet per minute in a piston-engined aircraft and 500 feet per minute in a gas turbine powered aircraft.

FIGURE 7-7 Effect of Altitude on Best Rate of Climb - Piston Aircraft

Chapter 7 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-8 Effect of Altitude on Best Rate of Climb - Jet Aircraft

23. The best angle of climb will be achieved at a speed where the margin between thrust and drag is the greatest. Reference to the diagram at Figure 7-4 will establish that the climbing angle () will be greatest when the margin between thrust and drag is greatest, since:

Chapter 7 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
Thrust Drag Sin = --------------------------------Weight

Forces in a Steady Descent


24. The forces acting on an aircraft in a straight steady descent (constant speed and angle of descent) are as shown in Figure 7-9.

FIGURE 7-9 Forces in a Steady Descent

Chapter 7 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
Assuming that thrust acts along the flight path, resolving along the flight path, T + W Sin = D or T = D - W sin

25. So, drag is balanced by thrust and the component of weight acting along the flight path and is therefore, greater than thrust. Now, resolving at right angles to the flight path, L = W cos and consequently, as in climbing flight, the lift required for the same weighted aircraft is less than was needed for level flight since in the descent lift supports only that component of the weight acting perpendicular to the flight path given by W cos .

Emergency Descent
26. The situation most likely requiring a rapid descent from altitude is a rapid loss of cabin pressurisation. If this were to be caused only by a failure of the pressurisation system itself, without being caused by any structural failure (such as window failure or pressure bulkhead failure), then the aircraft may be descended at the maximum possible rate.

Chapter 7 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-10 Rate of Descent

From Figure 7-10, Rate of Descent = V=sin From Figure 7-9, (D T) Sin === -----------------W therefore, rate of descent will be maximum when V(D T ) ---------------------- =is maximum, W ie. increase airspeed, increase drag, reduce thrust.

Chapter 7 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
27. This will be acheived with idle thrust, maximum speedbrake, and at limit speed values. Whilst this is the fastest method of losing height, it must be accomplished with care because of the high speeds involved. If there has been some form of structural failure, then it is not advisable to use this method since the high values of dynamic pressure involved may cause further failure, perhaps catastrophically. Therefore it will be necessary to descend at a lower speed. Whilst this has the disadvantage of incresaing the time required to get down, it is more easy to fly since the aeroplane will not be at limiting speeds, although if flap is used to increase drag then flap limit speeds must be observed.

Forces in a Steady Glide


28. The forces acting on an aircraft in a straight steady glide (no thrust) are as shown in Figure 7-11 from which it can be seen that the weight is balanced by the resultant of the lift and drag.

Chapter 7 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-11 Forces in a Steady Glide

29. To overcome the drag and maintain speed in the glide, the source of energy is no longer the engine but the aircrafts potential energy (ie. altitude). 30. Thus, resolving at right angles to the flight path, L = W cos =

Chapter 7 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
and along the flight path, D = W sin . 31. From Figure 7-12, triangles A formed by lift, drag and total reaction and B, formed by range, height and glidepath are geometrically similar and for maximum distance (range) gliding angle () must be minimum, Drag ie. Tan = -----------Lift or
min

Lift -----------Drag

max

32. Therefore the minimum glide angle is achieved at an angle of attack which gives the best L/D ratio and thus for maximum range, the aircraft should be flown for minimum drag. Furthermore, if trailing edge flaps were extended during a glide descent, the glide angle would increase due to the reduction in the lift/drag ratio. 33. The gliding range can be calculated by reference once more to Figure 7-12 from the two similar triangles A and B, Range ----------------Height Lift - . = -----------Drag

Chapter 7 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-12 Forces in a Steady Glide

Therefore, L Range = --- x Height. D

Chapter 7 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
34. On the other hand, for minimum rate of descent, ie. maximum endurance, we know that V Sin = must be a minimum. But, from paragraph 28, D = W Sin ,=========therefore, D -. Sin === ---W

DV - where (D x V) is power required. Therefore, V Sin === --------W Thus, maximum gliding endurance (for a given weight) is achieved at the speed where power required is a minimum as shown at Figure 7-13.

Chapter 7 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-13 Best Glide Endurance Speed

Effect of Wind
35. When gliding for maximum endurance, wind will have no effect since it is only the horizontal velocity and hence distance travelled which is affected.

Chapter 7 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
36. The glide or descent range is defined as the horizontal distance travelled during the descent. The best glide distance through the air is achieved by flying the aircraft for minimum drag. In still air conditions this will achieve the greatest ground distance covered during the descent. The true air speed for minimum drag is found from the graph of power required against airspeed by extending a line from the point of origin of the graph to make a tangent with the power required curve. The TAS perpendicularly beneath the point of intersection is the best TAS for range in still air. This is illustrated at Figure 7-14. 37. When descending against a headwind the ground distance covered per unit time is less and so, to increase the ground distance covered the TAS must be increased. Similarly a tailwind would increase the ground distance covered or, alternatively, the descent speed would have to be reduced to achieve a given glide range. The effect is also illustrated at Figure 7-14.

Chapter 7 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-14 Effect of Wind Component on Maximum Range Glide Speed

Chapter 7 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight

Effect of Weight
38. Providing that the aircraft is flown at the speed that gives the best lift/drag ratio, the weight of the aircraft will not affect the glide range. This is shown at Figure 7-15, which illustrates that the increased weight will cause the aircraft to move faster down the descent path, which will increase the lift without a requirement to change the angle of attack. 39. However since a heavier aircraft will glide at a higher speed, it will cover the same distance as a lighter aircraft in a shorter time and its endurance is therefore, less. Consequently, glide endurance reduces with increasing weight.

Chapter 7 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-15 Effect of Weight on Glide Endurance

Chapter 7 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight

Forces in a Steady Co-ordinated Turn


40. During a turn the aircraft weight still acts vertically downward, however the lift vector is now inclined by the bank angle. The amount of lift that was sufficient to maintain straight and level flight is therefore insufficient to maintain level flight with the aircraft turning, as shown at Figure 7-16.

FIGURE 7-16 Lift in a Turn

41. The horizontal component of the total lift vector is used to provide the centripetal force that causes the aircraft to follow a curved path.

Chapter 7 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
42. The additional lift required must be generated by increasing the angle of attack and this will result in increased induced drag and a subsequent decrease in airspeed, unless engine power is increased. 43. The increase in lift required in a turn may be considered as compensating for the apparent increase in aircraft weight during the turn. 44. The amount by which the weight apparently increases is known as the load factor (n) and is given by the formula; Lift n = -----------------Weight 45. From Figure 7-17 it can be determined that the bank angle is a trigonometrical function of the weight and lift vectors such that: Weight Cos = -----------------Lift 46. Therefore, by transposition of formulae: 1 n = -----------Cos 47. A turn employing an angle of bank of 60 will therefore produce a load factor of:

Chapter 7 Page 27

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
1 1 n = ------------------ = ------= 2 Cos60 0.5 48. The forces acting on an aircraft in a steady co-ordinated turn are shown in Figure 7-17. The vertical component of the lift supports the weight and the horizontal component is used to provide the centripetal force that causes the aircraft to follow a curved path.

Chapter 7 Page 28

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-17 Forces in a Steady Co-ordinated Turn

Chapter 7 Page 29

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
49. When any body is turning about a fixed point at constant speed (V) there is an acceleration towards the centre of the turning circle of: V2 -----r where r is the radius of the turning circle

50. If the body has a weight (W), then the centripetal force required to maintain the body at a given turning radius r is given as: WV 2 Centripetal Force = -----------gr 51. Referring to Figure 7-17 and resolving the forces vertically, WV 2 L Sin = -----------gr and horizontally L Cos = W V2 therefore Tan = -----gr and so the angle of bank is dependent upon TAS (V) and radius of turn (r) and independent of aircraft weight (W). Now, transposing this formula,

Chapter 7 Page 30

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
V2 r = ---------------gTan Thus the radius of turn can be calculated given aircraft speed and angle of bank. (Assuming a constant value of g or 9.81 m/s) 52. For example, if an aircraft enters a 30 bank co-ordinated turn at a TAS of 180kt, the turn radius will be: (assuming 1kt = 1.51 m/s i.e. 180kt = 92.7 m/s. V2 r = ---------------gTan ( 92.7 ) r = ----------------------------------9.81 Tan 30 8593.29 = ----------------------------9.81 0.577 = 1518m 53. Thus the turn radius for any given combination of speed and bank angle can be found. To reduce the turn radius the bank angle must be increased. However, the vertical component of lift must be maintained which means increasing the angle of attack to increase lift and therefore the load factor increases. Furthermore, to avoid stalling the speed must be increased as bank angle is increased. The minimum turn radius for any aircraft is determined by the engine power available to maintain speed above Vm, the manoeuvring stall speed, and the limiting load factor n.
2

Chapter 7 Page 31

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
54. In order to avoid overstressing the aircraft by "pulling too many gs" (exceeding the limiting load factor) in the turn the pilot needs a single parameter by which to judge the angular velocity of the aircraft. 55. The angular velocity is the rate at which the aircraft is turning and is referred to as the rate of turn or rate of change of heading. It is the ratio of TAS to turn radius and is usually measured in radians per second V i.e. Rate of Turn = --- radians per second R but r = V therefore Rate of Turn = --------------------------2 V gTan gTan and, Rate of Turn = --------------V 56. Thus it is a function of the two main variables, bank angle and TAS and is more conveniently measured in degrees per second. 57. For passenger transport aircraft the maximum rate of turn is usually limited to 3 per second, which is known as a Rate One turn. The pilot is thus able to maintain an indicated rate one turn by adjusting the bank angle to suit the airspeed. As a result the load on the aircraft and the g acceleration felt by the passengers will always be well within acceptable limits. V2 -----------------g Tan

Chapter 7 Page 32

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
58. In a steady co-ordinated (balanced) turn, the only acceleration on the aircraft is along the normal axis. In an unco-ordinated turn, however, there is a lateral acceleration acting towards the centre of the turn, slip, or away from the centre of the turn, skid. 59. Where the acceleration is to the inside of the turn, the aircraft is slipping and has too much bank for the speed and radius. Where the acceleration is to the outside of the turn, the aircraft is skidding and has insufficient bank for the speed and radius. 60. Therefore, an aircraft which is overbanked will slip-into the turn and this can be corrected by increasing the speed or reducing the angle of bank whilst applying more rudder. 61. On the other hand, an underbanked aircraft will skid-out of the turn and this can be corrected by reducing speed or increasing the bank angle whilst applying less rudder.

Asymmetric Thrust
62. Loss of a power plant on a multi-engine aeroplane produces a yawing moment about the normal axis due to the greater thrust on one side of the normal axis and the greater drag on the other. This was discussed in Chapter 6, on Control. The yawing moment due to asymmetric thrust must be countered by the use of rudder deflection to provide an equal and opposite yawing moment. 63. The side force produced by rudder deflection will depend upon the dynamic pressure, which of course depends upon the indicated airspeed. The lowest IAS at which the rudder at full deflection provides just sufficient force to create a yawing moment balancing the moment due to asymmetric thrust is known as the minimum directional control speed. Minimum directional control speed is usually no greater than 1.2 times the stall speed at take-off configuration and minimum all-up weight.

Chapter 7 Page 33

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
64. Figure 7-18 shows the force balance when the rudder is deflected towards the operational engine on a twin engine aircraft with one engine inoperative. Deflection of the rudder gives camber to the vertical fin, creating sideways lift. The distance between the CP of the fin and the aircraft CG is the moment arm of the fin force. The distance between the thrust line of the operating engine and the CG is the asymmetric thrust moment arm.

Chapter 7 Page 34

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-18 Yawing Moments Under Asymmetric Power

Chapter 7 Page 35

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
65. The asymmetric thrust moment arm will be greatest in the event of failure of the most outboard engine on the wing. However, with propeller powered aircraft, where the propeller rotation is in the same direction for each engine (ie. without counter rotating propellers) failure of one engine will be more critical, due to asymmetric blade effect (see Chapter 12). 66. For example, for a right hand tractor propeller, asymmetric blade effect causes the thrust line to be displaced to the right as in Figure 7-19 and the moment arm of the right engine is greater than that of the left.

FIGURE 7-19 Asymmetric Blade Effect

67. Consequently, failure of the left engine will result in a greater out of balance yaw and roll moment from the right engine. The left engine is referred to as the critical engine and is used when establishing minimum control speeds. 68. Due to the side force on the vertical tail surface it is necessary to maintain a slight angle of bank towards the operative engine. This raises the wing on the inoperative side and provides a horizontal component of lift to balance the side force of the tail.

Chapter 7 Page 36

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
69. It is important in asymmetric power situations to avoid over-banking the aircraft. It will be recalled that during a banked turn the load factor, and therefore the stalling speed, increases. Of even greater importance when flying with one engine inoperative, the induced drag increases and therefore speed decreases. Loss of speed is highly undesirable bearing in mind the need to stay above minimum directional control speed. Up to 15 of bank the increase in stall speed and induced drag is of no great significance, but beyond that bank angle both increase dramatically. 70. With the rudder at or close to maximum deflection to maintain directional control, the fin is inevitably close to its stalling angle of attack. Any loss of airspeed or tendency to sideslip under these circumstances is very liable to result in fin stall and complete loss of directional control. Since excessive use of bank angle renders sideslip more probable and is liable to cause speed loss due to greater induced drag it follows that it can lead to fin stall. 71. In addition to the asymmetric control effects with one engine inoperative on a multi engine aeroplane the effect of loss of thrust must be considered. Besides the reduction in level flight speed due to reduced thrust available, there is less excess thrust for acceleration or climbing. Figure 7-20 shows in graphical form the reduction in excess thrust available when only half the engines are operative. Clearly the climb performance of the aeroplane is greatly diminished under these circumstances. In conditions of high all-up weight, high altitude or high ambient temperature the reduction in excess power may mean that the aircraft not only has insufficient power to climb, but possibly insufficient to maintain height.

Chapter 7 Page 37

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
FIGURE 7-20 Reduction in Excess Thrust Available Under Asymmetric Power

Chapter 7 Page 38

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
72. When conditions are critical, as in the above paragraph, any increase in total drag whether due to excessive bank angle during a co-ordinated turn or to the extension of flaps or landing gear may create a thrust deficiency and make it impossible to maintain height. Figure 7-21 illustrates in graphical form the differences in thrust required in turning and level flight and clean versus landing configurations, against thrust available with one engine inoperative.

FIGURE 7-21 Thrust Defficiency Under Asymmetric Power and Different Configurations

Chapter 7 Page 39

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
73. Under asymmetrical power conditions the use of ailerons must be exercised with care since the secondary effect of roll is yaw. Yaw towards the operative engine may lead to sideslip and/or fin stall, yaw towards the dead engine may lead to loss of directional control. 74. Besides the yawing moments due to asymmetric thrust there are also roll moments, especially on propeller powered aircraft. With a twin engine propeller aircraft the torque reaction to the operative propeller is not balanced by the feathered propeller on the inoperative engine and this creates a rolling moment about the longitudinal axis. The direction of roll will depend upon the direction of rotation of the operative propeller. 75. When trailing edge wing flaps are extended the propeller wash adds to the lift force generated by them. Clearly, with no wash from the feathered propeller there is greater lift on the powered side, creating a rolling moment towards the dead engine. 76. Sideslip will also induce a rolling moment because of the unbalance of lift on the wings, the greater the sideslip angle the greater the roll moment resulting. 77. Furthermore, with an increase in weight, more power will be required to overcome the increased drag and consequently there will be an increase in the yawing moment, following engine failure, at this increased weight. 78. This increased yawing moment results in greater sideslip towards the live engine and a subsequent increase in the resulting rolling moment to the dead engine. This effect is more pronounced with dihedral and sweepback. 79. The minimum control speeds when operating in an asymmetric power situation are defined below.

Chapter 7 Page 40

G LONGHURST 1999 All Rights Reserved Worldwide

Forces in Flight
80. VMCA is the minimum control speed in the air in the take-off configuration. It is the minimum speed at which it is possible to suffer a critical engine failure and maintain control of the aeroplane within defined limits. 81. VMCG is the minimum control speed on the ground. It is the minimum speed at which it is possible to suffer a critical engine failure on take-off and maintain control of the aeroplane with the use of primary aerodynamic controls alone, within defined limits. 82. VMCL is the minimum control speed in the air in an approach or landing configuration. It is the minimum speed at which it is possible, with a critical engine inoperative, to maintain control of the aeroplane within defined limits while applying maximum variations of power.

Effect of Altitude on Minimum Control Speeds


83. As altitude is increased, the power available from any power plant, unless augmented, will reduce due to the reduction in density. Consequently, the yawing moment following failure of the most critical engine will be less. The minimum control speeds will, therefore, reduce with increasing altitude.

Chapter 7 Page 41

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Stability
Static Stability Static Longitudinal Stability Static Directional Stability Static Lateral Stability Dynamic Stability Longitudinal Dynamic Stability Lateral Dynamic Stability Asymmetric Effects Effect of Altitude

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Stability

1. The four forces acting on an aircraft in straight and level flight, we know, are lift, weight, and drag. When all the forces acting on it are in balance ie, the sum of the forces and moments is zero, the aircraft is in a state of equilibrium. 2. Newton's First Law states that a body remains in a state of rest or uniform motion unless acted upon by an external force. Thus equilibrium can be a state of rest or uniform motion, that is constant speed and direction. Following a disturbance, stability is concerned with the motion of the body after removal of the external force. This motion can be considered in two ways. Static stability describes the immediate reaction of the body, whilst subsequent reaction to the disturbance is dependent upon dynamic stability.

Static Stability
3. Static stability is the initial tendency of the aircraft to return to the equilibrium position following a disturbance. The response to the disturbance is termed positive, neutral or negative. If the body returns to the position it held prior to the disturbance it is said to be positively stable. If the body takes up a new position of constant relationship to the original position it is said to be neutrally stable. Negative stability indicates a continuous divergence from the original state. The analogy of a ball and bowl is used at Figure 8-1 to demonstrate the above concept.

Chapter 8 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-1 Static Stability

4. An aircraft is subject to disturbance about the three axes of movement, roll, pitch and yaw. Its stability, that is its tendency following a disturbance to return to its original position or to continue to diverge from that position is therefore considered in relation to those axes. Its stability in roll, about the longitudinal axis, is known as lateral stability. Its stability in pitch, about the lateral axis, is known as longitudinal stability. Its stability in yaw, about the normal axis, is known as directional stability.

Static Longitudinal Stability


5. When an aircraft suffers a disturbance about its lateral axis, it will, in the short-term, experience a change in angle of attack, and depending upon the relative positions of the CP and CG, the resulting pitching moment will either tend to restore equilibrium or be destabilising. Consequently, the necessary restoring moment to achieve balance is primarily provided by the horizontal stabiliser.

Chapter 8 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
6. The horizontal stabiliser is usually placed at the tail of the aircraft (tailplane) however, it can be placed ahead of the wings in what is referred to as a canard configuration. The main advantage of the candard design is that the foreplane produces positive lift and for a give speed, a lower angle of attack is needed on the mainplane resulting in less drag and a reduced stalling speed. 7. The greater the distance the stabilising surface is from the centre of gravity of the aircraft, the greater its moment arm and therefore the quicker the aircraft will be returned to its original attitude following a disturbance. This moment arm is often called the tail arm and its length is a measure of the aircrafts longitudinal stability. 8. Figure 8-3 shows the way in which the tailplane helps to produce a state of static longitudinal stability following a pitching disturbance. The success of the tailplane in achieving this relies on the difference between the angle of incidence of the wing and of the tail. This is known as longitudinal dihedral. See Figure 8-2.

FIGURE 8-2 Longitudinal Dihedral

Chapter 8 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
9. When the wing angle of incidence is greater than that of the tailplane the aircraft is said to have positive longitudinal dihedral, an essential requirement for static longitudinal stability. In conventional aircraft this is usually achieved with a slightly negative angle of incidence on the tailplane. From Figure 8-3, if the aircraft is disturbed in pitch, the tailplane will always tend to restore it to level flight and thus the longitudinal dihedral provides a stabilising influence. The same effect is achieved if the wing camber is greater than the tailplane camber. Often the tailplane is not cambered (a symmetrical aerofoil section) or it may even have reverse camber.

FIGURE 8-3 Tailplane Contribution

Chapter 8 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
10. The stabilising effect of a tailplane will be altered if there is a significant change in the downwash over it with changing angle of attack. The downwash angle determines the effective angle of attack of the tailplane, which in turn determines its balancing effect against the wing lift. If, for example, an increased incidence (pitch up) produces an increased downwash angle, the effective angle of attack of the tailplane is reduced and its restoring moment is reduced in consequence. Hence, the longitudinal stability of the aircraft has decreased. In cases where the design of the aircraft produces this effect, the CG would be moved forward to regain longitudinal stability. 11. It should be appreciated that, although the tailplane produces relatively little lift when compared with the wing, the distance from the centre of pressure of the tailplane to the centre of gravity of the aircraft (the arm) is considerable. Consequently the moment produced by the tailplane is significant. 12. The effect that turbulence will have in inducing a pitching moment of the aircraft will depend largely on the relative positions of the centre of gravity and the centre of pressure of the wing. At Figure 8-4 the centre of gravity is shown to be well forward of the centre of pressure. A disturbance that causes a nose up pitching moment will cause an increase in the angle of attack and an increase in the amount of lift produced. This will effectively increase the size of the lift/weight pitching moment (despite a small movement forward of the centre of pressure, which reduces the length of the arm), and this moment will act to push the nose downwards. In other words a state of static stability will tend to exist.

Chapter 8 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-4 Longitudinal Static Stability

13. At Figure 8-5 the centre of gravity is aft of the centre of pressure and now a disturbance which causes the nose to rise will cause an increase in lift which will cause a further (destabilising) movement of the nose in an upwards direction. Thus it can be seen that the aircraft will be both statically and dynamically unstable. This is obviously an undesirable situation, and provides a good reason for always operating any aircraft within the limits of the centre of gravity envelope specified by the manufacturer.

Chapter 8 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-5 Longitudinal Static Instability

14. To summarise, the factors that affect the static longitudinal stability of an aircraft are the design of the horizontal stabiliser and the position of the aircraft CG. The horizontal stabiliser in terms of area, planform, distance of its CP from the aircraft CG and the effect of the wing downwash upon it. The CG position in terms of movement; aft movement decreases positive stability, forward movement increases it. 15. However, it must be appreciated that there is a limit to the forward position of the centre of gravity. The greater the natural stability of an aeroplane due to a forward centre of gravity the more the elevators must move in order to pitch the nose up during take-off or landing flare. If the centre of gravity were too far forward there would be insufficient elevator movement available to achieve these essential pitch control requirements.

Chapter 8 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
16. If the centre of gravity is moved progressively aft there will come a point at which the destabilising effect of the aft centre of gravity will exactly offset the efforts of the tailplane to produce a state of positive stability. The position of the centre of gravity at this time is known as the neutral point, and, with the centre of gravity at this point, the aircraft is in a state of neutral stability. For obvious reasons the aft limit of the centre of gravity envelope of any aircraft is always forward of the neutral point. 17. The distance through which the centre of gravity can be moved aft from a given datum to the neutral point is known as the centre of gravity margin, and is a measure of the longitudinal stability of the aircraft. The greater the margin the greater the longitudinal stability of the aircraft. It is also known as the static margin and will be at a minimum when the CG and neutral point are co-incident. 18. If the CG is aft of the centre of pressure the aircraft will be longitudinally unstable. Conversely, if the CG is forward of the centre of pressure the aircraft will be stable. If the two are coincident a neutral condition of stability exists. 19. When an aircraft in straight and level flight is subjected to a disturbance that alters its angle of attack (incidence), there must inevitably be a change in the pitching moment about the CG. For static stability it is necessary that, if the initial change is nose-up (positive) the resultant pitching moment must be nose-down (negative), and vice versa. In other words, the rate of change of pitching moment with changing angle of attack must always be negative. This is illustrated in the graph at Figure 8-6 of pitching moment coefficient Cm against angle of attack .

Chapter 8 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-6 Graph

20. The negative slope indicates positive static longitudinal stability and the greater the slope, the greater the longitudinal stability. 21. The trim point shown on the graph at Figure 8-6 occurs at the angle of attack at which the overall moment about the CG is zero. It is the point where the wing moment is equal to the tail moment and the aircraft is said to be in trim. The actual angle of attack at which this occurs is dependent upon the amount of longitudinal dihedral of the aircraft.

Chapter 8 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Effect of CG Position
22. As the CG is moved aft, the aircraft static stability reduces, becoming neutral then unstable with continued rearward movement. This is shown at Figure 8-7 where the CG position is defined with reference to mean aerodynamic chord.

Chapter 8 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-7

23. With the CG at the neutral point, neutral static stability exists, the slope of the curve is zero and there is no change in pitching moment. The neutral point defines the most aft CG position without static instability.

Chapter 8 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
24. In cases where the CG of the aircraft is significantly moved from its normal position it may be restored by placing ballast weights in a balancing location. Suppose, for example, an aircraft is to be ferried empty of passengers or freight over a long distance. It may require a full fuel load that puts the CG too far aft and this would be balanced by placing ballast in a forward cargo hold. 25. In cases where the vertical force balance changes in flight the aircraft may be re-trimmed for stable horizontal flight by weight trimming, or the movement of weight within the aircraft. An example of this can be found in the chapter on High Speed Flight, where fuel is transferred aft in Concord to balance the aft movement of the centre of pressure during supersonic flight.

Effect of Elevator Deflection


26. The control surface connected with longitudinal stability is the elevator or, in some aircraft, all-moving stabiliser. If it is required to increase the angle of attack from the trim point the natural tendency of a longitudinally stable aircraft is to produce a stable nose-down pitching moment. In order to maintain the increased angle of attack a moment opposing the stabilising nose-down moment must be provided by the elevators. This is achieved by upward deflection of the elevators, which alters the camber of the tailplane providing an increased negative (downward) force and establishing a new trim point at a greater wing angle of attack. When a reduced angle of attack is required, the reverse procedure is adopted. Deflection of the elevators to achieve a changed angle of attack does not usually affect the longitudinal stability of the aircraft. 27. Figure 8-8 illustrates graphically the effect of elevator deflection. From the curves for various degrees of deflection it can be seen that the stable negative slope of the CM CL curves remains constant, only the position of the trim point varies.

Chapter 8 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-8 Effect of Elevator Deflection on Graph

Stick-Fixed and Stick-Free Stability


28. At each position of elevator deflection the static longitudinal stability of the aircraft is the same, but the lift coefficient at the trim point is different. Moving the elevator or stabiliser has no effect upon the contribution of the tailplane to stability, but it does change the pitching moment, and therefore alters the lift coefficient at which equilibrium (trim) will occur. Because these changes in lift coefficient for trim occur at fixed elevator positions, this is known as stick-fixed stability.

Chapter 8 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
29. If the elevators are allowed to float, as in hands-off (or stick-free) flight, they will tend to align with the free stream flow when the angle of attack of the tailplane is changed. If, for example, the tailplane angle of attack is increased the elevators will tend to float up, reducing the restoring moment of the horizontal tail surfaces. Consequently, the stick-free stability of the aircraft is usually less than the stick-fixed stability. 30. Another way of expressing this is to say that the position of the neutral point will be different for stick-fixed and stick-free conditions, which will in turn alter the CG margin. The CG margin will be greater under stick-fixed conditions.

Effect of Trim
31. With a stable aircraft, an increase in CL produces a nose down pitching moment change. In order to balance this in a new trim condition, the elevator must be deflected upwards to provide the necessary downward increase in tail lift. The resulting elevator hinge moment is nose-up and to balance this and trim out the stick force, the trim tab must be deflected downwards, ie. in the positive sense. 32. Therefore, a positive change in CL is accompanied by a positive change in trim tab angular deflection.

Chapter 8 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Wing/Tail Relationship
33. The ratio of the tail volume to the wing volume is the main parameter used by the aircraft designer in establishing the longitudinal stability of an aircraft. The tail volume is the product of the tail moment arm and the tailplane area, the wing volume is similarly the product of wing moment arm and wing area. In both cases the moment arm is the distance from the CP to the aircraft CG. The greater the tail volume in comparison with wing volume, the greater the longitudinal stability. 34. So what of tailless swept wing and delta planform aircraft? With swept wing aircraft the wing tips can be twisted to reduce their incidence (washout) and they therefore produce the same effect as a tailplane. With delta wings the reduced incidence is made towards the trailing edge.

Fuselage Contribution
35. The contribution of the fuselage, and to some extent the engine nacelles, is usually destabilising. If the fuselage aerodynamic centre is ahead of that of the wing, then the resulting wingfuselage combination aerodynamic centre will be further forward. Effectively, this means that the aircraft CG is relatively further aft and hence its longitudinal stability is reduced. 36. Furthermore, the induced upwash ahead of the wing increases the de-stabilising influence from the fuselage and nacelles ahead of the wing. The downwash behind the wing on the other hand, reduces the de-stabilising influence from the areas of the fuselage and nacelles aft of the wing.

Chapter 8 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Downwash Changes
37. With an increase in angle of attack, the downwash from the wings is increased which reduces the effective angle of the tailplane thus reducing its lifting force. Consequently, the restoring movement provided by the tailplane is reduced and the neutral point would therefore be further forward and the longitudinal static stability correspondingly reduced. 38. Any factor which alters the downwash at the tail will affect the tail contribution, notably trailing-edge flap extension and increased wing camber. However, extending trailing-edge flaps also affects the wing pressure distribution causing a rearward movement of the centre of pressure and a resulting nose-down pitching moment change, ie. a stabilising effect which now opposes the destabilising effect due to increased downwash. Usually, use of high lift devices is de-stabilising.

Configuration Changes
39. Extension of the landing gear lowers the position of the drag line and additionally, if the thrust line does not pass through the centre of gravity, variations in thrust with angle of attack changes, result in pitching moment changes. If the thrust line is high relative to the CG, it is stabilising, and if low, de-stabilising.

Chapter 8 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Elevator Position and Speed


40. The value of the lift coefficient depends upon, among other things, the dynamic pressure (IAS). Trim airspeed can therefore be correlated with elevator deflection, as shown in the graph at Figure 8-9. It is highly desirable that aft movement of the control column is needed to increase the angle of attack and trim at low airspeed and forward movement to decrease the angle of attack and trim at higher airspeed. This is known as stick-position stability and is the case when the CG is located forward of the stick-fixed neutral point. 41. It follows that if CL is increased, for example by the use of high lift devices, a reduced elevator deflection is required for a given trim airspeed. The effect on the elevator position speed graph is therefore a reduction in the trim speed for zero elevator deflection.

FIGURE 8-9 Elevator Position Speed Graph

Chapter 8 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Stick Force and Speed


42. Aircraft that use a tailplane and elevator system for longitudinal stability and control usually incorporate trim tabs in conjunction with the elevators. The functions of the two are quite different. The elevator is there to adjust the lift coefficient of the tail surface and thus the pitching moment coefficient CM. The trim tab is there to adjust the elevator hinge moment and therefore the stick force necessary to deflect the elevator. 43. The basic stick-free stability of the aeroplane will require a stick force to maintain a given attitude that is independent of airspeed, but the force to be balanced by the trim tab is dependent upon dynamic pressure, or IAS. Therefore, the trim tab setting necessary to maintain the correct attitude for level flight with zero control column force will vary with airspeed. The faster the speed the less the angle of attack necessary to maintain level flight, so the less the up elevator deflection required. The less the elevator deflection required, the less the opposite trim tab deflection needed to give zero stick force. This is illustrated in the graph at Figure 8-10. The same situation will exist for an all-moving stabiliser, although the need for force balancing is less, since it is usually moved by powered controls.

Chapter 8 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-10 Effect of Trim Tab on Stick Force Speed Graph

44. For certification, there are criteria of stick force versus speed characteristics laid down by the JAA and they are as follows: JAR 25.173(c) 45. The average gradient of the stable slope of the stick force versus speed curve may not be less than 1 pound for each 6 knots. (The average gradient is taken over each half of the speed range between 0.85 and 1.15 V trim ). JAR 23.173(c)

Chapter 8 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
46. The stick force must vary with speed so that any substantial speed change results in a stick force clearly perceptible to the pilot. 47. If the position of the CG changes whilst trimmed for flight at constant airspeed this must affect the stick force required to maintain the trim speed. An aft-moving CG will have the effect of reducing the stick force necessary to alter the pitch attitude of the aeroplane. This is illustrated in the graph at Figure 8-11. The further aft the CG the shallower the gradient of the stick force EAS curve and the less the stick position stability, until the gradient becomes zero at the neutral stability point. This is when the CG is at the stick-free neutral point, beyond this a condition of stick position instability will exist. In other words, the stick would have to be pushed at lower speeds and pulled at higher speeds to maintain attitude.

FIGURE 8-11 Effect of CG Position on Stick Force - Speed Graph

Chapter 8 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
48. Friction in the control system is highly undesirable and can seriously upset the feel of the controls since it makes it very difficult to determine when the correct trim setting has been applied. This is because the friction produces a range of force to be overcome, which translates into a range of trim speeds. The effect is illustrated in the graph at Figure 8-12.

FIGURE 8-12 Effect of Control System Friction on Stick Force Speed Graph

49. At Mach numbers in excess of 1.0 there is a marked aft movement of the CP and automatic Mach trim is necessary to cope with the ensuing Tuck Under. This is explained fully in Chapter 9, High Speed Flight.

Chapter 8 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Stick Force Modification


50. Devices have been developed to modify the forces acting on the pilots control column in order to achieve stick force stability, notably the downspring and the bob weight, both of which add to the pull force necessary on the control column independent of control deflection or airspeed. 51. The two devices are illustrated at Figure 8-13 and their effect on the stick force airspeed graph, which is the same in both cases, is illustrated at Figure 8-14.

Chapter 8 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-13 Downspring and Bobweight

Chapter 8 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-14 Effect of Downspring and Bobweight on Stick Force Speed Graph

Manoeuvring Stability
52. When an aeroplane is subjected to a pull-up it rotates about its lateral axis. In other words the nose is rotating up and the tail is rotating down. The effect upon the tailplane is to change the effective angle of attack so that it is opposing the rotation and providing a restoring moment. Thus, the stabilising effect of the tailplane is greater during manoeuvring than in steady level flight.

Chapter 8 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
53. The more longitudinally stable an aircraft is, the greater the force necessary to displace it in pitch. Pitch up is achieved by moving the elevator up, which produces a downward force at the tail to give the necessary pitching moment. The greater the rate of pitch change the greater the manoeuvring stability of the aircraft and so the greater the elevator/stick force required. Also, the greater the rate of pitch change the greater the load factor, or g acceleration. Thus it can be seen that, in a longitudinally stable aircraft, the stick force increase is in direct relation to the g acceleration, or load factor, n. This is illustrated in the graph at Figure 8-15, which shows a plot of stick force per g. For a large transport aircraft the stick force per g is deliberately high in order to prevent inadvertent overstressing of the aircraft through excessive control inputs. Such aircraft have a lower limit load factor than designs intended for high manoeuvrability and are certified accordingly.

Chapter 8 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-15 Stick Force per g Graph

54. The further aft the CG the lower the stick force per g, which is to be expected since we know that an aft CG decreases longitudinal stability. 55. Manoeuvring stick force can be reduced with adjustment of the trim tab setting, but this is not usually a recommended procedure since increasing stick force per g is a desirable feature. The bob weight, mentioned above, has the effect of increasing stick force per g. This is a useful feature if the basic aeroplane has too low a manoeuvring stick force stability, possibly because of an inadequate CG margin, and might be added to meet certification requirements. The downspring has no effect upon manoeuvring stick force stability.

Chapter 8 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Special Circumstances
56. Ice on an aircraft can greatly increase the surface friction drag and decrease the maximum lift coefficient of the wing and tail surfaces. The alteration to the drag line, especially where these surfaces are well displaced from the longitudinal centre line, may result in a significant change in pitching moment. The weight of the ice may also alter the position of the CG, particularly likely on swept wing aircraft. In either case the longitudinal stability of the aeroplane will be affected. 57. Similarly, reduction in the lift coefficient of either the wings or the tailplane may alter the position of the overall CP and thus the CG margin. It must be remembered that an increased CG margin may be no more desirable, with its decreased control effectiveness, than a decreased one with its decreased stability. Extending the flaps under these conditions could render a delicate situation worse and introduce control difficulties at a critical stage of flight. 58. Any other factor that increases local drag or decreases local lift may have effects similar to those described in the preceding paragraph. Rain can produce limited drag changes, especially at low speed, as can deformation of the airframe, where the effects are more noticeable at high speed.

Static Directional Stability


59. When an aircraft yaws it does so about the normal, or vertical, axis. When this yaw is caused by some disturbance, the ability of the aircraft to return to its original heading (without control surface movement) is a measure of the aircraft's static directional stability. It is often referred to as the weathercocking stability of the aircraft since it is the ability of the aircraft to align itself with the relative airflow, just as a weathercock aligns itself with the wind direction.

Chapter 8 Page 27

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
60. This is achieved by mounting a vertical stabilising aerofoil at the rear of the fuselage. If the aircraft is disturbed about the normal axis the vertical stabiliser, or fin, has an angle of attack to the relative airflow and the sideways lift thus produced has a restoring effect as shown at Figure 8-16. The fin is a symmetrical aerofoil that produces lift at positive angles of attack (up to the stall).

FIGURE 8-16

61. The larger the surface area of the fin the greater its restoring force, this is illustrated at Figure 8-17.

Chapter 8 Page 28

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-17

Chapter 8 Page 29

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
62. Alternatively, the further the fin is set back from the centre of gravity of the aircraft the greater its restoring moment. Hence, static directional stability may be achieved with a large fin and a short tail moment or a small fin and a long tail moment, the choice is usually determined by other design considerations.

Yaw Moment Coefficient


63. The yawing moment about the normal axis that tends to rotate the aircraft in the horizontal, or yawing plane is conventionally referred to as being positive when yaw is to the right and negative when it is to the left. The yawing moment, N, is defined as: 1 2 N = C n -- V Sb where b = wing span 2 and Cn = yawing moment co-efficient By transposition: N C n = -------------------1 2 -- V Sb 2

Chapter 8 Page 30

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Yaw Angle
64. The yaw angle is defined as the angle by which the aeroplane centreline (the longitudinal axis) is displaced in azimuth from a given reference. The angle is regarded as positive when the centreline is to the right of the reference azimuth direction. Yaw angle is only of real significance in wind tunnel experiments.

Sideslip Angle
65. The sideslip angle is the angle by which the aeroplane longitudinal axis is displaced from the relative airflow. The angle is regarded as positive when relative airflow is to the right of the aircraft centreline. Yaw and sideslip angles are illustrated at Figure 8-18.

Chapter 8 Page 31

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-18 Yaw and Sideslip Angles

Chapter 8 Page 32

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Graphical Explanation of Static Directional Stability


66. An aircraft exhibits static directional stability when, having been subjected to a sideslip angle, the weathercocking effect turns the nose toward the direction of sideslip. In other words, if the aircraft is subjected to a positive sideslip angle (+), a positive yawing moment coefficient (+Cn) results. Similarly, if the aircraft is subjected to a negative sideslip angle (-), a negative yawing moment coefficient (-Cn) results. 67. When this is plotted graphically as a curve of Cn versus , the resulting curve will have a positive slope, the greater the static directional stability of the aircraft, the steeper the slope. If the slope were to be zero this would indicate a situation of neutral static directional stability. In this case, if the aircraft were to be subjected to a sideslip angle there would be no tendency to return to the original heading (no weathercocking effect). An aircraft exhibiting static directional instability, upon being subjected to a sideslip angle, would continue to diverge from the original heading and this would appear as a negative slope on the Cn - graph. 68. A graph of Cn versus is shown at Figure 8-19. It will be noted that at high sideslip angles the curve becomes horizontal, indicating neutral static directional stability, and at greater angles becomes negative, indicating instability. This is due to the aerodynamic stalling of the fin and should not occur within the normal manoeuvring limits of the aircraft.

Chapter 8 Page 33

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-19 Cn Graph

Centre of Gravity
69. The position of the aircraft CG will influence the effectiveness of the fin in achieving static directional stability. The further forward the CG the longer the moment arm and the greater the restoring moment. Once again it is evident that it is highly undesirable to operate an aircraft with the CG aft of permissible limits.

Wing Effect
70. The two wing design features which contribute to the directional stability of the aircraft are dihedral and sweepback, although the former is relatively insignificant particularly when compared to the latter.

Chapter 8 Page 34

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
71. The inclination of the wings towards the tips associated with dihedral, effectively tilts the lift vector thus producing a destabilising contribution to the yawing movement due to yaw. On the other hand, with sweepback, it is the component of the relative airflow normal to the leading edge (or to the line of aerodynamic centres in the case of a wing whose leading and trailing edges are not parallel) which is responsible for the drag of the wing as shown in Figure 8-20. The drag therefore of the leading (starboard) wing is greater than that of the trailing wing due to its reduced effective sweep angle and is thus a stabilising influence. This effect is of course significant when the sweepback angle is large.

Chapter 8 Page 35

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-20 Effect of Sweepback on Static Directional Stability

Chapter 8 Page 36

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Fuselage Effect
72. The fuselage has a generally destabilising effect upon aircraft static directional stability. This is because, in subsonic flight, the CG is usually well aft of the centre of pressure of the fuselage. Hence, in the event of sideslip the relative airflow exerts a greater yawing moment forward of the CG than aft of it. This is the reason why an aircraft needs a vertical stabiliser (fin), to overcome the destabilising influence of the fuselage. The effect is often noticeable at high angles of attack, when the fin is stalled by the disturbed airflow from the fuselage and the aircraft becomes directionally unstable. 73. To counter this, strakes or ventral fins are sometimes fitted. These are flat plates or strips attached to the underside of the rear fuselage and running parallel to the aircraft centreline, forming a keel surface aft of the CG. They increase the directional stability in normal flight and help maintain it even if the fin begins to stall. Such devices are generally found on short-bodied training and military aircraft, rather than large transports. Another device is the use of multiple vertical tails, either in the form of twin fins as seen on many modern fighter aircraft or small tailplane mounted additional fins, as seen on some larger aircraft.

Dorsal Fin
74. The dorsal fin is an extension of the fin, running forwards along the top of the fuselage, as illustrated at Figure 8-21.

Chapter 8 Page 37

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-21 Dorsal Fin

75. The function of the dorsal fin is twofold. It significantly increases the side surface area aft of the CG and thereby increases the static directional stability of the aircraft by increasing the yawing moment and thus the weathercocking effect. Secondly, by reducing the effective aspect ratio of the vertical fin it increases its stalling angle of attack. Thus the fin remains effective at greater sideslip angles.

Fin Sweep
76. An alternative way of reducing the fin aspect ratio to increase its stalling angle is to sweep it backwards. This also has the benefit of moving its centre of pressure slightly rearwards, increasing the yawing moment of the tail.

Chapter 8 Page 38

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Effect of Major Components


77. Figure 8-22 illustrates in graphical form the effects of the major parts of the aircraft on directional stability. It will be seen that the fuselage on its own has a destabilising effect, evidenced by the negative slope of the Cn - curve for fuselage alone. The vertical tail has a strong stabilising influence and so the combined effect of tail and fuselage produces a stabilising, positive, curve. The addition of a dorsal fin increases the directional stability, as shown by the increased gradient of the Cn - curve.

Chapter 8 Page 39

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-22 Effect of Aircraft Components on Static Directional Stability

Propeller Slipstream
78. The propeller of a single-engine aircraft produces a swirling slipstream that is apt to strike one side of the vertical tail more strongly than the other. This creates a directional control requirement from the rudder, which must balance the asymmetric airflow and prevent sideslip. The result is usually that the aircraft is more prone to sideslip in one direction.

Chapter 8 Page 40

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Static Lateral Stability


79. When an aircraft is subject to a disturbance that imparts a rate of roll, one wing rolls downwards and the other upwards. The downgoing wing consequently has a greater effective angle of attack than the upgoing wing, so it creates greater lift. Providing that the angle of attack of the downgoing wing does not exceed the stalling angle, the differential lift thus produced will tend to resist the rolling movement. This is known as roll damping and is shown at Figure 8-23. It should be noted that it does nothing to restore the aircraft to a wings level condition. In other words a state of neutral static stability will tend to exist.

Chapter 8 Page 41

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-23 Roll Damping

80. However, in addition to the rolling motion produced by a disturbance about the longitudinal axis, the aircraft also experiences a sideslipping motion due to the inclination of the lift vector as at Figure 8-24.

Chapter 8 Page 42

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-24 Sideslipping Motion

81. The static lateral stability of an aeroplane is a measure of its tendency to return to wings-level flight when it has been disturbed in roll. The tendency to recover from such a disturbance is always reliant upon the effects of sideslip upon the aeroplane.

Chapter 8 Page 43

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
82. As well as the rolling moment due to sideslip, the sideslip will also produce a yawing moment, the strength of which will depend upon the static directional stability. However, when considering the static lateral stability we need only consider the relationship between sideslip and rolling moments.

Roll Moment Coefficient (CL)


83. As with most aerodynamic considerations it is convenient to consider rolling moments in coefficient form, independent of such things as weight, speed, altitude and so on. The rolling moment, L, is defined in coefficient form by the equation: 1 2 L = C1 -- V Sb 2 Therefore by transposition: L C 1 = -------------------1 2 -- V Sb 2

Sideslip Angle and Rolling Moment Coefficient


84. When an aircraft is subject to a positive sideslip angle , that is with the relative airflow to the right of the aircraft centreline, it exhibits lateral stability if a rolling moment to the left results. The sideslip will have caused the right wing to drop and the rolling moment produces a correcting roll to the left.

Chapter 8 Page 44

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
85. When plotted as a curve of rolling moment coefficient Cl against sideslip angle the slope of the curve should be negative. If the slope is zero this represents neutral lateral stability and if it is positive the aircraft is exhibiting lateral instability, as shown at Figure 8-25.

FIGURE 8-25 Rolling Moment Coefficient Sideslip Angle

86.

The aircraft design features which affect static lateral stability are:

Chapter 8 Page 45

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Dihedral
87. Dihedral has been defined in an earlier chapter as the upward inclination of the wing to the plane through the lateral axis. The purpose of this wing form is solely to make improvements to the lateral stability of the aircraft. Figure 8-26 shows the effect of dihedral on lateral stability.

FIGURE 8-26 Dihedral Effect

88. Because of the wing dihedral and the sideslip, the effective angle of attack of the lower wing is greater than that of the upper. Thus the lift produced by the lower wing is now considerably greater than that produced by the upper, producing a rolling moment to roll the wings back towards the level position. Hence, dihedral improves the lateral stability of an aircraft and is in fact one of the most important contributions to the overall stability. It is for this reason that lateral static stability is often termed the dihedral effect. Other design features which also contribute to static lateral stability therefore can be considered to increase the effective dihedral of the wing.

Chapter 8 Page 46

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Anhedral
89. From the foregoing it should be apparent that anhedral, a downward inclination of the wing to the plane through the lateral axis, will have the opposite effect to dihedral. Where laterally stabilising features such as large fin surfaces and swept wings are essential for other purposes, such as high-speed flight, it may be necessary to deliberately introduce laterally destabilising features. Anhedral is often used for this purpose and many examples are currently in service.

Swept Wings
90. Wings which are swept backwards from root to tip also have a stabilising effect, as illustrated at Figure 8-27. The swept wing increases lateral stability because, when an aircraft rolls to one side, it also tends to sideslip in that direction. In a swept wing aircraft this has the effect of decreasing the effective chord length, and increasing the effective span, of the downgoing wing (i.e. increasing its aspect ratio compared to the upgoing wing). The resultant extra lift generated by the high aspect ratio wing creates a rolling moment to roll the aircraft back to wings level. It is generally reckoned that 10 of sweepback will provide the same stabilising effect as 1 of dihedral.

Chapter 8 Page 47

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-27 Effect of Sweepback on Lateral Stability

Wing Location
91. High-winged aircraft, that is aircraft with the wings mounted above the centre of gravity and longitudinal axis, are more laterally stable than mid-wing or low wing aircraft. 92. It is often incorrectly thought that mounting the wings above the centre of gravity improves lateral stability due to a pendulum effect. In fact it is the effect of sideslip that provides the restoring force. The lift of the lower wing will be greater than that of the upper, since the sideslip flow component induces a circulation tendency around the fuselage, increasing the angle of attack of the lower wing.

Chapter 8 Page 48

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
93. Often high wings are chosen by the designer so that ground clearance can be maintained below the engine pods or the propeller tips. If the lateral stability due to the high wings is greater than required anhedral may once again be used to reduce lateral stability. No doubt the student can think of several examples.

Fin Size
94. A large fin area will increase lateral stability, since the relative airflow in the sideslip provides a righting moment. The higher the centre of pressure of the fin is above the centre of gravity the greater the rolling moment and therefore the greater the lateral stability of the aircraft. This is illustrated at Figure 8-28.

FIGURE 8-28 Effect of High Fin on Lateral Stability

Chapter 8 Page 49

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Ventral Fin
95. If there are side surfaces below the centre of gravity these will produce a moment, during sideslip, which increases the rate of roll. Such keel surfaces increase lateral instability. Some aircraft have, for example, a tall fin to support a tee tailplane clear of turbulent downwash from the wing. Such an aircraft may be laterally stable to such a degree that roll response is poor, so a keel surface (or ventral fin) is introduced deliberately to improve roll response.

Flaps and Lateral Stability


96. The effect of lowering flaps, which are invariably mounted close to the inboard end of the wings, is to move the centre of pressure inwards towards the wing roots. Since the lift/weight arms are now laterally shorter this necessarily has a laterally destabilising effect which can, however, be offset by the use of drooped ailerons, which tend to shift the centre of pressure spanwise towards the wing tips. Drooped ailerons can also be used to increase C Lmax and this, along with the improved spanwise lift distribution, reduces take-off speeds and improves short field take-off performance. When used as such, they are often referred to as flaperons.

Asymmetric Propeller Slipstream


97. The propeller slipstream is normally in balanced flight, symmetrical along the longitudinal axis but, when sideslip is present, this is no longer the situation. The increased dynamic pressure in the slipstream when compared to the free stream now affects the up-going or trailing wing due to sideslip following a disturbance in roll. This increase in the rolling moment, due to the asymmetric slipstream is laterally destabilising.

Chapter 8 Page 50

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Cross-Wind Landings
98. It is desirable that an aircraft should exhibit a degree of lateral stability. However, if roll due to sideslip is too great this can create handling difficulties during crosswind landing and take-off and may render roll control difficult as well. 99. On an approach to land with the aircraft crabbing into a cross-wind, the aircraft has to be yawed to align it with the runway prior to touchdown. Sideslip will thus be increased in the opposite direction to the yaw and, with high lateral stabilising design features, the aircraft will roll away from the sideslip, i.e. sideslip right, roll left. 100. Generally speaking it is desirable from a handling point of view that the aeroplane should have weak positive lateral stability.

Dynamic Stability
101. Dynamic stability considers the change in the state of stability of a body with the passage of time. If, following a disturbance, the deviation of a body from its original position diminishes, the body is said to be dynamically stable. If the induced oscillation remains constant a state of neutral dynamic stability exists. If the induced oscillations increase in magnitude the body is said to be dynamically unstable. These conditions are illustrated at Figure 8-29.

Chapter 8 Page 51

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-29 Dynamic Stability

Chapter 8 Page 52

G LONGHURST 1999 All Rights Reserved Worldwide

Stability

Longitudinal Dynamic Stability


102. When an aircraft is disturbed in pitch from its trimmed level flight condition it usually begins to oscillate about the intended level, the oscillations being in terms of height, speed and load factor. If the aircraft has positive dynamic stability these oscillations will die away due to the damping effect of the horizontal stabiliser. Oscillations in pitch take two forms, the characteristics of which are significantly different. They are the phugoid and the short-period oscillation, both of which are illustrated at Figure 8-30.

FIGURE 8-30 Longitudinal Dynamic Stability

Chapter 8 Page 53

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
103. The phugoid, a rather pompous term invented by the eminent nineteenth century scientist Lanchester, is a long period oscillation in which both aircraft speed and height vary considerably with its change in angle of attack. For example, following a disturbance causing an increase in forward speed, there will be an increase in drag, resulting in a reduction in speed. In addition there is an increase in lift which results in an increase in height and this increase in height further reduces speed until it is below the pre-disturbed value. Now the aircraft begins to descend once more and an oscillatory motion is observed where height gain and speed loss are followed by height loss and speed gain. The extent to which the phugoid oscillation is damped by the horizontal stabiliser depends largely upon the drag characteristics of the aircraft. Modern low drag jet transporters are particularly susceptible to phugoidal oscillation but since it is a long period oscillation, is easily corrected. 104. The short-period oscillation comprises rapid oscillations in pitch, involving large changes in load factor but only small changes in speed and height. Such an oscillation may result from a vertical gust causing an increase in angle of attack and lift, resulting in a change in pitching moment. Since the frequency of this level of oscillation is usually high, it is termed a short-period oscillation. However, if this oscillation is dynamically unstable, any correction necessary to eliminate it will be continually changing in sense, in phase with the oscillation itself. Furthermore, because of the short period, it may not be possible for the pilot to apply unstable in-phase corrections due to pilot response lag and control system lag producing a negative damping effect. This is referred to as pilot induced oscillations. Their magnitude can reach alarming proportions extremely quickly and are most likely to occur under conditions of high dynamic pressure eg, high speed/weight at low altitude and excessive pilot response lag or tendency to over control. If pilot induced oscillations are encountered, the conditions should be controlled or released to allow inherant dynamic stability of the aircraft to damp out the oscillations.

Chapter 8 Page 54

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
105. The damping effect of the tailplane, essential to the maintenance of longitudinal stability, is reduced in the low-density atmosphere at altitude. For a given IAS, the TAS increases with altitude and therefore so does inertia. In other words the aerodynamic damping will have less effect

Lateral Dynamic Stability


106. Lateral motion of an aircraft is made up of three components that occur simultaneously, they are sideslip, yaw and roll. Although all three occur at the same time they can conveniently be considered separately, since their overall dynamic effects will vary according to the strength of each component.

Roll Damping
107. The rolling motion will initially only affect the angle of bank and will be quickly damped as explained in Static Lateral Stability, Paragraph 79.

Spiral Instability
108. When an aircraft drops a wing the magnitude of the righting forces due to fin area, dihedral, high wing, etc, determines the degree of lateral stability. The fin and keel surfaces, however, tend to yaw the aircraft into the airflow, in the direction of the lower wing. The higher wing, which is now on the outside of the turn (the result of the yaw) is travelling slightly faster than the lower wing and developing more lift. This adds to the rolling moment of the original disturbance, opposing the righting forces and, possibly, exceeding them. This in turn causes the bank angle to increase further.

Chapter 8 Page 55

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
109. If the situation described above exists, the aircraft will enter a progressively steeper diving turn, or spiral dive. This is known as spiral instability, and it occurs when an aircraft's directional stability is greater than its lateral stability. However, the roll and yaw rates will be reduced by roll and yaw damping, the cross-coupling moments will be consequently of a reduced magnitude and the spiral dive tends to become neutral or even stable. A reduction in fin area, reducing the directional stability, will reduce spiral instability. 110. Spiral instability is a common feature in many aircraft and is not of particular importance, since it is easily overcome with normal control correction. Nevertheless, if uncorrected, the spiral divergence will persist and speed increases as pitch attitude reduces, roll attitude increases and altitude is lost.

Dutch Roll
111. Dutch roll is a combination of yawing and rolling motions, and occurs more readily with swept wing than with straight wing aircraft. 112. The secondary effect of yaw is roll. This occurs because, with a yaw to the left, the left wing effectively slows down (less lift), whilst the right wing speeds up (more lift), causing a roll to the left. Wing sweepback increases the secondary effect due to a change in aspect ratio. This reduces the tendency of the wings to dampen the rolling motion, which then leads to sideslip. If the oscillatory yawing and rolling motion thus induced maintains constant amplitude a state of neutral Dutch roll exists. If the oscillations diminish in magnitude (without control input) a state of stable Dutch roll exists. If the amplitude of oscillations increases the aircraft is unstable. Dutch roll is difficult to counter manually and may be exacerbated to a hazardous degree if control inputs are out of phase with the rolling and yawing motion.

Chapter 8 Page 56

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
113. The factors leading to Dutch roll are a combination of relatively weak directional stability compared with lateral stability. Hence, any departure from wings level attitude is rapidly corrected, but the aircraft has little or no tendency to turn into the sideslip (and may even turn away from it). Thus the aircraft has a wallowing tendency. 114. It is invariably considered necessary with modern jet transport aircraft to relieve the pilot of the requirement to control Dutch roll by incorporating yaw dampers into the system. 115. A yaw damper is an electro-hydraulic system. The yaw is sensed by gyros, and the direction and rate of yaw is used to operate the hydraulic jacks controlling the rudder. Thus any unintentional yaw is dealt with before it becomes oscillatory.

Roll Dampers
116. In addition to yaw dampers, some aircraft have roll dampers that are also gyroscopically controlled to correct roll by applying either aileron or asymmetric spoiler. This is particularly useful in turbulent conditions and, whilst not specifically fitted to counter Dutch roll, will obviously assist in doing so.

Asymmetric Effects
117. The stability of an aircraft (its ability to recover from a disturbance) will be degraded if it is in an asymmetric condition, in terms of power or drag, at the time of the disturbance. The asymmetry can cause one wing tip to stall, resulting in roll and yaw to such an extent that the aircraft enters a descending spiral path, or spin. This was discussed in depth in the section on Stalling. 118. The engines of multi-engine aircraft are positioned equidistant from the aircraft centreline so that their thrust is symmetrical about the normal axis and does not produce a yawing moment.

Chapter 8 Page 57

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
119. In the event of an engine failure the resultant asymmetry of thrust and drag produces (with the exception of a few three-engine examples), a strong yawing moment towards the failed engine. The aircraft is thus sideslipping away from the failed engine and with high static lateral stability characteristics, would roll towards the failed engine. Under conditions of high power setting and low forward speed, i.e. on take-off or go-around, the ensuing wing drop may be further increased by a large corrective aileron deflection. The yawing moment created by the engine failure has to be countered with the use of considerable rudder, accentuated by the higher power/lower speed situation which adds more drag and reduces the available control authority. The reduction in power available increases response times to demands for increased power.

Effect of Altitude
120. With increasing altitude, aerodynamic damping decreases due to the reduction in the density of the air. The damping moment in pitch, provided by the tailplane for longitudinal stability, in yaw, provided by the fin for directional stability and in roll, by the differential lift caused by the upgoing and downgoing wings for lateral stability, is proportional to dynamic pressure and inversely proportional to TAS. As altitude increases, at a constant EAS, the TAS increases and consequently damping reduces, as shown at Figure 8-31.

Chapter 8 Page 58

G LONGHURST 1999 All Rights Reserved Worldwide

Stability
FIGURE 8-31 Effect of Altitude on Static Stability

121. As we know, a tendency to Dutch roll occurs when lateral stability is greater than directional stability, whereas spiral instability is common when directional stability is strong in comparison to lateral stability. With increasing altitude both lateral and directional stability reduce, however, directional stability reduces at a greater rate than lateral stability. Consequently, the tendency to Dutch roll increases and to spiral instability reduces with increasing pressure altitude.

Chapter 8 Page 59

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

High Speed Flight


Mach Number Definition Speed of Sound Indicated Mach Number Compressibility Wave Drag Control Deflection Angle of Attack Aerofoil Thickness Angle of Sweep Area Rule Influence of Mach Number on Stalling Angle Influence of Mach Number on CLmax Influence of Mach Number on Lift/Drag Ratio

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight Aerodynamic Heating Mach Buffet Buffet Boundary Tuck Under Mach Trim Lateral Stability Vortex Generators The Supercritical Wing Oblique Shock Waves Mach Cone Aircraft Weight Expansion Waves Centre of Pressure

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

High Speed Flight

1. Low speed aerodynamics is based on the assumption that air is incompressible. At low speeds the errors which result from this assumption are negligible, it is at true airspeeds in excess of 300 knots that the air is compressed sufficiently to warrant attention. By the time that the speed of sound is approached, the compression and subsequent expansion of the air is sufficient to affect the streamline pattern of airflow past the aircraft. 2. We already know all about the streamlines involved in low speed flight. Although the flow is somewhat different at supersonic speeds, the behaviour of the airflow is stable and predictable. It is principally in the transonic envelope (the range from the critical Mach number up to an aircraft speed slightly in excess of Mach 1.0), that control and stability problems arise, since the airflow consists of both subsonic and supersonic elements.

Mach Number Definition


3. The True Mach Number (M) is the ratio of the speed of an object to the speed of sound in the same part of the atmosphere. 4. True Mach number is therefore the ratio of true airspeed to the local speed of sound, and may be expressed as: TAS True Mach No = ----------LSS

Chapter 9 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Speed of Sound
5. The speed of sound is the speed at which a very small pressure disturbance is propagated through a fluid medium at a given temperature. 6. The speed of sound is proportional to the absolute temperature such that: The Local Speed of Sound (LSS) = 38.94 Temperature ( A ) 7. LSS. 8. Remember that 0C = +273A. Therefore, as temperature decreases with altitude so does the Combining the two formulae it can be seen that: TAS True Mach No = ----------------------------------------------------------------38.94 ( Temperature ( A ) )

Example 1
9. From the formula above it can be seen, for example, that if TAS is constant then true Mach number will change with varying altitude in the standard atmosphere since temperature, and thus LSS, will change. 10. 11. Find the local speed of sound at sea level in the International Standard Atmosphere. ISA msl temperature = +15C = +273 + 15 = +288A

Chapter 9 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


12. An aircraft flying at Mach 1.0 at ISA mean sea level would have to be flying at a true airspeed of 661 knots, since: LSS = 38.94 288 = 661knots 661 Mach1.0 = -------661

Example 2
13. An aircraft is to fly at Mach 0.85 at FL 350, where the ambient air temperature is -55C. What will be its true airspeed? 14. Absolute air temperature (A) = +273 55 = +218A TAS MachNo = ----------LSS LSS = 38.94 218 = 575knots TAS = MachNo LSS TAS = 0.85 575 = 489knots

Chapter 9 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Indicated Mach Number


15. As already explained, True Mach Number is the ratio of TAS to LSS. An instrument which would be capable of measuring this ratio would therefore need to be supplied with values of TAS and LSS. TAS can be derived from an air data computer, and LSS can be computer-modelled. An instrument which would measure True Mach Number would thus be computer based. 16. TAS is a function of pitot excess pressure (the difference between pitot and static pressure) and density. LSS is a function of static pressure and density. Density is common to both functions. Thus TAS and LSS may be expressed as pressure ratios, and it is this that the Machmeter measures. It employs an airspeed capsule (CAS) and an altimeter capsule (static pressure) working in combination. dynamic pressure Indicated Mach Number is thus proportional to ----------------------------------------static pressure 17. From this it may be seen that, for example, flight at constant CAS and constant flight level will give a steady machmeter reading, since neither dynamic nor static pressure are changing, even if temperature changes. It may also be seen that, for example, at constant CAS in the standard atmosphere the machmeter reading will change if altitude changes since the static pressure will change. 18. A full explanation of the working of a machmeter may be found in the Flight Instruments book, Chapter 1.

Chapter 9 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Compressibility
19. When considering flight at low subsonic speeds it is convenient to regard air as an incompressible fluid, since the pressure changes which occur are relatively small and the consequent changes in density are negligible and can be ignored. At higher flight speeds the pressure variations which occur as air flows over the aeroplane are much greater and give rise to significant changes in air density. 20. Hence, when considering high-speed flight the air must be regarded as a compressible fluid and account taken of the compressibility effects on the aircraft and particularly on the lift-producing surfaces. Any body moving through the air creates pressure waves, which travel away from the body in all directions at the speed of sound. If an aircraft is travelling at less than the speed of sound the pressure waves ahead of the aircraft can travel away from it, as illustrated in the diagram at Figure 9-1.

Chapter 9 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-1 Pressure Wave Propagation

21. However, if the aircraft is travelling at the speed of sound the pressure waves ahead of the aircraft can no longer escape and a compression wave builds up at the leading edge. All changes in pressure and velocity occur across this compression wave and consequently occur much more rapidly than in the subsonic case, where changes begin ahead of the aircraft. Because of this the compression, or pressure, wave is often referred to as the shock wave. The effect is illustrated in the diagram at Figure 9-2. The shock wave is formed perpendicular, or normal, to the direction of flight and is known as a Normal Shock Wave. The flow immediately behind a normal shock wave is subsonic.

Chapter 9 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-2 Normal Shock Wave

22. Since the shock wave forms at the leading edge, the pressure and velocity distribution over the wing behind the wave is altered. This can cause significant variations in local values of lift and drag, altering the pitching moment on the wing and therefore the aircraft trim. These variations and the disturbed airflow arising from them may also affect control operation.

Chapter 9 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


23. It should be noted that the aircraft does not necessarily have to be flying at Mach 1.0 to experience compressibility effects. Because the function of an aerofoil is to accelerate the airflow over it, the air speed at some point on the aerofoil can be sonic or supersonic even though the aerofoil is moving through the air at less than the speed of sound. In other words, it is possible to have both subsonic and supersonic airflow on different parts of the aircraft at the same time. This is known as transonic flight and, as a general rule, occurs between Mach 0.75 and Mach 1.2. Below Mach 0.75 the airflow on all parts of the aircraft is subsonic, whilst above Mach 1.2 it is supersonic. 24. The free stream Mach number (MFS) is the Mach number of the airflow at a point unaffected by the pressure of the aircraft, but measured relative to the aircraft. 25. The local Mach number (ML) is the speed of the air relative to LSS measured at a specified point on the aircraft. The local Mach number may be greater than, the same as, or less than MFS. 26. The critical Mach number (MCRIT) is the lowest free stream Mach number at which a local Mach number of 1.0 will occur at any point on the aircraft. MCRIT varies with the angle of attack, since the higher the angle of attack the greater the acceleration of air over the top of the wing, and therefore the sooner the air over the wing will go supersonic. MCRIT marks the lower end of the speed range within which ML may be either subsonic or supersonic. 27. Compressibility Mach number is the Mach number at which compressibility effects cause significant control problems, and beyond which loss of control is probable. 28. Finally, critical drag Mach number relates Mach number to an appreciable increase in drag that is associated with compressibility effects. The critical drag Mach number is normally 110% to 115% of MCRIT.

Chapter 9 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


29. Remember that the free stream Mach number (MFS) can be well below Mach 1.0 for a shock wave to form since, providing the speed is in excess of MCRIT, ML on the upper surface will be in excess of 1.0. 30. When this occurs a shock wave forms over the upper surface of the wing because the pressure waves from the rear of the wing are trying to move forward, and are meeting air travelling backwards at exactly the same speed. The point at which this shock wave usually forms is just aft of the point of maximum camber of the wing where the acceleration of the air is greatest. In front of the shock wave the flow is at or beyond Mach 1.0, whilst behind the shock wave the flow is still subsonic. It should be appreciated, therefore, that a shockwave forms where the airflow goes subsonic, not supersonic 31. At the shock wave, the normal laws of physics seem to break down, and as the air passes through the shock wave the pressure increases, as does the temperature and density. If the speed of the aircraft is increased still further the region of supersonic flow on top of the wing also increases and the shock wave will start to move back towards the trailing edge. On the under surface the curvature of the wing is usually less than on the upper surface, and the shock wave will form later. However, once having formed, if the actual speed of the aircraft is further increased, the lower surface shock wave will also move rearward. 32. As MFS reaches 1.0 both the upper and lower shock waves arrive at the trailing edge of the wing. At some time another shock wave, known as the detached bow wave, forms just in front of the leading edge. As MFS is increased well above 1.0 the forward shock wave moves back to actually touch the leading edge, and is now known logically enough, as the attached bow wave. At this point both the forward and aft shock waves are at the position at which they will remain as speed is further increased. All that will happen now is that, as speed is further increased, the waves will bend further backwards at points removed from the wing itself.

Chapter 9 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


33. These effects are summarised at Figure 9-3.

Chapter 9 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-3 Shock Wave Formation

Chapter 9 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


34. The bow wave and tail wave of an aircraft at supersonic speed are sufficiently strong to reach the surface of the earth causing a sonic boom. Whilst normally heard as one disturbance, there are, in fact, two very close together. The actual pressure disturbance is small. For example, Concorde at mach 2 and 52000 ft altitude leaves an overpressure of 1.94 pounds per square foot over the surface, which is not great compared to the standard atmospheric pressure of 14.7 pounds per square inch. The mach cone angle at mach 2 is 60 degrees. This is illustrated at Figure 9-4.

FIGURE 9-4 Mach Cone

Chapter 9 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Wave Drag
35. For an aircraft that is not designed for transonic or supersonic flight, the formation of shock waves will have a dramatic effect on both lift and drag. The stability of the aircraft may become critical, and control difficult, or in the extreme impossible. 36. Basically, the cause of these problems is the separation of the airflow behind the shock wave due to the rise in pressure that occurs. Separation within the boundary layer creates a considerable increase in drag and reduction in lift. The drag thus produced is termed wave drag, and is particularly marked at MCRIT. As the speed is further increased the shock wave moves towards the trailing edge, with a resultant decrease in wave drag and increase in lift. As a MFS of 1.0 is achieved the shock wave reaches the trailing edge. Beyond a MFS of 1.0 the problems of transonic flight are left behind (literally), although the amount of lift produced will be lower than at subsonic speed for a given angle of attack. This is because the co-efficient of lift is reduced in supersonic flight, due to the presence of the bow wave. 37. Figure 9-5 shows the variation of the co-efficient of drag through the transonic phase and Figure 9-6 shows the variation of the co-efficient of lift against Mach number.

Chapter 9 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-5 Variation of C D Through the Transonic Range

Chapter 9 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-6 Variation of C L Through the Transonic Range

38. In order to reduce the effect of wave drag, shock waves must be kept as weak as possible. To achieve this the wings of supersonic aircraft are designed with a sharp leading edge to minimise the bow wave, together with a thin wing with minimal camber to reduce the adverse pressure gradient. This is discussed in further detail in the subsequent section entitled Aerofoil Thickness.

Chapter 9 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Control Deflection
39. The effect of the conventional hinged flap type of control surface when deflected is to create an aerodynamic force ahead of the control that is usually greater than that produced by the control surface itself. When a shock wave is established on the associated lifting surface ahead of the control surface this is no longer the case, since pressure variations cannot be transmitted forward through a shock wave. Consequently the effect of deflecting the control surface is reduced, since only the flow over the control surface is modified. When the shock wave becomes established at the trailing edge, the effectiveness of the control is even further reduced. 40. This has led to the replacement of the tailplane and elevator configuration with the allmoving horizontal stabiliser in high-speed aircraft. As has been previously mentioned, instead of ailerons at the wing tips, differential spoilers and/or inboard ailerons are used in the transonic region of flight. This is partly due to the effects described in the previous paragraph and partly due to the wing twist caused by aileron deflection at high speed, which can negate or even reverse the effect of the ailerons. 41. When the shock wave is actually located on the control surface its deflection will cause the wave to move, with consequent large alterations of the position and magnitude of the aerodynamic forces acting on the control. This may result in high frequency vibration of the control surfaces, known as control buzz. 42. Deflection of the controls during transonic flight usually requires much greater input forces from the pilots controls. This is because the hinge moment of the control surface is changed by the changing forces as the shock wave forms first ahead of the control, then at the hinge and finally at the trailing edge. Almost invariably power-operated controls are essential in order to overcome the large and variable stick forces encountered in aircraft designed to operate in this regime.

Chapter 9 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Angle of Attack
43. Increasing the angle of attack has the effect of reducing the critical Mach number (MCRIT) of the aircraft. This is because as angle of attack increases, the local peak velocity over the wing is increased. This is of particular significance when one considers that effective angle of attack increases during manoeuvring, hence manoeuvres when flying at high speed may induce the formation of a shock wave at a free stream Mach Number lower than the steady level flight value.

Aerofoil Thickness
44. In order to avoid the complications of flight in the transonic region, it is clearly highly desirable for aircraft that are to fly at high subsonic speed to have a high critical Mach number. One of the ways of increasing the critical Mach number is to reduce the thickness/chord (t/c) ratio of the aerofoil sections, which reduces the flow acceleration over the section. The benefits are further enhanced if the camber is kept low, with the point of maximum thickness at about 50% chord. This has the effect of reducing the intensity of the shock waves when they form and ensuring that they form well back towards the trailing edge. Thus less of the wing surface is affected by the adverse conditions behind the wave. 45. It will be remembered that thickness/chord ratio is the ratio of maximum aerofoil thickness to chord length, expressed as a percentage. This can therefore be achieved with a thin wing of narrow chord, or a thicker wing of broad chord. The disadvantages of a thin aerofoil section are that it is susceptible to flexure and twisting and it has a lower maximum lift coefficient. Because of its thinness it is difficult to incorporate the chord-extending trailing edge flaps necessary to improve the lift coefficient at low flight speeds.

Chapter 9 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


46. A broad chord wing, such as the delta planform of the Vulcan bomber, can be made thicker for the same t/c ratio, but has its own structural and low-speed handling limitations, that render it generally unsuitable for high volume passenger transport aircraft. 47. Consequently a compromise has to be reached, giving the highest practicable t/c ratio in a wing capable of incorporating adequate lift augmentation systems for landing and take-off. The maximum MCRIT for subsonic transport aircraft is usually of the order of 0.85. From the graph at Figure 9-7 it will be seen that this can be achieved with a straight wing of 7.5% thickness/chord ratio.

Chapter 9 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-7 Effect of Thickness/Chord Ratio on MCRIT - Straight and Swept Wing

Chapter 9 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Angle of Sweep
48. From the graph at Figure 9-7 it will be seen that the critical Mach number for a wing of given thickness/chord and aspect ratios can be greatly increased by sweep back. Given the stowage requirements for fuel, flaps and possibly landing gear in the wings of a transport aircraft, together with the requirement for a reasonably high aspect ratio, a t/c ratio of at least 10% - 15% is necessary. A 45 sweep back as shown in Figure 9-7 is generally impractical for a large transport aircraft, but even with a much more modest sweep angle a satisfactorily high critical Mach number can be obtained at the t/c ratio required. 49. The control effect produced by ailerons on a swept back wing can be considerably reduced due to the fact that they are not mounted perpendicularly to the relative airflow, and this is illustrated at Figure 9-8. Also, the effect of ailerons can be reduced in high-speed flight due to the twisting effect on the aircraft's structure, as previously mentioned. 50. For the foregoing reasons many jet aircraft are fitted with two ailerons per wing. The outboard ailerons are used for low speed flight in the conventional manner, and have the advantage of a long force couple. During high-speed flight the outboard ailerons are locked and the inboard (or high-speed) ailerons are used. 51. The force couple is very much reduced with the inboard ailerons, but this is offset by the high speed, and anyway rapid rates of roll are not required during high-speed flight. The important thing is that, since the high-speed ailerons are mounted on a structurally rigid (thick and deep) part of the wing, any wing twist that is produced will be minimal. The function of the spoilers, which are shown at Figure 9-8 was discussed in Chapter Six.

Chapter 9 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-8 Primary and Secondary Controls on a Swept Wing

Chapter 9 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


52. Broadly speaking, the major advantage of a swept back wing is that of the increase in the critical Mach number, but it does also reduce both the rate at which the drag coefficient rises with increasing Mach number, and the maximum drag coefficient. This is illustrated in the graph at Figure 9-9. It will also be noted from Figure 9-9 that beyond about Mach 2.0 the advantage of sweep back is minimal.

FIGURE 9-9 Effect of Mach Number on C D

Chapter 9 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


53. Many of the disadvantages of swept wings are solved by the use of the variable geometry wing (as in the Tornado) whereby the sweep can be altered mechanically. This is a rather complex system involving an increase in weight and structural complexity and for this reason, to date, it has been confined purely to military aircraft.

Area Rule
54. It can be shown theoretically that there is a certain ideal shape for a transonic or subsonic aircraft that will produce minimum drag. It was stated earlier that reducing the thickness and camber of the wing, thereby avoiding sharp pressure and velocity peaks, could reduce the intensity of the shock wave on a wing. The same is true of the aeroplane as a whole and the ideal graphical distribution of cross-sectional area along the longitudinal axis is a smooth curve, as illustrated at Figure 9-10.

Chapter 9 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-10 Ideal Body of Revolution

55. If the cross-sectional area of a conventional aircraft is plotted against its longitudinal axis, the wings and empennage (tail surfaces) tend to appear as lumps on the curve. However, if the fuselage contours are increased ahead of and behind the wings and empennage, a smooth curve can be obtained as shown at Figure 9-10.

Chapter 9 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-11 Area Rule

Chapter 9 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Influence of Mach Number on Stalling Angle


56. As Mach number increases the stalling angle of attack decreases and a large increase in Mach number can induce a premature stall. This is because of the loss of lift that occurs due to the disturbed airflow behind a shock wave. The effect of flight at high subsonic speeds on lift coefficient, compared to the theoretical incompressible flow case, is shown in the graph of CL - at Figure 9-12.

FIGURE 9-12 Effect of Mach Number on Stalling Angle

Chapter 9 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Influence of Mach Number on CLmax


57. In addition to the reduction in stalling angle with increasing Mach number, the maximum lift coefficient is also reduced, as shown in the graph at Figure 9-13.

FIGURE 9-13 Effect of Mach Number on


CLmax

Chapter 9 Page 27

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Influence of Mach Number on Lift/Drag Ratio


58. The variation of CL and CD in the high subsonic and transonic regions was illustrated at Figure 9-6 and Figure 9-5. From these it can be seen that, in this range, CL falls dramatically at the formation of the shock wave (the shock stall) and never fully recovers to its subsonic value. At the same time CD is rising, so it is clear that, in the transonic range, the lift/drag ratio (CL CD) falls markedly.

Aerodynamic Heating
59. As any body moves through the air the surface friction creates heating. The faster the body moves the greater the friction and therefore the greater the heat generated. In addition to this, pressure builds up, particularly at the stagnation points, and from the gas laws we know that when a gas is compressed its temperature increases. Once again, the higher the airspeed the greater the total pressure and the greater the resultant temperature. 60. At speeds below Mach 2.0 this heating effect is of no great importance from a structural point of view, but beyond that it is necessary to use heat resistant structural materials. Between Mach 1.0 and Mach 2.0 the surface temperature of the aircraft can rise to a point where maintaining a comfortable cabin temperature for passengers becomes problematical. 61. Figure 9-14 shows graphically the surface temperature rise with Mach number at an altitude of 28,000 ft, where the ambient air temperature is -40C.

Chapter 9 Page 28

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-14 Effect of Mach Number on Surface Temperature

62. Various methods can be used to dissipate the heat generated, or to insulate the aircraft interior from it. Very high-speed military aircraft and space re-entry vehicles employ surface materials able to withstand the high temperatures and which, in some cases, encourage heat dissipation by radiation. Concorde employs a system in which fuel is circulated through channels just beneath the exterior surface to dissipate heat by conduction.

Chapter 9 Page 29

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Mach Buffet
63. The turbulent airflow that occurs behind a shock wave is liable to cause severe buffeting, especially if this turbulent flow affects the horizontal stabiliser surfaces. The symptoms are very similar to that of the stall, hence the term, shock stall, applied to the phenomena. The major difference between this and the stall discussed in Chapter Four is that the conventional stall occurs because of excessive angle of attack and is correctly referred to as the high-incidence stall. The shock stall is much more likely to occur at low angles of attack and is caused by shock-induced separation. (See Figure 9-3). 64. Another cause of Mach buffet, noticeable particularly with supercritical aerofoil sections, is a tendency for the relatively weak shock wave to move rapidly backward and forward on the surface of the wing as MCRIT is reached.

Buffet Boundary
65. Because of the high altitude at which jet transport aircraft normally operate the cruising speed, which is ideally just below MCRIT, and the high-incidence stalling speed are remarkably close to each other, sometimes less than 20 knots separates the two. Because of the low temperature at the operating altitude the local speed of sound is low and the cruise TAS will be very close to that needed to achieve MCRIT. The very narrow margin between the clean aircraft high-incidence stall speed and the shock stall speed is known as the buffet boundary. It is also known colloquially as coffin corner.

Chapter 9 Page 30

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Tuck Under
66. In the transonic range there is usually a marked change in the pitching moment, producing a nose-down effect. This can be due to a number of factors. The position of the shock waves on the upper and lower surfaces is not usually coincident between MCRIT and Mach 1.0, the lower one usually being forward of the upper. This tends to move the centre of pressure aft, creating a nosedown moment. In swept wing aircraft shock waves tend to form on the thicker inboard wing sections first, which means that there is a loss of lift near the roots and so the centre of pressure moves aft because of the sweep back. Finally, loss of downwash over the horizontal stabiliser reduces its balancing downward force resulting in an increased nose-down pitching moment. This nose-down pitching tendency at high subsonic/transonic speeds is sometimes referred to as tuck under.

Mach Trim
67. The effect of the pitch down moment described in the preceding paragraph is to tend to make the aircraft speed unstable as well as longitudinally unstable. That is to say, failing any corrective action, the nose-down pitch causes the speed to increase, which causes further nose-down pitch, which causes the speed to increase and so on. The problem continues beyond the transonic envelope, since the centre of pressure continues to move rearward. 68. Since the condition of speed instability is inherently a very dangerous one, supersonic, transonic and high-speed subsonic aircraft are fitted with a Mach trimming system. 69. The Mach trim is a fully automatic trimming device which is sensitive to Mach number, and which moves the elevator/stabiliser to compensate for the pitching movements which are expected at any given Mach number.

Chapter 9 Page 31

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Lateral Stability
70. Shockwave formation does not usually occur at precisely the same Mach number and location on each wing and consequently one wing starts to drop when MCRIT is exceeded due to the reduction in lift it produces. Furthermore, laterally stationary design features (dihedral, sweepback) which increase the lift of the downgoing wing in the sideslip produce a greater local acceleration of the flow and this higher local Mach number may well cause further wing drag by creating a shockwave or intensifying an already existing shockwave or that downgoing wing. 71. Any asymmetry in the formation of shock waves on the wings will result in a tendency for a wing drop to occur between MCRIT and a MFS of 1.0.

Vortex Generators
72. These devices, which we have already met in previous chapters, are not only useful in limiting spanwise flow to prevent tip stalling. It will be recalled that the boundary layer breaks away when it is no longer able to overcome the adverse pressure gradient over the upper surface of the wing. The formation of a shock wave makes matters worse since the speed of the boundary sub-layer is subsonic so the high pressure at the wave can be transmitted forward. 73. As we know, the vortex generators are small plates projecting from the upper surface of the wing and they create vortices that re-energise the boundary layer. In addition, they appear to weaken the shock wave and so reduce shock drag. Shock buffet is reduced by the vortices they produce. 74. Other vortex inducing devices such as the saw tooth leading edge, the wing fence and the notched leading edge achieve the same result. A thickened trailing edge is also sometimes used to create vortices at the point where the boundary layer is thickest and most sluggish.

Chapter 9 Page 32

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

The Supercritical Wing


75. Figure 9-15 shows a supercritical wing section overlaid upon a conventional aerofoil.

FIGURE 9-15 Supercritical Wing

76. The upper surface is relatively flat, whereas the lower surface has more pronounced curvature. The airflow does not achieve the same increase of speed over the upper surface and therefore the formation of a shock wave is delayed until a higher MFS. The critical drag Mach number (MCD) is thus also increased. 77. If the upper surface leading edge is very carefully profiled then any expansion waves will reflect from the supersonic/subsonic boundary back to the shock wave and reduce its intensity. The principle is shown at Figure 9-16.

Chapter 9 Page 33

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-16 Effect of Leading Edge Shape on Shock Wave Intensity

78. Some of the lost lift from the flattened upper surface is recovered by a pronounced reflex camber near the lower surface trailing edge. The section is rear-loaded in its lift generation, and is therefore often referred to as an aft-loaded wing. 79. The section profile is very critical, particularly the leading edge and the early part of the upper surface. In order to keep the upper surface free of joints, any leading edge devices will be of the Krueger type, which fold out from the under surface. Note the blunt leading edge. 80. The advantage of the supercritical wing is that high Mach number cruising speeds can be maintained with a thicker (stronger) wing, which requires less sweepback.

Chapter 9 Page 34

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Oblique Shock Waves


81. When a body is moving through the air at sonic speed (the speed of sound) the shock wave produced is normal to the direction of travel, as we have seen. When speed is increased beyond the speed of sound (i.e. supersonic) a different wave pattern is produced. 82. The shock wave moves away from its source at the speed of sound. Imagine now a body moving at greater than sonic speed along a straight path. As it travels the pressure wave it generates is not only moving outwards from the body, but it is also being left behind. This is illustrated at Figure 9-17.

Chapter 9 Page 35

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-17 Oblique Shock Waves

83. From Figure 9-17 it can be seen that the shock wave produced is oblique to the direction of travel of the source and it is consequently called an Oblique Shock Wave. The angle it makes with the direction of travel is known as the Mach Angle. The higher the supersonic speed the shallower the mach angle will be.

Chapter 9 Page 36

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight

Mach Cone
84. If the shock source is a single point, such as the nose of a bullet, the shock pattern will be three-dimensional and so the shock boundary will be cone-shaped. If the source is a straight line, such as the leading edge of a wing, a wedge will be formed. In either case the angle at the apex will be twice the mach angle.

Aircraft Weight
85. Buffet boundary and all reference speeds, such as MCRIT and maximum operating Mach number (MMO), are absolutely dependent upon aircraft weight. Basically, the greater the weight, the lower the reference speed because a greater angle of attack is needed for any given speed. This means that, as fuel is burned off during flight and aircraft weight reduces, reference speeds gradually increase permitting increased aerodynamic ceiling.

Expansion Waves
86. When a supersonic flow meets a convex corner it is free to expand, which results in a decrease in pressure, density and temperature as the flow lines become further apart. At the same time the velocity of the flow increases. These features are indicated in the diagram at Figure 9-18.

Chapter 9 Page 37

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-18 Expansion Wave

87. From Figure 9-18 it will be seen that, because of the velocity increase that takes place as the air flows around the corner, the Mach number of the flow increases. Also, the angle between the new Mach line and the direction of flow has become more acute. This is to be expected since, as we now know, the higher the speed of the supersonic flow the more acute the mach angle. 88. It should be emphasised at this point that the Mach lines are not shock waves. Shock waves are formed at a point where a number of mach lines converge, as would be the case when a supersonic flow meets a concave corner, as at Figure 9-19.

Chapter 9 Page 38

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


FIGURE 9-19 Shock Wave

89. Mach lines are weak waves that may form at any point on a surface in a supersonic flow and the flow passes through them without undergoing any sudden physical changes. When the supersonic flow meets a concave corner as at Figure 9-19 the convergence of the Mach lines produces a strong oblique shock wave, through which there is a sudden increase of pressure, temperature and density and a decrease of velocity. This is known as compressive flow. 90. Conversely, the phenomenon occurring at a convex corner is called expansive flow and the weak wave associated with it is known as an expansion wave. The changes in pressure, temperature and density that occur at an expansion wave are relatively gradual, when compared to a compression shock wave. In comparison with subsonic flow however, where changes begin well before the corner is reached and continue well after it is passed, they are still relatively rapid.

Chapter 9 Page 39

G LONGHURST 1999 All Rights Reserved Worldwide

High Speed Flight


91. It is possible to turn supersonic flow through a large angle by means of a series of expansion waves created by a succession of convex corners. A curved surface is, of course, an infinite succession of corners and over such a surface there will be a series of expansion waves. As supersonic flow passes over the curved surface there will be a succession of expansion waves in which the physical changes will be even more gradual than at a single corner. 92. The designer of the supersonic aircraft seeks to include convex curves in the design wherever possible, since this will not only reduce the shock wave effects in supersonic flight, but is also beneficial in subsonic flight.

Centre of Pressure
93. During transonic flight the centre of pressure moves progressively aft, typically from about 20% chord at Mach 0.75 to about 50% chord at Mach 1.4. In supersonic flight this rearward migration of the centre of pressure continues and is one of the factors limiting the maximum Mach number achievable with a swept wing planform. 94. By careful design it is possible to retard the rearward movement of the CP. In Concord, for example, the aft movement of the CP is limited to about 8% of total chord, but this is still a considerable distance and has to be counteracted by transference of fuel to move the centre of gravity (CG) aft a corresponding amount.

Chapter 9 Page 40

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Limitations
Operating Limitations Control Reversal Flap/Landing Gear Operation Limiting Speeds The Manoeuvring Envelope The Gust Envelope The Buffet Onset Boundary Chart

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations

10

Limitations

Operating Limitations
1. Any aircraft is designed in the first instance to meet certain performance requirements in terms of payload, operating speed, operating altitude and so on. The designer goes to great lengths to ensure that there is adequate structural integrity to meet those design requirements and then formulates the parameters within which the aircraft can safely be operated. In so doing he will incorporate due allowance for unusual factors such as gusts due to rough air or turbulence, by making the airframe structurally flexible. 2. The necessary flexibility of the structure can of itself cause control problems and, in extreme cases, lead to structural failure.

Control Reversal
3. Control reversal is a problem that arises from the aerodynamic force overcoming torsional stiffness and causing wing twist. When an aileron is deflected downwards it increases the lift coefficient of the wing at that point. The increased lift is acting a point on the chord well to the rear of the flexural centre and there is a tendency for the wing to twist. This is especially so with outboard ailerons, located as they are at the point of least torsional stiffness. 4. The greater the airspeed the greater the lift produced by a downgoing aileron and so the greater the wing twist, eventually to the extent that the overall angle of attack at the tip is negative the reverse of the intended effect! It is only likely to occur on aircraft not designed for high subsonic or transonic flight when a limiting airspeed is exceeded.

Chapter 10 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations

Flap/Landing Gear Operation


5. Extending the flaps significantly increases the wing lift generated and therefore increases the wing bending moments. There is a limiting speed above which the flaps must not be extended, to ensure that the structural limitations of the wings and flaps are not exceeded. 6. Extending the landing gear considerably increases the profile drag of the aircraft and alters the thrust/drag relationship. There is a limiting airspeed for landing gear extension to avoid imposing excessive loads on the gear attachment points and to avoid possible pitch control problems.

Limiting Speeds
7. For every aircraft there are operating speeds designed to ensure that potentially hazardous structural and control problems are never encountered.

VMO
This is the maximum operating speed and is normally expressed as IAS. It is the limiting speed at low altitude because it will be achievable before limiting mach number in dense air. When descending at constant mach number IAS will be increasing, and therefore VMO may be exceeded. It is the maximum permitted speed for all operations.

VNO
This is the normal operating speed. It is the maximum permitted speed for normal operations.

Chapter 10 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations

VNE
This is the Never Exceed Speed. It is a higher maximum permitted speed when operationally desirable.

MMO
The same as VMO, but quoted as a Mach number. It is the limiting speed at high altitude because it will be achievable before limiting IAS in less dense air. When climbing at constant IAS the mach number will be increasing, and therefore MMO could be exceeded.

VLO
This is the maximum speed at which the landing gear may be lowered or retracted.

VLE
This is the maximum speed with the landing gear extended. It may be the same as VLO or may be different, depending on the design of the gear and doors. On some designs the doors are open whilst the gear is moving and retract when the gear is locked down and in such cases it may be found that VLE exceeds VLO.

Chapter 10 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations

VFE
This is the maximum speed for lowering flaps to various flap settings. It is possible to incorporate automatic flap positioning and load-limiting devices based upon airspeed or dynamic load inputs to ensure that the flaps cannot be overstressed.

The Manoeuvring Envelope


8. The manoeuvring envelope represents in graphical form the operating limits of an aircraft. It is developed initially during the design stage for a new aircraft type and is then used subsequently to define the limits which, if exceeded, will result in stalling, structural damage or even structural failure. 9. The aeroplane must be constructed to withstand the loads to which it will be subjected during normal operations. These are defined as the load factor, or g force, acting upon the aircraft. In straight and level flight in calm air the load factor, or g force, is 1 because the lift (L) is equal to the weight (W). From chapter four (Stalling) it will be recalled that: L n = ---W 10. During manoeuvres the load factor (n) increases. Imagine the situation when, for example, the aircraft is pulling out of a dive. In order to arrest the dive the angle of attack is increased and the lift force acting upon the wings exceeds the aircraft weight during the period of pull-up, increasing the loading on the wings.

Chapter 10 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
11. The more abrupt the pull-up the greater the angle of attack and therefore the greater the wing loading, or load factor n. Exactly the same is true during a tight turn or, for that matter, during any manoeuvre. When manoeuvring, the lift and weight will not be equal. When lift exceeds weight the load factor, or g acceleration is said to be positive, when weight exceeds lift the load factor is said to be negative. The latter is the case during a bunt (or outside loop) manoeuvre or when making an outside turn. So, in straight and level (non-manoeuvring) flight the load factor is +1. 12. As load factor increases the stalling speed increases and this establishes the first of the manoeuvring operating limits. 13. From chapter four we know that the maximum lift at any airspeed occurs at maximum lift coefficient (CLmax) and that maximum lift occurs at the stall. Therefore: 1 2 -V S L max = C Lmax -2 m 14. 15. Where Vm = manoeuvre stall speed. At the basic stalling speed (Vs) weight (W) must be equal to maximum lift (Lmax). Therefore: 1 2 -V S W = CLmax -2 s

Chapter 10 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
16. Therefore: 1 2 2 - V S C Lmax -Vm 2 m L - = ------n = ---- = ---------------------------------2 1 2 W Vs CLmax -- Vs S 2 17. So: Vm = Vs n 18. For example, the manoeuvring stall speed at a load factor of 2 for an aircraft with a basic Vm =
110

stalling speed (Vs) of 110 knots would be: 19.

2 = 155 knots

At a load factor of 3 it would be 190.5 knots, at load factor 4, 220 knots and so on.

20. From this it is possible to plot the manoeuvring stall speeds against load factor on the manoeuvring load diagram, or V n diagram as it is sometimes known, as shown at Figure 10-1. It will be noted from this that it is possible to fly at less than the basic stalling speed provided the load factor is less than one. This is of course exactly what is happening when we see simulated weightlessness during astronaut training, as the training aircraft flies a zero g bunt after initiating a steep pitch angle climb.

Chapter 10 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
FIGURE 10-1 Manoeuvre Stalling Speed

21. The load factors to which an aircraft is designed to safely operate depends upon the role it is to perform. Aeroplanes of large mass will, because of structural limitations, have lower limits than smaller aeroplanes. The values are published in JAR 25 (large aeroplanes) and JAR 23 (normal, utility, aerobatic, and commuter). They are as follows:

Chapter 10 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations JAR 25
22. The positive limit manoeuvring load factor (n) for any speed up to VD may not be less than 24000 2.1 + -------------------------- except than n may not be less than 2.5 and need not be greater than 3.8 where W W + 10000 is the design maximum take off weight (lbs). The negative limit manoeuvring load factor: (1) (2) may not be less than -1.0 at speeds up to VC and must vary linearly with speed from the value at VC to zero at VD.

Manoeuvring load factors lower than those specified may be used if the aeroplane has design features that make it impossible to exceed these values in flight.

Chapter 10 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations JAR 23
23. The positive limit manoeuvring load factor may not be less than: (1) 24000 2.1 + -------------------------- for normal and commuter category except that n need not W + 10000 be more than 3.8 4.4 for utility category aeroplanes or 6.0 for aerobatic category aeroplanes.

(2) (3)

The negative limit manoeuvring load factor may not be less than: (1) (2) 0.4 times the positive load factor for the normal, utility, and commuter categories or; 0.5 times the positive load factor for the aerobatic category.

Manoeuvring load factors lower than those specified may be used if the aeroplane has design features that make it impossible to exceed these value in flight. Note. In all cases a 50% safety factor is usually built in. 24. From this we can now add two more limits to the manoeuvring load diagram, the design maximum positive and negative load factors. Beyond these load factors some structural damage or deformation may occur.

Chapter 10 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
25. If the maximum load factor, positive or negative is exceeded by more than 50% then structural failure will occur, giving us a further two limits to add to the diagram, the design ultimate positive and negative load factors. This is illustrated at Figure 10-2.

FIGURE 10-2 Manoeuvre Load Factor Boundaries

Chapter 10 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
26. Finally, every aircraft has a maximum operating speed that must not be exceeded. Adding this vertical line to the manoeuvring load diagram completes the limits that form the Manoeuvring Envelope, as shown at Figure 10-3.

FIGURE 10-3 Manoeuvre Speed Limit

Chapter 10 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
27. Figure 10-4 shows the manoeuvre envelope for a complete aeroplane in the large category. Note that the stall buffet boundary (CN MAX) with flaps down has been added. Also shown are significant speeds.

FIGURE 10-4 Manoeuvre Envelope

VD is the design dive speed. At altitude it may be limited by mach number.

Chapter 10 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
VC is the design cruising speed. At altitudes where VD is limited by mach number, VC may be limited to a selected mach number. VS1 is the unaccelerated stall speed in the clean configuration. VA is the design manoeuvring speed. It is the highest speed at which the aeroplane will stall before exceeding the maximum load factor. Its maximum value may be calculated by multiplying VS1 by the square-root of the maximum load factor. As aeroplane mass reduces VA will reduce as VS1 reduces. It may be calculated as follows: Given VS1 160 kts. Load factor limit +2.5. Calculate VA. VA = 160 2.5 = 253kts Changes of VA with changes of weight may also be calculated as follows: Given VS1 160 kts. Load factor limit +2.5. VA 253 kts. Weight 8000 kgs. Calculate VA at weight 6500 kgs. It is necessary to calculate the new VS. 6500 New V S = 160 ----------- = 144kts 8000 New V A = 144 2.5 = 227kts

Chapter 10 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
28. The load placed upon the aircraft structure during flight is dependent upon aircraft weight and the manoeuvring load diagram is based upon a particular design gross weight. If the aircraft is operated at a higher gross weight the load factor for any given airspeed will be greater. In other words, the maximum load limit will be encountered at a lower airspeed. 29. It will be noted that the airspeed used for the design case is EAS, whereas at altitude the pilot is usually much more concerned with TAS. Since the relationship between the two airspeeds is relative air density, at altitude any given value of TAS is equivalent to a much lower value of EAS. The design EAS values of VA, VC and VD are reduced for altitudes above 20,000 feet, usually as a step reduction. 30. As altitude is increased the TAS/EAS at which MCRIT is encountered decreases to a value less than VD, modifying the manoeuvring envelope. For aircraft designed to operate in the high subsonic or transonic regime a complete envelope will show Mach number limits for each altitude.

The Gust Envelope


31. It is possible to calculate the approximate load factor resulting from a change in angle of attack since the lift coefficient curve is mostly linear (refer to Chapter 2, Figure 2-7). From an initial condition of load factor 1 at a given angle of attack, then a given percentage increase in angle of attack will result in the same percentage increase in load factor. 32. During flight in rough air the aircraft will encounter gusts which will place loads upon the aircraft structure. It is the vertical component of any gust that is important in terms of load factor, since a vertical gust increases the effective angle of attack and therefore the magnitude of the lift generated. This, of course, increases the loading on the wings. The situation is illustrated at Figure 10-5.

Chapter 10 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
FIGURE 10-5 Effect of Vertical Gust

33. In determining the load factor that will result from an aeroplane encountering a vertical gust the designer must decide upon two factors, the speed of the aeroplane and the vertical speed of the gust. He therefore chooses a design speed at which he expects the aircraft to operate in normal circumstances (VB), and applies a vertical gust velocity based upon many years of recorded values. This arbitrary gust speed is 66 feet per second. 34. Were the aircraft to encounter a gust when flying at a higher speed, clearly the resultant load factor would be greater so the designer must allow for this also. However, the mathematical probability of this happening is less and so the arbitrary gust velocities applied are lower. At the design cruise speed (VC) the assumed gust value is 50 feet per second and at the design dive speed (VD) it is 25 feet per second. If these values are plotted on a load, or V n diagram as shown at Figure 10-6, a gust load diagram is the result. VB will be less than VD and need not be greater than VC.

Chapter 10 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
FIGURE 10-6 Gust Envelope

Chapter 10 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
35. The arbitrary vertical gust velocities are plotted from the load factor 1 point on the vertical axis of the graph, since this is the load factor in straight and level flight in smooth air. Where they intersect the chosen aircraft velocities the load limits are established. Joining these points sets the Gust Envelope within which the aircraft must operate in order to avoid stalling or exceeding the limiting load factor. 36. It should be noted that the gradient of the positive stall curve is shallower than in the case of the manoeuvring envelope. This is because a vertical gust, by increasing the effective angle of attack, increases the airspeed at which the stalling angle is reached. 37. From the gust load diagram another important reference airspeed is established, the recommended speed for flight in turbulence, or rough air speed (VRA). VRA must lie within the range of speeds for which the requirements associated with VB are met and must be sufficiently less than VMO to ensure that likely speed variations in rough air will not cause the overspeed warning to operate too frequently. 38. As with the manoeuvring envelope, the effects of aircraft gross weight, altitude and Mach number must be taken into account in the case of the gust envelope. The gust load diagram is drawn up for particular values of each and the complete envelope will show the effects of flight at varying altitudes and gross weights. In addition, wing aspect ratio and sweepback must be considered since a high aspect ratio will generally be less resistant to bending loads.

Chapter 10 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations

The Buffet Onset Boundary Chart


39. An aeroplane at high altitude in the cruise configuration is in a condition that the forces acting upon it are in a state of equilibrium. Because of the high TAS at altitude any disturbing force will cause a large deviation from the original state. The dynamic stability of the aeroplane at altitude is reduced and the damping of such deviations is diminished. The aeroplane is less stable and any control movements to recover the original altitude and/or attidue must be made slowly and smoothly. 40. Buffet is vibration of the airframe which, if excessive and prolonged, could cause structural damage, interfere with control of the aeroplane, and may result in excessive crew fatigue. Whilst in the cruise, buffet may be experienced for two reasons. At low speed it will be felt just prior to the aeroplane stalling, known as the pre-stall buffet, which occurs at approximately 1.2 VS. When approaching the maximum speed limit the high speed buffet will be experienced, due to shockinduced separation (see chapter on high speed flight). Except for low-speed stall warning buffet, there should be no perceptible buffet at any speed up to VMO/MMO (JAR 25.251(d)). 41. The speed range available at which to cruise is, therefore, between 1.2 VS and VMO/MMO, and this speed range will decrease as altitude is increased. When this speed range has reduced to zero the aeroplane is at its aerodynamic ceiling (not to be confused with the service ceiling, which is dicated by power available). Although a low-speed stall warning is fitted, it is unlikely that the pilot would allow the speed of the aeroplane to fall to this speed inadvertently. However, it is possible to inadvertently exceed VMO/MMO and encounter the high-speed buffet. This increase in speed may be

Chapter 10 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
caused by gust upsets, uninentional control movement, or when levelling off from a climb or descent from mach to airspeed limit values such as during an emergency descent. For such an eventuality a high-speed aural warning device is fitted to most aeroplanes and in addition a maximum speed needle on the airspeed indicator shows VMO up to the altitude at which VMO = MMO where the datum becomes MMO. The warning will sound at approximately 10kts above VMO or 0.01M above MMO. 42. For normal operations it is possible to construct a graph against altitude to produce an envelope within which it is safe to operate in the cruise which is limited by the buffet onset. Such a Buffet Onset Boundary Chart is shown at Figure 10-7. 43. The lower limit is defined by the pre-stall buffet and the upper limit by VMO/MMO. This should produce a sufficient range of speeds and load factors for normal operations. Inadvertent excursions beyond the boundaries of the buffet envelope will not result in unsafe conditions without prior adequate warning to the pilot (JAR 25.251(c)). 44. Although the maximum limit of the buffet boundary is fixed for all weights, the lower limit increases with increased weight. Therefore the range of speeds available for normal operations increases as the weight reduces with fuel burn. For any given weight, centre of gravity position and airspeed the maximum operating altitude is that at which it is possible to achieve a positive normal acceleration increment of 0.3g without exceeding the buffet onset boundary (ACJ 25.1585(c)). It is therefore common practice to draw the buffet onset chart for 1.3g. The intersection of the pre-stall

Chapter 10 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
buffet and VMO/MMO lines defines the aerodynamic ceiling (JAR 25.251(e)2). It follows then that as altitude increases manoeuvrability becomes increasingly restricted. In good weather this is of little consequence apart from encountering clear air turbulence but in bad weather it has great significance because of the possibility of turbulence disturbing thte status quo and requiring correction which becomes increasingly more difficult as altitude increases. Operating areas near the limits are often referred to as coffin corner.

Chapter 10 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
EXAMPLE 10-1

EXAMPLE
Using Figure 10-7. Given: Airspeed - M0.72 Altitude - 35,000ft Gross weight - 50,000 kg C of G - 10% MAC Find: Low and high speed buffet for 1g flight. Manoeuvre margin to initial buffet (load factor and bank angle).

SOLUTION
Enter graph with mach number (A) Move up to altitude line (B) Move across to reference line (C) Follow %MAC line to 10% (D) Move across to weight line (E) Move vertically down to find load factor and bank angle for initial buffet. = 1.47g and 47 degrees of bank Front point E move down the weight line to point F Move across to altitude line (G) via %MAC line Move down to speed scale to find 1g high speed buffet boundary: = M0.814 Move left to intercept altitude line again (H) to find 1g low speed buffet boundary: = M0.54 (178kts)

Chapter 10 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
FIGURE 10-7 Buffet Onset Boundary Chart

Chapter 10 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
EXAMPLE 10-2

EXAMPLE
Using Figure 10-8. Given: Load factor - 1.3g Gross weight - 59,000 kg C of G - 10% MAC Find: Aerodynamic ceiling.

SOLUTION
Enter graph with load factor (A) Move up to gross weight (B) Move across to %MAC and back to reference line (C) Move across to intercept peak of altitude line (D) = 34,000ft.

Chapter 10 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Limitations
FIGURE 10-8 Buffet Onset Boundary Chart

Chapter 10 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Special Circumstances
Effects of Ice on a Propeller Heavy Tropical Rain Loss of Controllability Effects on High Lift Devices Effect of Surface Damage, Modifications and Age Effect on the Lift/Drag ratio Windshear

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances

11

Special Circumstances

1. It has been shown in previous chapters that the performance of the boundary layer is fundamental to the lift coefficient of the wings and other lifting surfaces. Clearly anything which modifies the lifting surfaces in a manner detrimental to the boundary layer will reduce the lift coefficient.

Effects of Ice on a Propeller


2. The formation of ice on a propeller blade can have similar effects to those described in Chapter 2 with reference to the aircraft wing. 3. The aerodynamic shape of the blade will change and result in a reduction in its capability to convert the power from the engine into thrust. In addition to the reduction in thrust, there will also be an increase in drag due to the relatively rough nature of the ice causing an increase in skin friction. 4. The increase in weight and its distribution can lead to increased loading and out of balance vibration and should the ice adhering to the blade now come free, damage to the airframe and/or engine can ensue. 5. Furthermore, the pitch change mechanism of a variable pitch propeller can be affected by the formation of ice and hence a reduction in efficiency, which will already have been reduced by a reduction in the thrust produced.

Chapter 11 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances

Heavy Tropical Rain


6. During heavy rain there is a rapid accumulation of water which distorts the upper wing surface shape. This can cause total lift to reduce by up to 30%. The drag rise can be considerable, especially if the landing gear is down, and jet engines may be slower to respond to rapid demands for more power, or even flame out altogether. In addition the weight of the aeroplane may increase considerably. The sum effect of these factors will be to increase the stalling speed of the aeroplane. Under such conditions it is prudent to delay takeoff or landing. 7. Heavy rain may cause standing water on the runway to such a depth that take-off acceleration may be affected and engine ingestion of water from tyre spray may be a hazard. Standing water also increases the danger of aquaplaning by the wheel tyres, especially during landing.

Loss of Controllability
8. Any reduction of boundary layer energy reduces the aircraft speed at which the adverse pressure gradient will cause boundary layer separation and therefore increases the stalling speed and reduces the stalling angle of attack. This in turn means that use of the ailerons at low speeds or high angles of attack is more likely to lead to tip stall and loss of lateral control. Similarly, high rates of turn are more likely to lead to manoeuvring stall, because of the higher effective angle of attack necessary during the turn. Clearly, ice or frost contamination of the elevator and rudder surfaces increases the likelihood of their stalling when large control demands are made, as may be necessary during flight at low speed.

Chapter 11 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances
9. The control surfaces move about a fulcrum, which is not usually coincident with the centre of pressure of the control. Hence there is an aerodynamic moment, which must be balanced by the input force from the pilots controls. The changed aerodynamic characteristics due to ice or frost formation on the control surfaces may significantly alter the aerodynamic moment. This, in turn, means that the force required to move the surfaces would be altered.

Effects on High Lift Devices


10. Ice formation in the flap guide tracks may inhibit the deployment of the flaps. Deployment of the types of trailing edge flap which reduce the stalling angle of attack (plain flap and split flap), on a wing with serious ice contamination, could conceivably induce a stall during low speed flight on the landing approach. 11. Leading edge ice formation on the flaps themselves is particularly detrimental in the case of slotted flaps, since the airflow through the slots may be affected. The consequent loss of the reenergising effect on the boundary layer results in loss of lift and an increase in drag. The effect on the take-off will be to increase the length of the take-off run and on landing to increase the landing speed and landing distance required. 12. Surface frost or ice on leading or trailing edge flaps will have the same effect as on the wing surface, encouraging boundary layer separation and increasing drag. This will reduce the effectiveness of the high lift devices.

Chapter 11 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances

Effect of Surface Damage, Modifications and Age


13. Just as the surface roughness of ice and frost increases skin friction and destroys boundary layer energy, so will any other form of surface roughness have the same effect, to a greater or lesser degree. Flaking or scratched paint, abraded surfaces and surface indentations all increase surface friction and slow down the sub-layer and encourage separation. Repairs and modifications to the aircraft surface, unless carefully faired and smoothed, also modify the boundary layer flow and may lead to local separation. All these factors tend to increase as an aircraft grows older in service. Minor abrasions and local damage, whilst insignificant individually, become cumulatively significant, degrading the aerodynamic performance to an extent where it no longer matches the original manufacturers specifications. 14. Furthermore, as aircraft grow older they are often repainted in new liveries with only partial removal, if any, of the old. This can add considerably to the overall weight of the aircraft with a consequent reduction in performance from that advertised.

Effect on the Lift/Drag ratio


15. All surface contamination increases drag and any contamination on the lifting surfaces reduces lift. Consequently surface contamination, whether due to ice, frost, modification or wear and tear, inevitably reduces the lift/drag ratio.

Chapter 11 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances

Windshear
16. Windshear is caused by variations in the direction and/or speed of the local wind with changes in height and/or horizontal distance. It is almost always present but normally does not cause undue difficulty in controlling an aeroplane. It is abnormal windshear which is dangerous, and tends to displace an aeroplane abruptly from its intended flight profile such that substantial control input is required. A full explanation of the causes may be found in the Meteorology notes.

Effects of Windshear
Energy Loss
17. An aircraft encountering windshear tends to maintain its speed over the ground due to its own momentum (the larger the aircraft the more momentum it will have). If the windshear is due to a reduction in headwind component (or increase in tailwind component) this reduction manifests itself as an energy loss and a reduction in indicated airspeed. Lift is therefor reduced and the aircraft will, without correction, suffer a loss of height/increase in rate of descent/decrease in rate of climb. The situation is shown at Figure 11-1.

Chapter 11 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances
FIGURE 11-1 Effect of Reduction in Headwind Component

Energy Gain
18. An increase in headwind component (or decrease in tailwind component) results in an energy gain and increase in indicated airspeed, as shown at Figure 11-2.

Chapter 11 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances
FIGURE 11-2 Effect of Increase in Headwind Component

19. These events become critical when the aircraft is being flown close to the ground during the final stages of an approach or shortly after take-off. In the energy loss case the engine reaction time when additional thrust is demanded can be critical. 20. The energy gain/loss situations described above can occur as a result of either vertical windshear or as horizontal windshear, in other words the aircraft can either climb/descend or fly horizontally into air flowing at a different speed or from a different direction, in either event changing the head/tail wind component. In simple terms a change in the wind component will, in the short term, change the airspeed rather than the groundspeed of the aircraft.

Downdraught
21. Figure 11-3 shows an aircraft taking off in the vicinity of a thunderstorm.

Chapter 11 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances
FIGURE 11-3 Effect of Downdraught on Take-off

22. The situation illustrated is the critical case where the headwind component decreases sharply and/or becomes a tailwind component shortly after take-off (energy loss). In this case, because of inertia, the groundspeed remains constant but the airspeed decreases sharply. The loss of lift associated with the resulting low airspeed may cause the aircraft to strike the ground.

Chapter 11 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances Approach Under Thunderstorm


23. At Figure 11-4 an aircraft is approaching to land in the vicinity of a thunderstorm.

FIGURE 11-4 Effect of Thunderstorm on the Approach to Land

Chapter 11 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances
24. Initially, at position A, the aircraft is stabilised on the glideslope maintaining target airspeed. Suddenly the aircraft flies through a gust front into the cold air which is spilling from the base of the cloud. What was a slight tailwind component is now a very marked headwind component but, because of inertia, the groundspeed will momentarily remain constant. The result is that the airspeed is increased by an amount equal to the change in wind component. The amount of lift generated will increase with the increased airspeed, and the aircraft will initially make a rapid excursion above the desired glideslope at point B. The natural reaction of the pilot in this situation is to reduce thrust and steepen the approach. Now, the increased headwind component will give a lower groundspeed once the effect of inertia has been overcome, and this will require an increase in thrust and a decreased rate of descent.In other words the pilots initial response is 100% incorrect, and the aircraft is now descending rapidly in an underpowered configuration to point C and a go-around must be initiated.

Actions to be Taken on Encountering Windshear


25. If windshear is likely to be encountered then one of the easiest measures to be taken is to increase airspeed on the approach, bearing in mind that this could lead to stopping problems if the runway is short. If loss of airspeed occurs unexpectedly then the recommended actions are: (a) (b) (c) (d) Rapidly increase power (full go-around power if necessary). Raise the nose to check descent. Coordinate power and pitch. Be prepared to carry out a missed approach.

26. If downbursts are encountered near thunderstorms, then the recommended actions during approach or take-off phases are:

Chapter 11 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Special Circumstances
(a) (b) (c) Use the maximum power available as soon as possible. Adopt a pitch angle of about 15 degrees and hold that attitude. Do not chase airspeed. Be guided by stick-shaker indications when holding or increasing pitch attitude, easing the back pressure as required to attain and hold a slightly lower attitude.

27. It is difficult to be specific on the use of autopilots, autothrottles and flight directors, but in a severe encounter it may be best to revert to manual handling.

Chapter 11 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

081 Principles of Flight

Propellers
Propeller principles Power Absorption Fixed Pitch Propellers Variable Pitch Propellers Propeller Effects Propeller Forces and Moments

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

12

Propellers

1. The aircraft propeller is a device for converting the shaft power of a piston engine or gas turbine into thrust power. It functions by imparting a relatively small velocity increase to a relatively large mass of air.

Propeller principles
2. The aircraft propeller comprises a number of blades mounted upon a central shaft, rotating about an axis that is parallel to the direction of required thrust, i.e. the direction of flight. The rotating propeller blades form a disc that is perpendicular to the direction of thrust. Figure 12-1 shows the propeller disc of a four-bladed propeller.

Chapter 12 Page 1

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-1 Propeller Disc

Blade Angles
3. Each propeller blade has an aerofoil section similar to that of a wing as shown at Figure 12-2.

Chapter 12 Page 2

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-2 Propeller Blade Terminology

4. Because the blade section of a propeller is an aerofoil section similar to that of a wing, it generates lift as it rotates in rather the same way as a wing creates lift. 5. When the aircraft is stationary in still air conditions, with the propeller rotating, the airflow relative to the propeller blades is parallel to the plane of rotation. Under these circumstances the propeller blade angle or geometric pitch angle, which is the angle between the plane of rotation and the chord line of the aerofoil and the angle of attack, are the same thing. When the aircraft is moving forward however, there is a forward velocity vector in addition to the rotational velocity and their resultant is the path of the blade element which of course determines the airflow relative to the propeller blade. 6. The angle between the plane of rotation and the relative airflow is referred to as the helix angle, or angle of advance, and thus: Blade angle = (Geometric pitch angle) Helix angle + (Angle of advance) Angle of attack ( )

This is illustrated at Figure 12-3.

Chapter 12 Page 3

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-3 Angle of Attack

Chapter 12 Page 4

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
7. The geometric pitch is the distance that a propeller should advance in one complete revolution. A useful analogy is the distance that a wood screw would advance into a piece of timber in one complete revolution. The effective pitch is the distance that the propeller actually advances in one revolution. The difference between the two is known as propeller slip. The principle is illustrated at Figure 12-4. Because air is not a perfect propulsion medium propeller slip is inevitable and it accounts for a significant energy loss in the conversion of torque to thrust.

FIGURE 12-4 Propeller Slip

Chapter 12 Page 5

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

Forces on a Propeller Blade


8. When set at a small positive angle of attack, the blade produces a total aerodynamic reaction forces whose components perpendicular and parallel to the relative airflow are, of course, lift and drag respectively. However, if the total reaction is resolved relative to the flight path of the aircraft, the parallel component is the thrust force and that which is perpendicular is the propeller torque force, as shown at Figure 12-5.

Chapter 12 Page 6

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-5 Forces on a Propeller Blade Element (Section)

Chapter 12 Page 7

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
9. Note from Figure 12-5 that the better the lift/drag ratio the further forward is the total reaction vector, and consequently the greater the amount of engine power that is translated into thrust, whereas propeller torque acts in the plane of rotation and opposes engine torque, ie. it is the resistance to rotary motion and as such is an undesirable force vector since it is trying to rotate the aircraft in the opposite direction to that of the propeller rotation.

Blade Twist
10. It can be seem from Figure 12-3 that the angle of attack is governed by the relationship between the propeller's rotational velocity and the aircraft's forward speed. For the best lift/drag ratio, the most desirable propeller angle of attack lies between 2 and 4. At angles of attack greater than 15, the blades will stall and the propeller will become ineffective. 11. Clearly, the rotational velocity of the propeller blade is greater at the tip than at the root and consequently, the rotational velocity depends upon the point on the blade radius where the rotational velocity is considered. Since the rotational velocity is one of the vectors that determine the direction of the relative airflow, and therefore the blade angle of attack, the blade is twisted to maintain a constant angle of attack along its entire length. Blade twist is illustrated at Figure 12-6.

Chapter 12 Page 8

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-6 Blade Twist

Propeller Efficiency
12. The mechanical efficiency of any system is the ratio of power out to power in. 13. The force rotating the propeller is the torque, or twisting force developed by the engine and delivered to the propeller shaft. The power delivered to the propeller is the product of this twisting force, or torque, and the angular velocity of the propeller shaft. Engine Power Delivered = Torque x 2N

Chapter 12 Page 9

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
where N = angular velocity 14. The force exerted by the propeller is the product of the mass of air moved and the acceleration given it by the propeller. This force is called thrust. The power developed by the propeller is the product of thrust and the true airspeed (TAS) of the aircraft. Propeller Power Developed = Thrust x TAS Therefore: Thrust x TAS Propeller Efficiency ( ) = -------------------------------Torque x 2N

Effect of Airspeed on Propeller Efficiency


15. If the blade angle, or pitch, of a propeller is fixed its angle of attack will vary with varying aircraft forward speed. The greater the forward speed of the aircraft the less the angle of attack and, of course as angle of attack decreases so does thrust. Figure 12-7 shows the situation with a fixedpitch propeller blade rotating at constant rpm, but at different forward speeds.

Chapter 12 Page 10

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-7 Effect of Speed on a Fixed Pitch Propeller

16. At low forward speed the angle of attack is high and, although the blade will be producing a large amount of thrust there will also be high drag. The forward speed (TAS) is low, and the torque required to maintain propeller rpm is high in order to overcome the high drag. Since: Thrust x TAS Propeller Efficiency ( ) = -------------------------------Torque x 2N 17. It can be seen that propeller efficiency will be low under these circumstances.

18. At high forward speed the angle of attack is reducing to the incidence at which no lift is generated, which means that the thrust element of the efficiency equation is low and therefore propeller efficiency will be low.

Chapter 12 Page 11

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

Power Absorption
19. The amount of engine power that the propeller is able to absorb and convert into thrust depends upon the lift-producing capability of the blades. As with a wing, the greater the aspect ratio the greater the lift generated and aspect ratio is the ratio of span to mean chord. In the case of a propeller blade it is the ratio of blade length to mean blade chord, so increasing propeller diameter increases aspect ratio. 20. However, the greater the propeller diameter the higher the tip speed for any given rpm, so that the tips reach sonic velocity at relatively low rpm. The compressibility effects at sonic velocity have a very detrimental effect upon propeller efficiency, since the thrust is considerably reduced and the blade drag (torque required) increased. Hence there is an aerodynamic limit on the practical propeller diameter and upon the TAS at which a large diameter propeller can be used. 21. Another reason for avoiding sonic velocities at the blade tips is the unacceptably high noise levels produced which, apart from environmental considerations, can have damaging fatigue effects upon the aircraft structure.

Propeller Solidity
22. The usual way in which the power absorption ability of the propeller is increased is to increase what is called the solidity of the propeller disc. The solidity at any point on the radius of the propeller disc is the ratio of total blade chord at that radius to total circumference at that radius. The radius normally used for the calculation is 70% tip radius. The concept of solidity is shown at Figure 12-8.

Chapter 12 Page 12

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-8 Propeller Solidity

Chapter 12 Page 13

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
Number of Blades x Chord at 70% Tip Radius Solidity = -------------------------------------------------------------------------------------------------------------70% Tip Radius Circumference 23. From the above it can be seen that the solidity of the propeller disc can be increased either by increasing the number of blades or by increasing the blade chord. 24. There is a practical limit to the number of blades that can be used, partly due to aerodynamic interference between blades and partly due to the complexity of the pitch control mechanism, which must be contained within the propeller hub. However, developments in both blade and pitch control technology in recent years have made six and more blades possible when, up until the early 1980s, five blades was generally considered to be the limit. 25. Increasing the blade chord, to produce the paddle blade seen on most C-130 Hercules transport aircraft, has the disadvantage of low aspect ratio but was considered the best choice at the time for use with a high rpm power plant.

Fixed Pitch Propellers


26. There is clearly a very limited range of flight speeds at which the propeller blade angle of attack will be close to the optimum and at which propeller efficiency is maximum. The blade angle of a fixed pitch propeller must therefore be chosen to suit the performance requirements of the aircraft to which it is to be fitted. 27. If a small blade angle, or fine pitch, is used the propeller will be at its most efficient at low forward speeds. Thus it will be ideal for take-off and low flight speeds, but incapable of operating efficiently at high TAS.

Chapter 12 Page 14

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
28. Conversely, if a large blade angle, or coarse pitch, is chosen the propeller will be inefficient and will demand excessively high torque during take-off, but will be suitable for high flight speeds. 29. Examination of Figure 12-3 will show that the blade angle of attack could be adjusted by varying the rotational velocity (propeller rpm), but this is a very limited option, especially with a piston engine powered aircraft. By far the better solution is to operate at constant rpm and alter the blade angle to produce an angle of attack to suit the forward speed of the aircraft; fine pitch for takeoff and landing and progressively coarser pitch for progressively higher TAS. This is illustrated at Figure 12-9.

Chapter 12 Page 15

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-9 Effect of Speed on Angle of Attack

Chapter 12 Page 16

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

Variable Pitch Propellers


30. Historically this led to the development of the two-pitch propeller, offering fine pitch for takeoff and climb and coarse pitch for cruise and high-speed flight. Subsequently the fully variable pitch constant-speed propeller was introduced, enabling optimum blade angle of attack to be maintained over the entire range of flight speeds, with the engine/propeller rpm governed at a constant value. 31. This is shown at Figure 12-10 which compares the efficiency of a two-pitch propeller and a variable pitch propeller with varying airspeeds.

FIGURE 12-10 Propeller Efficiency

Chapter 12 Page 17

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
32. Reference to the propeller efficiency equation will show that efficiency increases with increasing TAS. However there is a limit to this because there is clearly a finite limit to the extent to which the blade angle can be increased and still produce a forward component of thrust. At about 300-350 knots TAS the blade angle is at its greatest practical value and any further increase in airspeed will result in a reduction of thrust, so propeller efficiency begins to fall. This is illustrated at Figure 12-11, which compares the efficiency of various forms of propulsion against airspeed.

Chapter 12 Page 18

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-11 Propulsive Efficiency Comparison

Chapter 12 Page 19

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
33. Deformation of the aerofoil section of the propeller blade will alter the thrust developed at any airspeed. Whilst propeller blades are susceptible to deformation due to stones, bird strike and other foreign object damage by far the greatest risk of deformation is from propeller icing. The accretion of ice on the blades will significantly change their aerodynamic characteristics and drastically reduce the thrust produced and propeller efficiency at any given airspeed and propeller rpm. Most propeller-powered aircraft are equipped with some form of propeller anti-icing or deicing.

Propeller Effects
Torque Effect
34. Newtons third law tells us that to every action there is an equal and opposite reaction. The force rotating the propeller is torque so it follows that there must be a reaction force in the opposite direction of rotation. In a single engine aircraft this torque reaction is attempting to rotate the aircraft about the propeller shaft axis. This can create a rolling (and secondary yawing) tendency, especially at take-off. 35. Torque effect can be eliminated by mounting two propellers, one behind the other, each driven by its own shaft but in opposite directions. The shafts are concentric and as the propellers rotate in opposite directions, their torque effects cancel each other out. Such an arrangement is referred to as counter-rotating propellers. 36. Contra-rotating propellers, on the other hand, are two propellers mounted in tandem but driven by a single engine through a gearbox to ensure the propellers rotate in opposite directions and likewise their torque effects are cancelled out.

Chapter 12 Page 20

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

Slipstream Effect
37. An asymmetric slipstream effect occurs with single engine aircraft, and is illustrated at Figure 12-12. The slipstream corkscrews around the fuselage and strikes the fin predominantly on one side, creating a yawing moment. The propeller shown at Figure 12-12 is a right hand tractor propeller (it rotates in a clockwise direction when viewed from behind), and the yaw produced would be to the left. This can be compensated for by mounting the fin at a slight angle to the aircraft centre line, or by the application of right rudder, especially at high power settings.

FIGURE 12-12 Slipstream Effect

Chapter 12 Page 21

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

Asymmetric Blade Effect


38. Asymmetric blade effect occurs because the axis of rotation (the thrust axis) is inclined in relation to the horizontal path of the aircraft. It is therefore most noticeable with tail wheel aircraft on take-off and landing, or with any aircraft when a high angle of attack is employed. 39. The essence of asymmetric blade effect is that the down going blade has a greater angle of attack than the up going blade, and therefore generates more lift (thrust) than the up going blade. This is shown at Figure 12-18.

Chapter 12 Page 22

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-13 Asymmetric Blade Effect

Chapter 12 Page 23

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
40. Asymmetric blade effect is further exacerbated by the fact that the down going blade travels further (in relation to the direction of travel of the entire aircraft) than the up going blade. Since it does so in the same time, its relative speed is higher and thus its aerodynamic force higher. Figure 12-14 illustrates this phenomenon.

FIGURE 12-14 Asymmetric Blade Effect

Chapter 12 Page 24

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

Gyroscope Effect
41. Since the rotating propeller displays certain of the characteristics of a gyroscope, whenever a force is effectively applied to the propeller, a precessed force will result which is transposed through 90 in the direction of rotation of the propeller. 42. If a tail wheel aircraft with a right hand tractor propeller is accelerating for take-off, the act of raising the tail will be as if a force was applied downwards onto the propeller hub. This applied force will be precessed 90 clockwise (when viewed from behind) and will cause a yaw to the left. Conversely, when either a nose- or tail-wheeled aircraft with a right hand tractor propeller rotates into the climbing attitude, gyroscopic effect will cause a yaw to the right. Similarly, yaw will be precessed to produce a pitching moment.

Propeller Forces and Moments


43. A rotating propeller is acted upon by centrifugal, twising and bending tensile forces. Centrifugal force tends to stretch the blades away from the hub, air resistance to rotation applies a torsional force that tends to bend the blades opposite to direction of rotation and the thrust load on the blades tend to bend them forwards, as shown at Figure 12-15.

Chapter 12 Page 25

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-15 Forces Acting on a Rotating Propeller

Chapter 12 Page 26

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
44. In addition, variable pitch propellers have further forces acting upon them which create twisting moments about their pitch change axis, thus altering the blade angle.

Aerodynamic Twisting Moment (ATM)


45. The total aerodynamic reaction force acting about the pitch change axis (pivot point) creates an aerodynamic twising moment which tends to increase (coarsen) the blade angle, as shown at Figure 12-16.

Chapter 12 Page 27

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-16 Aerodynamic Twisting Moment

Centrifugal Twisting Moment (CTM)


46. The centrifugal force acting on a propeller is shown at Figure 12-17 resoloved into its radial and tangential components considered at two points, A at the leading edge and B, the trailing edge of the blade.

Chapter 12 Page 28

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
47. The radial components tend to stretch the blade whereas the tangential components produce a twising moment about the pitch change axis. This centrifugal twisting moment tends to reduce the blade angle (fine of the blade) as shown at Figure 12-7, and it is always greater than ATM.

Chapter 12 Page 29

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-17 Centrifugal Twisting Moment

Chapter 12 Page 30

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

Engine Failure
48. In the event of engine failure, or the need to shut down an engine, the torque driving the propeller is clearly lost. In the case of the constant speed (variable pitch) propeller, the propeller speed governor will drive the blades to fine pitch in an attempt to maintain rpm. The aircraft still has forward speed however, either due to the glide or the thrust from its remaining propellers. The airflow relative to the blades of the unpowered propeller will now be at a negative angle of attack, producing torque to rotate the propeller in the normal direction of rotation. The force component perpendicular to the plane of rotation now acts rearwards and is therefore no longer thrust, but drag. This is illustrated at Figure 12-18.

Chapter 12 Page 31

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-18 Forces on a Windmilling Propeller

Windmilling
49. The rotating blades at fine pitch present almost a flat disc at right angles to the flight direction and therefore very high drag forces are present. Because the propeller is behaving exactly like a windmill in the wind, this is known as windmilling drag.

Chapter 12 Page 32

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
50. In a twin-engine aircraft a windmilling propeller leads to asymmetric forces which can render the aeroplane uncontrollable. In order to maintain flight speed above stalling the pilot is obliged to increase power on the good engine. The increased thrust from this, coupled with the drag of the windmilling propeller, can produce a yawing moment too great for the rudder to compensate, especially at low airspeed. 51. Additionally, the windmilling propeller driving a damaged engine may lead to overheating and seizure. 52. Note also, that the total reaction force now produces a twisting moment about the pitch change axis which tends to reduce the blade angle ie. both CTM and ATM will fine off the blade, as shown at Figure 12-19.

Chapter 12 Page 33

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-19 Twisting Moments on a Windmilling Propeller

Chapter 12 Page 34

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers

Propeller Feathering
53. To prevent the propeller from windmilling in the event of loss of engine torque the pitch control mechanism usually incorporates a feathering facility. When the propeller is feathered the blade angle is increased until the angle of attack is 0 and the blade angle is approaching 90. With the blades thus edge on to the airflow there is no lift force to produce a torque reaction and the propeller is not rotated. In practice, because the aerofoil section of a propeller blade is cambered, the zero lift, or feathering blade angle is typically about 85=as shown at Figure 12-20.

FIGURE 12-20 Feathered Propeller Blade

Chapter 12 Page 35

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
54. With the propeller blades feathered the propeller is no longer presenting a disc to the airflow and the drag is very significantly reduced when compared to that of a windmilling propeller. This has a very beneficial effect on the glide performance of the aircraft since the greater the drag the steeper the glide angle necessary to maintain airspeed. 55. Similarly, the relatively low drag of a feathered propeller reduces the yawing moment, even with high power on the remaining engine, leaving sufficient rudder authority to control the aircraft in yaw.

Reverse Thrust
56. In order to provide aerodynamic braking from the propeller during the landing roll, the blade is turned until a negative blade angle is achieved which, when power is applied, will produce thrust acting in the opposite direction to that normally produced by the propeller. This is shown at Figure 12-21.

Chapter 12 Page 36

G LONGHURST 1999 All Rights Reserved Worldwide

Propellers
FIGURE 12-21 Reverse Thrust

57. The blade angle in reverse is restricted by the blade twist effect now also becoming reversed. The blade tip is at a greater angle than the root when reverse pitch is selected and thus the thrust is greatest at the tip. This creates a significant bending moment and limits the amount of reverse thrust which the propeller can tolerate. Typically, in reverse the blade angle is limited to about -25.

Chapter 12 Page 37

G LONGHURST 1999 All Rights Reserved Worldwide

Você também pode gostar