Вы находитесь на странице: 1из 491

FOREWORD

1. The field of military aviation is expanding at a rapid pace,


therefore it is essential that the wealth of knowledge is continuously
updated, amongst our operators.

2. In Indian Air Force, a young trainee pilot is groomed from the


nascent stage, either at the National Defence Academy (NDA) or the
Air Force Academy (AFA), well before taking the first flight. In order to
achieve the desired quality of training amongst pilots, there is a need
to have a single, authentic and comprehensive, source of knowledge.

3. After extensive efforts, the Flying Instructors School (FIS) has


prepared four volumes of quality training notes. These books are
aimed to cater to the knowledge requirements, of not only the ab-
initio, but also to those of an instructor and beyond.

4. In order to cater to the future advancements in the field of


aviation there would be a requirement to update this knowledge. I am
confident that these books would ensure standardization in training
and enrichment of knowledge at all levels. I compliment FIS for an
excellent contribution in this regard.

-sd-
(S Bhojwani)
Air Marshal
Date : Nov 05 AOC-in-C
FIS BOOK 1

AERODYNAMICS

INDEX

Chapter Title Page No.


1. Mechanics 1
2. The Atmosphere and Airspeed 17
3. Theories of Development of Aerodynamic Forces 31
4. Viscosity and Boundary Layer 57
5. Aerofoils 69
6. Lift 83
7. Drag 91
8. Stalling 113
9. Spinning 129
10. Wing Planform 147
11. Lift Augmentation 177
12. Flight Controls 191
13. Trimming and Balance Tabs 209
14. Stability 215
15. Level Flight 245
16. Climbing and Gliding 249
17. Manoeuvres 261
17 (a). Operating Strength Limitations 279
18. Range and Endurance 291
19. Transonic and Supersonic Aerodynamics 325
20. Design for High Speed Flight 363
21. Propeller Theory 391
22. Application of Aerodynamics to Specific Problems of Flight 413
23. Basics of Rotor dynamics 445
24. Hovering and Horizontal Movement 451
25. Control in Forward Flight 461
26. Power Requirements 477
27 Autorotation 483
CHAPTER 1

MECHANICS

Scalars and Vectors

1. Scalar. A quantity that has only magnitude and no direction is called a scalar quantity, e.g.
length, area, volume, mass, density, speed, work, energy etc.

2. Vector. A quantity that has both magnitude and direction is called a vector quantity, e.g.
displacement, velocity, acceleration, force, momentum etc.

3. A vector quantity is usually represented by a line


with arrowhead. The magnitude of the vector is shown by
the length of the line and the direction of the arrow
represents the direction. To find the resultant of scalar
quantities, they are simply added algebraically. To find
the resultant of two or more vectors, the triangle law,
parallelogram law or polygon law is applied. A single Fig 1-1: Triangle Law of Vectors

vector may also be resolved into two or more


components.

4. Triangle Law of Vectors. If two vectors (a &


b) are represented in magnitude and direction by two
sides of a triangle (BC & CA) taken in order, the resultant
is represented by vector c in magnitude and direction by
the third side of the triangle (BA) taken in the opposite Fig 1-2: Parallelogram Law of
Vectors
order (Fig 1-1).

5. Parallelogram Law of Vectors. If two vectors


are represented in magnitude and direction by the two
adjacent sides of a parallelogram (AB & AD) drawn from
a point, the resultant vector r is represented in magnitude
and direction by the diagonal (AC) passing through that
point (Fig 1-2).
Fig 1-3: Polygon Law of Vectors

6. Polygon Law of Vectors. If the vectors are represented by the sides of a polygon, taken in
order, the resultant r is represented by the closing side of a polygon (AF) taken in opposite order (Fig
1-3).
FIS Book 1: Aerodynamics 2

Basic Definitions

7. Speed. The rate at which a body is displaced in position is called its speed. Speed is a
scalar quantity.

8. Velocity. The rate of displacement of a body in a given direction is called its velocity. It is a
vector quantity. Unit of both speed and velocity is m/s.

9. Acceleration. It is the rate of change of velocity. The change may be in speed or direction
2
or both. Unit: m/s .

10. Mass. The mass of a body is the quantity of matter contained in it. Unit: kilogram (kg).

11. Inertia. There is a natural tendency for things to continue doing what they are already
doing. A body that is at rest tends to remain at rest. A body that is moving tends to continue moving at
the same speed in the same direction. This tendency of a body in equilibrium, to continue in the same
state is due to its inertia. It is a property possessed by all bodies, which shows the reluctance to
change their state of equilibrium. It is a quality and has no units. Inertia can be measured only in
terms of mass, which is a scalar quantity.

12. Force. Application of force changes or tends to change the state of rest or uniform motion
of a body in a straight line. In other words, force is that which changes or tends to change the
momentum of a body. F = ma. Unit: Newton.

13. Momentum. The momentum of a body is the quantity of motion contained in a body and is
the product of the mass and velocity of the body. Momentum is a vector quantity having the same
direction as the velocity vector, and may be resolved into components, or combined with another
momentum to give a resultant, like other vector quantities. Momentum should not be confused with
inertia which is merely a quality related to the mass of the body, not its velocity. Momentum = mv.
Unit: kg.m/s.

14. Laws of Motion. Sir Issac Newton propounded three laws of motion:

(a) First Law. A body will continue in its state of rest or uniform motion in a straight line
unless compelled to change that state by an external force.

(b) Second Law. The rate of change of momentum is proportional to the applied force,
and the change of momentum takes place in the direction of the applied force.

(c) Third Law. Action and reaction are equal and opposite. (Note that action and
reaction refer in this case to different bodies).
3 Mechanics

15. Newton’s second law is equivalent to defining force as that which causes acceleration. From
the second law we can say that F ∝ rate of change of momentum. But momentum = mv, and if m is
constant for a given body, F ∝ m X rate of change of v, or F ∝ ma. If the unit of force is chosen at that
force which gives unit acceleration to unit mass, the equation can be written as F = ma.

16. Impulse of a Force. Suppose a force F acts on a body of mass m for a short time t (e.g.
when a bat strikes a ball), such that the velocity of the body changes from u to v.

As F = ma,

The acceleration ‘a’ produced by the force will be F/m.

Substituting ‘a’ in v = u + at.


v = u + _F_ δt,
m
or F . δt = mv – mu.

17. The expression on the left (F.δt) is known as the impulse of the force, and is equal to the
change of momentum caused by the force.

18. Conservation of Momentum. Consider two bodies travelling in the same direction which
collide with each other, the duration of the collision being a short time say δt. Throughout the collision
each will experience a force equal and opposite of that experienced by the other (Newton’s third law).
The impulse of the force is F δ t and is the same for each body. Thus the change of momentum will be
the same for each body. If at the time of collision, body A was overtaking body B, it is apparent that
the effect of the impact will be to decrease the momentum of A and increase that of B, and the total
momentum of the system of two bodies will remain unchanged.

19. The Law of Conservation of Momentum. The effects of the interaction of parts of a closed
system are summarized in the law of conservation of momentum, which states that the total
momentum in any given direction before impact is equal to the total momentum in that direction after
impact.

20. Work. A force is said to do work on a body when it moves the body along its line of action.
The amount of work done is equal to the product of the force and the distance moved (s) in the
direction of the force. W = Fs. Unit: Nm or Joules.

21. Energy. The energy of a body is its capacity to do work. E = Fs. Unit: Joules.

22. Power. The power of an engine is its rate of doing work.


FIS Book 1: Aerodynamics 4

P = Fs = Fv Where: F = Force
T s = Distance the body has moved
T = Time
Unit: Watts or Joules / second.

23. Newton’s Universal Law of Gravitation. It states that any two bodies in the universe
attract each other with a force directly proportional to the product of their masses and inversely
proportional to the square of the distance between them.

F ∝ M1 X M2 (1.1)
s2

24. Acceleration Due to Gravity ‘g’. It can be seen from the universal law of gravitation that
any body of some mass will be attracted towards any other body. If ‘M1’ be the mass of the earth and
‘M2’ be the mass of a body then it will be attracted towards the earth (assuming this is the only body
attracting it) with a force which is given by equation (1.1). Here ‘s’ is the distance between the given
body and the centre of the earth. This implies that different bodies at the same distance from the
centre of the earth, having different masses will have correspondingly different forces of gravity acting
upon them. However all of them will experience the same acceleration due to gravity, provided that
only the force of gravity acts upon them. This might sound a little complex however the following proof
shall make the understanding easier.

As per Newton’s law of gravitation we know that:


F ∝ M1 X M2
s2
Also if a body of mass ‘M2’ is being accelerated then the force acting on that body is:
F = M2 X a
Where in this case ‘a’ is acceleration due to earth’s gravity. Hence ‘a’ may be replaced by ‘g’
which denotes acceleration due to gravity. Therefore, we have:
M2 X g = M1 X M2
s2

Since M2 gets eliminated on both the sides of the equation, in all the cases where ‘S’ remains same
the acceleration due to gravity ‘g’ shall remain same as the mass of the earth is constant.

25. Weight. The force with which a body is attracted towards the centre of the earth is called its
weight. It is equal to the product of mass and acceleration due to gravity at that point. W = mg, where
g is the acceleration due to gravity at that point. Since g is not constant over the earth’s surface (more
at poles and lesser at equator), therefore the weight of a body is variable, unlike its mass which
remains constant.
5 Mechanics

26. A spring balance measures weight. A see-saw type of balance or a steel yard or a
weighbridge measures mass as in these cases force of gravity acts on both the sides and gets
eliminated.

27. Density. Density is the mass per unit volume of a substance. It is denoted by the symbol ρ.
Unit: kg/m³.

28. Pressure. Pressure is the force exerted per unit area. Unit: N/m2 or Pascals.

29. Stress. Stress is the force exerted between two contacting bodies or parts of a body. It is
measured as the force per unit area. Unit: N/m2.

30. Strain. Strain is the deformation caused by stress. It is recorded as the change of size over
the original size. Since it is a ratio, it does not have any unit.

31. Motion on Curved Path.

(a) Refer Fig 1-4. Consider a body


travelling along a circle of radius ‘r’ with
velocity ‘v’.

(b) In a small time say ‘t’ the body


travels from point A to point B describing an
arc S which subtends an angle ‘θ’ at the
centre O.

(c) Considering the radian theory, the


relationship between the arc, angle
subtended and the radius is:- Fig 1-4: Motion Along a Curved Path

s=rθ
Or θ = s / r

(d) By definition we know that the rate of change of angle per unit time is angular velocity
(ω), therefore:-

ω = θ/t
ω = s X 1
r t
ω = s X 1
t r
FIS Book 1: Aerodynamics 6

Now s / t is the rate of change of displacement i.e. the velocity ‘v’. Therefore, we have

ω = vX1
r
ω = v (1.2)
r

(e) Now ‘v’ is the linear velocity of the body. If θ is vary small, it can be considered, that
the two vectors (AD and BC) representing motion at these two points (Pt A and Pt B) are of
the same length. Though the magnitude has remained the same, the direction of the velocity
has changed. Therefore, by definition, the body is subjected to acceleration. Vector DC
represents the change in velocity between Pt A and Pt B in time ‘t’ and is represented by ‘∆v’.
Therefore, the acceleration is given by DC/t. i.e. ∆ v / t, and this is the centripetal acceleration
(CPA) ‘a’.

(f) From the figure, using radian theory, it can be said that:

DC = ∆ v = v X θ (1.3)
CPA = a = ∆ v / t (1.4)

Substituting the value of ∆ v from equation (1.3) into equation (1.4), we have:

a = vXθ/t
a = vXω (ω = θ / t)
a = v X v /r (ω = v / r) (From equation 1.2)
Therefore, a = v² (1.5)
r

32. Centripetal Force. It was shown that a body travelling with uniform speed in a circle has an
acceleration of v²/r towards the centre of the circle. The force producing this acceleration is termed as
the centripetal force, and for a body of mass ‘m’ the centripetal force is mv²/r towards the centre of the
circle.

As force (F) = ma
But in circular motion a = v2 therefore,
r
Centripetal Force (CPF) = mv² (1.6)
r
But m = w / g
where w = weight and g = acceleration due to gravity
Therefore, CPF = Wv2 (1.7)
rg
7 Mechanics

33. Centrifugal Force. It should be noted that in the case of a body travelling in a circular path
at the end of a string, while the mass is experiencing centripetal force towards the hand, there is an
equal and opposite reaction on the hand holding the string, known as centrifugal force. It is the
outward force on the body travelling along a curved path and represents the resistance of the mass to
centripetal acceleration. Centrifugal Force is an inertia force and it exists only as an equal and
opposite reaction to centripetal force.

34. Let us take the simple example of a stone on the end of a piece of string. If the stone is
whirled round so as to make one revolution per second, and the length of the string is 1 meter, the
distance travelled by the stone per second will be 2πr, i.e., 2π X 1 or 6.28 m.

Therefore v = 6.28 m/s, r = 1 m.


Therefore, acceleration towards centre = v²/r
= (6.28 X 6.28) / 1
= 39.5 m/s² (approx)

35. Notice that this is nearly four times the acceleration due to gravity, or nearly 4g. Since we are
only using this example as an illustration of principles, let us simplify matters by assuming that the
answer is 4g i.e. 39.24 m/s².

36. This means that the velocity of the stone towards the centre is changing at a rate 4 times as
great as that of a falling body. Yet it never gets any nearer to the centre! No, but what would have
happened to the stone if it had not been attached to the string? It would have obeyed the tendency to
go straight on, and in doing so would have departed farther and farther from the centre. The
acceleration of 4g may, in a sense be taken as the rate at which it is being prevented from doing this.
Now lets see what centripetal force will be required to produce this acceleration of 4g?

CPF = The mass of the stone X 4g.


So, if the mass is ½ kg, the centripetal force will be ½ X 4g = ½ X 4 X 9.81 = 19.62, say 20
Newtons.
Therefore the pull in the string is 20 N in order to give the mass of ½ kg, an acceleration of
4g.
Notice that the force is 20 Newtons, very roughly 2X9.81 m/s² or 2kgf, the acceleration is 4g.
There is a tendency to talk about "g" as if it were a force, it is not, it is acceleration.

37. Now this is all very easy provided the centripetal force is the only force acting upon the mass
of the stone - but is it? Unfortunately, no. There must, in the first place, be a force of gravity acting
upon it.
FIS Book 1: Aerodynamics 8

38. If the stone is rotating in a horizontal circle its


weight will act at right angles to the pull in the string,
and so will not affect the centripetal force. But of
course a stone cannot rotate in a horizontal circle, with
the string also horizontal, unless there is something to
support it. So let us imagine the mass to be on a table,
but it will have to be a smooth, frictionless table, or we
shall introduce yet more forces. We now have, at least
in imagination, the simple state of affairs as illustrated
in Fig 1-5.
Fig 1-5: Stone Rotating in Horizontal
Circle Supported on a Table
39. Now suppose we rotate the stone in a vertical
circle, like an aeroplane looping the loop, the situation
is rather different (Fig 1-6). Even if the stone were not
rotating, but just hanging on the end of the string, there
would be a tension in the string, due to it’s weight, and
this as near as matters would be ½ kgf, or very roughly
5 Newtons, for a mass of ½ kg. If it must rotate with an
acceleration of 4 g, the string must also provide a
centripetal force of 20 Newtons. So when the stone is
at the bottom of the circle D, the total pull in the string
will be 25 Newtons. When the stone is in the top
position, C, it’s own weight will act towards the center
Fig 1-6: Stone Rotating in a Vertical
and this will provide 5 N, so the string need only pull Circle
with an additional 15 N to produce the total of 20 N for
the acceleration of 4g. At the side positions, A and B, the weight of the stone acts at right angles to
the string and the pull in the string will be 20 N.

40. To sum up: The pull in the string varies between 15 N and 25 N, but the acceleration is all
the time 4g and, of course, the centripetal force is all the time 20N. From the practical point of view
what matters most is the pull in the string, which is obviously most likely to break when the stone is in
position D and the tension is at the maximum value of 25 N.

41. Co-relation of Linear and Angular Motion. The following formulae for circular motion,
similar to those for linear motion, may be derived for uniform angular velocity and uniform angular
acceleration.
Let’s assume: ω1 = Initial velocity in radians per sec.
ω2 = Velocity in radians per sec after t sec.
α = Angular acceleration in radians per sec² .
θ = Angle traversed, in radians.
9 Mechanics

Using the basic equations of motion we have:-


(a) ω2 = ω1 + αt (as v = u +at)
(b) θ = ( ω1 + ω2 ) t (as θ = Average Velocity X Time)
2
(c) θ = ω1t + ½ α t² (as s = ut + ½at²)
(d) ω1² - ω2²= 2αθ (as v² - u² = 2as)
(e) Since ω = v/r, therefore angular acceleration α will be
α = ω
t
α = v X 1
r t
α = a
r

42. Vector representation of Angular


Quantities. Angular velocity and acceleration
are vector quantities. By convention they are
represented by vectors. The vector length
represents the magnitude of the quantity and it’s
direction is perpendicular to the plane of rotation,
i.e. parallel to the axis of rotation. The direction of
the arrow is such that, on looking in its direction,
the rotation is clockwise. The convention is known
as the ‘right handed screw law’. Such vectors can
be combined and resolved according to the normal
principles of vectors.

43. Centre of Gravity (C.G.). It is that point Fig 1-7: Vector Representation of Angular
Moments
through which the weight of the body can be
considered to act.

Moments and Couples

44. Moment of a Force. The moment of a force around a point is the product of the force and
the perpendicular distance between that point and the line of action of the force.

45. Addition of Moments. The moment (or turning effect) of a force can be either clockwise or
anticlockwise. The net effect of a system of force can be calculated by adding all the moments
algebraically. The clockwise moments by convention are taken as positive while the anticlockwise
moments are designated negative.
FIS Book 1: Aerodynamics 10

46. Couples. A system of two equal, unlike, parallel forces is called a couple. A couple can
cause no motion in translation but causes only rotation. The moment of a couple is equal to the
product of one force and the perpendicular distance between the two forces. The work done by a
couple is the product of the moment of the couple and the angle in radians described by the arm.

47. Moment of Inertia. It was shown that a body travelling with uniform speed in a circle has an
acceleration of v²/r towards the centre of the circle. The force producing this acceleration is termed as
the centripetal force, and for a body of mass ‘m’ the centripetal force is mv²/r towards the centre of the
circle. Now consider a rigid body, such as a flywheel, free to rotate about an axis through its centre, at
an angular velocity ‘ω’. When the wheel is rotating, all the particles of the wheel have the same
angular velocity, but their linear velocities will depend on their individual distances from the axis, those
on the rim moving much faster than those near the axis.

48. A particle of mass ‘m’, at distance ‘r’ from the centre and with linear velocity ‘v’, has a kinetic
energy equal to ½ m v², or ½ m r² ω².

The total kinetic energy of all the particles is ½ ω² X Σ m r²


(Σ = Epsilon which denotes summation).

49. The sum of the products mr² for all the particles in a rigid body rotating about a given axis is
called the Moment of Inertia (I) and I = mr2.

50. Moment of inertia can be considered as the rotational equivalent of mass. As in linear motion,
if a force is applied to a body then force = mass x acceleration, likewise in circular motion if a torque is
applied to a wheel, then

Torque = Moment of Inertia X Angular Acceleration.

51. Radius of Gyration. The radius of gyration is a useful concept whereby the mass M (= Σ
m) of the wheel is considered to be concentrated in a ring of radius ‘k’ from the axis, such that Σ mr² =
Mk². (k is then known as the radius of gyration and I = Mk²).

52. Angular Momentum. If the mass (m) of a body is concentrated at a radius (r) from the axis
of rotation, the angular momentum is defined as the product of moment of inertia and angular velocity.

Angular momentum = IXω


= m r² ω

Alternatively, angular momentum can also be calculated from the product of the instantaneous linear
momentum (m v) and the radius.
11 Mechanics

Angular momentum = m v r or m r² ω

53. Conservation of Angular Momentum. If no external forces are applied then the angular
momentum of a system remains constant. As we know that angular momentum is the product of
moment of inertia and angular velocity, therefore, if one is decreased, the other must increase to
conserve the angular momentum.

54. Equilibrium. When two or more co-planar forces act on a rigid body, the body (or system of
forces) is said to be in equilibrium if the following conditions are satisfied:

(a) The algebraic sum of the resolved parts of all the forces in same fixed direction
should be equal to zero.
(b) The algebraic sum of the resolved parts of all the forces in the direction at right
angles to the first direction should be equal to zero.
(c) The algebraic sum of the moments of all the forces around a given point in their plane
should be equal to zero.

Equations of Motion: Kinematics

55. Kinematics is the study of the bodies in motion.

The symbols and units used for the derivations shall be as follows:
Time = t (sec)
Distance = s (metres)
Initial Velocity = u (m/s)
Final Velocity = v (m/s)
Acceleration = a (m/s²)

Uniform Velocity

If velocity is uniform at u meters per sec then clearly:


Distance travelled = Velocity X time
or s = ut

Uniform Acceleration

Final velocity = Initial velocity + Increase of velocity

Increase in velocity will be the product of acceleration and time i.e. a X t and therefore,

v = u + at (1.8)
FIS Book 1: Aerodynamics 12

Distance travelled = Average Velocity X Time

s = (v + u) X t (1.9)
2

Substituting the value of v from equation (1.8) we get

s = (u + at + u) X t
2
i.e., s = u t + ½ a t² (1.10)

Extracting the value of t form equation (1.8) we get

t = v-u and putting this value into equation (1.9) we get


a
s = (v + u) (v -- u) = v² - u²
2a 2a
or v² = u² + 2as (1.11)

With the aid of these simple formulae it is easy to work out problems of uniform velocity or uniform
acceleration.

Forms of Energy

56. Energy can exist in various forms e.g. kinetic energy, potential energy, pressure energy, heat
energy etc. The capacity to do work is called Energy. The amount of work the body can do, will be the
measure of its energy.

57. Kinetic Energy. It is the energy a body possesses by virtue of its motion and is measured
by the amount of work that the body can perform against the impressed force before its velocity is
destroyed or if a body starts from rest then the energy required to accelerate it to the final velocity
would be the measure of its kinetic energy.

If a body of mass m starts from rest under a force F, then


v² = u² + 2as
v² = 0 + 2as
s = v2
2a
KE = Fs = ma.s = m a v2
2a
KE =½ mv2 (1.12)

58. Potential Energy. It is the energy a body possesses by virtue of its position and is
measured by the amount of work that the body can perform in moving from its initial position to some
standard position called the zero position. For bodies moving under the influence of the earth's gravity
13 Mechanics

the zero position is usually the earth’s surface.

Work or Energy = Force X distance


But F = m.a where a = g for a body under the influence of earth’s gravitational field or
F = m.g = Weight
h is the height of the body above the earth’s surface (Zero position).
Therefore, PE = mgh (1.13)

Also PE = Wh (∵ m g = W)

59. Pressure Energy. It is the energy a fluid possesses by virtue of its compression. Suppose a
Pressure P is applied to a plate of area A causing it to move through a distance s then,

Work or Energy = Force X distance

∵ Force = Pressure X Area = PA


∴ Energy = (PA) . s

∵ Area X Length (Distance) = volume

Energy = Pressure X Volume (∵ volume = m / ρ )


Energy = P X m / ρ where ρ is the density
Pressure energy = PV or P X m (1.14)
ρ

Units and Dimensions

60. There are various systems in use for measuring physical quantities like length, mass and
time, e.g., the CGS, FPS and metric systems. The most important system used now-a-days is the SI
or System International.

61. As per the SI system there are three basic physical quantities i.e. mass, length and time.
These have the dimensions M, L and T respectively. Many other quantities are merely combinations
of these fundamental dimensions M, L and T.

Base Units

Quantity Name of Unit Symbol Dimension


Length metre m L
Mass kilogram kg M
Time second s T
Temperature Kelvin k -
Plane angle radian rad 1
FIS Book 1: Aerodynamics 14

Derived Units

Quantity Name of the unit Symbol Unit Equivalent to Dimension


Speed / Velocity - - m/s LT-¹
Acceleration - - m/s² LT-²
Density - - kg/m³ ML-³
Force Newton N Kg m/s² MLT-²
Pressure/ Stress Pascal Pa N/m² M L-¹ T-²
Work / Energy Joule J Nm ML²T-²
Power Watt W Nm/s or J/s ML²T-³
Weight Kilogram force kg f kg m/s² MLT-²
Strain - - - 1
Moment - - N-m ML²T-²
Momentum - - Kg m/s MLT-¹
Angular Velocity - - Rad /s T-¹
Angular - - Rad / s² T- ²
Acceleration

Gas Laws

62. BoyIe’s Law. Under constant temperature conditions, the volume of a given quantity of a
perfect gas is inversely proportional to its pressure.

i.e. P ∝ 1/V
or PV = constant

63. CharIe’s Law. If the pressure of a given mass of a perfect gas remains constant, then it’s
volume increases by 1/273 of its value at 0º C for every degree centigrade rise in temperature. Thus,
if temperature is measured in Kelvin then,

V0 = V1 = V2 or V = constant (1.15)
T0 T1 T2 T

The law further states that if the volume of a given mass of a perfect gas remains constant, then its
pressure increases by 1/273 of its value at 0ºC for every degree centigrade rise in temperature. If
temperature is measured in Kelvin then,

P0 = P1 = P2 or P = constant (1.16)
T0 T1 T2 T

However, when a gas is heated, P, V and T all change. Therefore, combining these laws, we get
15 Mechanics

P1V1 = P2V2 or PV = constant (1.17)


T1 T2 T

∵ Density, ρ = m/V
So for a given mass of gas (m = 1),
ρ ∝ 1 or V ∝ 1
V ρ
Therefore, _P1 = _P2_ = constant (1.18)
ρ1T1 ρ2T2

64. Gas Equation.

Since PV = constant, it can be expressed as


T
PV = RT where R is called the gas constant. (1.19)
The numerical value of R depends upon the molecular weight and quantity of the gas in
question (e.g. R for air = 288 Jules per Kg Kelvin)

65. Laws of Fluid Pressure. The following are the laws of fluid motion:-

(a) In a fluid at rest the pressure at any point acts equally in all directions.
(b) In a fluid at rest the pressure acts at right angles to the surface with which the fluid is
in contact.
(c) In a fluid at rest the pressure increases with depth. If the fluid is incompressible the
increase in pressure is directly proportional to depth. If the fluid is compressible the rate of
increase of pressure increases with depth.

66. Adiabatic Changes. A change is said to be adiabatic if no heat is taken from or given to the
surroundings during the change.

67. Isothermal Changes. A change is said to be isothermal if the temperature of the object
under change remains the same as that of the surrounding throughout the change.

68. Friction. When two solid surfaces which


are in contact move, or tend to move, relative to
each other, a force acts in the plane of contact of the
surfaces in a direction opposing the motion (Fig 1-8).
This is known as friction. Various theories have been
propounded to explain this phenomenon, but it will Fig 1-8: Friction between Two Bodies
suffice here to summarise the observed effects.

69. Experiments have shown that:


FIS Book 1: Aerodynamics 16

(a) If the two surfaces are not in relative motion, the force of friction balances the applied
force and is known as static friction. As the applied force is increased so the force of static
friction increases until it is at a maximum when the surfaces are just on the point of sliding.
This is called the limit of static friction or limiting friction for the surfaces.

(b) Once motion has started, the force needed to maintain motion without acceleration is
less than that needed to overcome the static friction. The friction between the moving
surfaces is called sliding or kinetic friction.

(c) Different materials experience different friction forces, and friction is dependent on the
smoothness of the surfaces.

70. The generally accepted laws of friction are:

(a) When the surfaces are at rest, friction can take any value up to a certain maximum
called the limiting friction.

(b) The friction force F, when the surfaces are about to slide or are sliding, is within limits,
proportional to the normal reaction (R) between the surfaces. Thus it can be stated that F =
µ R (Fig 1-8) where, µ is a constant (over a wide range of values of R) called the coefficient
of friction for the surfaces. If at rest but just at the verge of moving, the constant of friction
considered is the called coefficient of static friction. When a body is in motion then the
constant of friction is called the coefficient of sliding friction.

(c) The force of friction is independent of the area of contact between the surfaces.

(d) The sliding force is independent of the sliding velocity. The force of friction depends
on the nature of the surfaces.

Machines

71. A machine is any device which enables energy to be used in a convenient way to perform
work. Typical examples of simple machines are levers, winches, inclined planes, pulleys and screws.

Efficiency = Work done on load


Work done by effort
= Out put
Input

Efficiency is usually expressed in percentage. As friction must be overcome and as a part of the
applied effort is expended in making parts of a machine move, the efficiency of a machine is always
less than 100%.
17

CHAPTER 2

THE ATMOSPHERE AND AIRSPEEDS

Fig 2-1: The Atmosphere


FIS Book 1: Aerodynamics 18

THE ATMOSPHERE

Introduction

1. Atmosphere is the term given to the layer of air, which surrounds the Earth and extends
upwards from its surface to about 500 miles (800 km). The flight of all objects using fixed or moving
wings to sustain them, or air breathing engines to propel them, is confined to the lower layers of
atmosphere. The properties of atmosphere are therefore of great importance to all forms of flight.

2. The Earth's atmosphere can be said to consist of four concentric gaseous layers. The layer
nearest to the surface is known as the troposphere, above which are the stratosphere, the
mesosphere and the thermosphere. The boundary of troposphere, known as the tropopause, is not at
a constant height but varies from an average of about 25,000 ft at the poles to 54,000 ft at the
equator. Above the tropopause, stratosphere extends to approximately 30 miles. Ionosphere is the
region of our atmosphere extending from roughly 40 to 250 miles altitude in which there is appreciable
ionisation. The presence of charged particles in this region, which starts in the mesosphere and runs
into the thermosphere, profoundly affects the propagation of electromagnetic radiations of long
wavelengths (radio and radar waves).

3. Through these layers the atmosphere undergoes a gradual transition from its characteristics
at sea level to those at the fringe of the thermosphere, which merges with space. The weight of the
atmosphere is about one millionth of that of Earth, and an air column one square metre in section
extending vertically through the atmosphere weighs 9800 kg. Since air is compressible, the
troposphere contains a greater part (over three quarters in middle latitudes) of the whole mass of the
atmosphere, while the remaining fraction is spread out with ever-increasing rarity over a height range
of some hundred times that of the troposphere.

4. Average representative values of atmospheric characteristics are shown in Fig 2-1. It will be
noted that the pressure falls steadily with height, but that temperature falls steadily upto the
tropopause, where it then remains constant through the stratosphere, but increasing for a while in the
warm upper layers. Temperature falls again in the mesosphere and eventually increases rapidly in
the thermosphere. The mean free path (M) in Fig 2-1 is an indication of the distance of one molecule
of gas from its neighbours. Therefore in the thermosphere, although the individual air molecules have
the temperatures shown, their extremely rarefied nature results in a negligible heat transfer to any
body present.

Physical Properties of Air

5. Air is a compressible fluid and as such it is able to flow or change its shape when subjected
even to minute pressures. At normal temperatures, metals such as iron and copper are highly
19 The Atmosphere and Airspeeds

resistant to deformation by pressure, but in liquid form they flow readily. In solids the molecules
adhere so strongly that large forces are needed to change their position with respect to other
molecules. In fluids, however, the degree of cohesion of the molecules is so small that very small
forces suffice to move them in relation to each other. A fluid in which there is no cohesion between
the molecules and therefore no internal friction i.e. inviscid, and which is incompressible would be an
"ideal" fluid - if it were obtainable.

6. Fluid Pressure. At any point in a fluid the pressure is same in all directions, and if a body is
immersed in a stationary fluid the pressure on any point of the body acts at right angles to the surface
at that point irrespective of the shape or position of the body.

7. Composition of Air. Since air is a fluid having a very low internal friction it can be
considered, within limits, to be an ideal fluid. Air is a mixture of a number of separate gases, the
proportions of which are:

Element By Volume % By Weight %


Nitrogen 78.08 75.5
Oxygen 20.94 23.1
Argon 0.93 1.3
Carbon dioxide 0.03 0.05
Hydrogen
Neon
Helium
Krypton Traces Only
Xenon
Ozone
Radon

Table 2-1: Composition of Atmosphere

It can be seen from table 2-1 that for all practical purposes the atmosphere can be regarded as
consisting of 21% oxygen and 78% nitrogen by volume. Up to a height of some five to six miles water
vapour is found in varying quantities. The amount of water vapour in a given mass of air depends on
the temperature and whether the air is, or has recently been, over large areas of water. Higher the
temperature, greater is the amount of water vapour that the air can hold.

Standard Atmosphere

8. The values of temperature, pressure, density, viscosity etc. are never constant in any given
layer of the atmosphere, in fact, they are all constantly changing. Experience has shown that there is
a requirement for a standard atmosphere for the comparison of aircraft performances, calibration of
FIS Book 1: Aerodynamics 20

altimeters and other practical uses. Some standard values have been established by the International
Civil Aviation Organization (ICAO), and this set of figures, giving variation of physical properties of air
with altitude, is known as the International Standard Atmosphere (ISA).

9. In the ISA sea level values are denoted by the suffix 0. It is also useful to refer to relative
pressure, relative temperature and relative density, by which is implied in each case the local value
divided by the sea level value. The manner in which these parameters vary in the standard
atmosphere is given in the succeeding paragraphs.

Temperature

10. Measurement of Temperature. The Celsius scale is normally used for recording
atmospheric temperatures and working temperatures of engines and other equipment. On this scale
water freezes at 0° and boils at 100°, at sea level. In scientific measurement of temperature absolute
zero has a special significance. At this temperature a body is said to have no heat whatsoever.
Temperatures relative to absolute zero are measured in Kelvin (also known before 1967 as degrees
Kelvin) and are used in all formulae dealing with density and pressures. Kelvin (K) zero occurs at -
273.15° Celsius (C).

11. Conversion Factors. Occasionally the Fahrenheit (F) scale may still be encountered. It is
not suitable for scientific purposes, as Fahrenheit zero has no particular significance. Water at sea
level freezes at 32° F and boils at 212° F. To convert °F to °C, subtract 32° and multiply by 5/9. To
convert °C to °F, multiply by 9/5 and add 32°. To convert °C to Kelvin or °K (Absolute) add 273° (or
more precisely 273.15°).

12. The standard sea level temperature of air, T0, is 15°C or 288.15 K. The temperature falls
steadily upto the tropopause, where it then remains constant through the stratosphere, but increasing
for a while in the warm upper layers. Temperature falls again in the mesosphere and eventually
increases rapidly in the thermosphere. The relative temperature θ (theta) = T / T0, where T and T0 are
both expressed in K.

Pressure

13. The standard sea level pressure is 1013.25 mb. The pressure at any point in a stationary fluid
is determined by the weight of fluid in question. Same is the case with air in the atmosphere. It follows
that the pressure must fall steadily with increasing altitude. However, because the density also falls,
the rate at which the pressure falls must decline as the altitude increases. The relative pressure
δ (delta) = p/p0.
21 The Atmosphere and Airspeeds

14. Pressure Altitude. If the sea level pressure should be other than 1013.25 mb, then there is
a requirement to relate the actual pressure experienced at a given point to a height (above or below
MSL) in the ISA at which the same pressure will be found. This is pressure altitude, and is widely
used in Operating Data Manuals and performance charts. Pressure altitude can be found either by
calculating a height change of 30 ft for every 1 mb difference in pressure away from 1013.25 mb at
the surface, or by setting the sub-scale of an ICAN calibrated altimeter to 1013.25 mb and reading
pressure altitude directly from the instrument.

Density

15. Density (ρ) is mass per unit volume. The unit of density used in this book is kg per m³. The
relationship of density to temperature and pressure can be expressed as:

p / ρT = R (2.1)
Where p = pressure in mb
T = absolute temperature
R = gas constant (= 8313 J/Mol °K for any gas and
for air R = 287 J/kg °K).
The relative density σ = ρ / ρ0.

16. Effect of Pressure on Density. When air is compressed, a greater amount can occupy a
given volume, i.e. the mass, and therefore the density, has increased. Conversely, when air is
expanded less mass occupies the original volume and the density decreases. From equation (2.1), it
can be seen that, provided the temperature remains constant, density is directly proportional to
pressure, i.e. if the pressure is halved, so is the density, and vice versa.

17. Effect of Temperature on Density. When air is heated it expands so that a smaller mass
will occupy a given volume, therefore the density will have decreased, assuming that the pressure
remains constant. The converse will also apply. Thus the density of the air will vary inversely as the
absolute temperature, as indicated by equation (2.1). In the atmosphere the fairly rapid drop in
pressure as altitude is increased has the dominating effect on density, as against the effect of the fall
in temperature, which tends to increase the density.

18. Effect of Humidity on Density. The preceding paragraphs have assumed that the air is
perfectly dry. In the atmosphere some water vapour is invariably present. This may be almost
negligible in certain conditions but in others the humidity may become an important factor in the
performance of an aircraft. The density of water vapour under standard sea level conditions is 0.760
kg per m³. Therefore water vapour can be seen to weigh 0.760/1.225 as much as air, roughly 5/8 as
much as air at sea level. This means that under standard sea level conditions the portion of a mass
of air, which holds water vapour, weighs (1 – 5/8), or 3/8 less than it would if it were dry. Therefore air
FIS Book 1: Aerodynamics 22

is least dense when it contains a maximum amount of water vapour and most dense when it is
perfectly dry.

19. Density Altitude. For aircraft operations air density is usually expressed as a density
altitude. Density Altitude is defined as that height (above or below mean sea level) in the standard
atmosphere to which the actual density at any particular point corresponds. For standard conditions
of temperature and pressure, density altitude is the same as pressure altitude. Density Altitude equals
pressure altitude ± (120 X t), where t is the difference between local air temperature at pressure
altitude and the standard temperature for the same pressure altitude. If the air temperature is higher
than standard, then (120 X t) is added to pressure altitude, if it is lower, subtracted.

Viscosity

20. Coefficient of Viscosity (µ). The value of coefficient of viscosity of air at sea level in the
-5
standard atmosphere, µ0, is 1.79 X 10 kg / ms. The coefficient, µ, for any given gas, depends

directly on its temperature (µ is proportional to T¾). Thus µ falls steadily with increasing altitude until
the tropopause is reached and remains constant in the stratosphere.

21. Kinematic Viscosity. An alternative measure of the viscosity of a gas is its kinematic
–5
viscosity, ν, defined by the formula ν = µ / ρ. Its standard sea level value is 1.46 X 10 m²/s. In the
troposphere µ & ρ both reduce with increasing altitude but ρ does so more rapidly. In the
stratosphere ρ falls while µ is constant. Therefore ν increases with altitude in both regions.

Altitude (ft) Tempe- Pressure Pressure Density Relative


rature (°C) (mb) (psi) (Kg per m³) Density (%)

0 +15.0 1013.25 14.7 1.225 100.0


5,000 +5.1 843.1 12.22 1.056 86.2
10,000 -4.8 696.8 10.11 0.905 73.8
15,000 -14.7 571.8 8.29 0.771 62.9
20,000 -24.6 465.6 6.75 0.653 53.3
25,000 -34.5 376.0 5.45 0.549 44.8
30,000 -44.4 300.9 4.36 0.458 37.4
35,000 -54.3 238.4 3.46 0.386 31.0
40,000 -56.5 187.6 2.72 0.302 24.6
45,000 -56.5 147.5 2.15 0.237 19.4
50,000 -56.5 116.0 1.68 0.186 15.2

Table 2-2: ICAO Standard Atmosphere


23 The Atmosphere and Airspeeds

22. The ISA. To summarise the main parameters, the ISA assumes a mean sea level
temperature of +15°C, a pressure of 1013.25 mb (101.325 KN / m2 or 760 mm of Mercury) and a
density of 1.225 kg per m3. The temperature lapse rate is assumed to be uniform at the rate of 6.5° C
per kilometre (1.98° C per 1,000 ft) up to a height of 11 km (36,090 ft) above which height it remains
constant at -56.5° C (Table 2-2).

23. Speed of Sound. The speed of sound in a gas, a, is an important parameter which will be
formally defined in the chapters on high speed flight of this book. In a given gas a = α0 √T. Its
standard sea level value, a0, is 340 m/s or 661.2 kts.

Dynamic Pressure

24. Because it possesses density, air in motion must possess energy and therefore exerts a
pressure on any object in its path. This dynamic pressure is proportional to the density and the
square of the speed. The energy due to movement, the kinetic energy (KE), of one cubic metre of air
moving at a stated speed is given by the formula:

KE = ½ ρV2 joules Where: ρ is the local air density in kg per m³


V is the speed in metres per second.

One Joule = the work done when one Newton force displaces the point of its application
through a distance of 1 metre in the direction of the force.

If this volume of moving air is completely trapped and brought to rest by means of an open-ended
tube, the total energy remains constant. In being brought to rest the kinetic energy becomes pressure
energy (small losses are incurred because air is not an ideal fluid) which, for all practical purposes, is
equal to ½ ρ V2 Newton per m2, or if the area of the tube is S square metres, then:

Total Pressure (dynamic + static) = ½ ρ V2 S Newton.

One Newton = that force which applied to a mass of 1 kilogram, gives it an acceleration of 1
metre per second per second (1 kg force = 9.81 Newton).

25. The term ½ρV2 is common to all aerodynamic forces and fundamentally determines the air
loads imposed on an object moving through the air. It is often modified to include a correction factor
or coefficient. The term stands for the dynamic pressure imposed by air of a certain density moving at
a given speed and which is brought completely to rest. The abbreviation for the term ½ ρ V2 is the
symbol "q".

Note: Dynamic pressure cannot be measured on its own, as the ambient pressure of the
atmosphere (known as static) is always present. This total pressure is also known as
stagnation or pitot pressure. It can be seen that (Dynamic + Static) - Static = Dynamic.
FIS Book 1: Aerodynamics 24

MEASUREMENT OF SPEED

Air Speed and Ground Speed

26. When we speak of the speed on an aeroplane we mean its speed relative to the air, or air
speed as it is usually termed. The existence of a wind simply means that portions of the air are in
motion relative to the earth, and although the wind will affect the speed of the aeroplane relative to the
earth, i.e. its ground speed, it will not affect its speed relative to the air.

27. For instance, suppose that an aeroplane is flying from A to B (60 km apart), and that the
normal speed of the aeroplane (i.e. its air speed) is 100 kmph. If there is a wind of 40 kmph blowing
from B towards A, the ground speed of the aeroplane as it travels from A to B will be 60 kmph, and it
will take one hour to reach B, but the air speed will be 100 kmph. If, when the aeroplane reaches B, it
turns and flies back to A, the ground speed on the return journey will be 140 kmph. The time to regain
A will be less than half an hour, but the air speed will still remain 100 kmph, that is, the wind will strike
the aeroplane at the same speed as on the outward journey. Similarly, if the wind had been blowing
across the path, the pilot would have inclined his aeroplane several degrees towards the wind on both
journeys so that it would have travelled crabwise, but again, on both outward and homeward journeys
the air speed would have been 100 kmph and the wind would have been a head wind straight from
the front as far as the aeroplane was concerned.

28. An aeroplane which encounters a head wind equal to its own air speed will appear to an
observer on the ground to stay still, yet in reality it is moving with considerable velocity. A free balloon
flying in a wind travels over the ground, yet it has no airspeed (a flag on the balloon will hang vertically
downwards).

Method of Measuring Air Speed

29. We have talked about air speed, but so far we have not explained the method by which it is
usually measured, since it is essential that an aircraft have some means of measuring the speed at
which it is passing through the air. The method of doing this is by comparing the total pressure (static
+ dynamic) with static pressure (Para 24). Let’s see how and why. The resistance of a body placed
in an air stream depends, among other things, on the velocity of the stream and there is a connection
between the resistance and the velocity. If we can eliminate the “other things” we should be able to
devise an instrument, which by measuring resistance would give, if suitably calibrated, readings of
velocity.

30. If we were to place a flat plate at right angles to the airflow and behind this plate we fitted a
compression spring balance, then there would be a direct connection between the velocity of the
airflow and the reading of the spring balance, provided that the area of the plate and the density of the
25 The Atmosphere and Airspeeds

air remained constant. As regards the area, there is not much difficulty except that we must
remember that it is the “frontal” area facing the airflow which matters, and therefore we must ensure
that the plate is at all times at right angles to the airflow. The question of the density, however, is not
quite so simple, in fact, the ordinary method of measuring airspeed is dependent on the density of the
air and therefore readings of air speed can only be correct when the density is of some definite value.

Pitot Static Tube

31. The system, which is almost universally employed, is the pitot-static head. This consists of
two metal tubes fixed to some exposed part of the aeroplane and facing directly into the airflow. One
of these, the pitot (or pressure) tube, has an open end facing the wind, while the other tube, called the
static, is closed at the end but pierced with small holds or slits farther back. On modern aircraft the
tubes are usually concentric, the pitot tube forming the centre and being surrounded by the static.
The other end of each tube is connected by tubing to the air speed indicator in the cockpit.

32. The pitot tube is connected to the inside of a capsule (a flat circular box of a corrugated metal
such as is used in an altimeter of aneroid barometer), while the static tube is connected to the casing
of the instrument. When there is no air speed the normal atmospheric pressure will be communicated
to both sides of the diaphragm or capsule, but when the air blows up against the open pitot tube the
extra pressure due to the air velocity plus the normal atmospheric pressure will act on one side, while
the static tube still conveys the ordinary atmospheric pressure to the other. The difference in pressure
causes a deflection of the capsule, which, by suitable gearing, makes a pointer move round the scale,
which is marked off in air speeds.

33. It might be expected that the open pitot tube would effectively bring the air to rest. The extra
force on the open end of the pitot tube is ½ ρV2S, where S is the area of the opening. Therefore the
pressure increases on the open end due to the airflow will be force divided by the area, i.e. ½ ρV2.

34. This dynamic pressure is used to indicate the speed using an instrument known as an Air
Speed Indicator (ASI). It should be noted that both the dynamic and static pressures act on the pitot
tube while only the static pressure acts on the static tube. The ASI therefore reads the difference
between the two i.e. the dynamic pressure and gives an indication of air speed.

Relationship Between Air Speeds

35. The general term air speed is further qualified as:

(a) Indicated Air Speed (IAS) (Vi). The reading on the ASI is the indicated air speed.
Note: In practice, any instrument error is usually very small and can, for all practical
purposes, be ignored.
FIS Book 1: Aerodynamics 26

(b) Calibrated Air Speed (CAS) (Vc). When the IAS has been corrected by the
application of pressure error correction (PEC) and instrument error correction, the result is
known as Calibrated Air Speed (CAS). PEC can be obtained from the Aircrew Manual or
ODM for the type. When the PEC figures for individual instruments are displayed on the
cockpit, they include the instrument error correction for that particular ASI. CAS is also known
as RAS (Rectified Air Speed).

(c) Equivalent Air Speed (EAS) (Ve). Equivalent air speed is obtained by adding the
compressibility error correction (CEC) to CAS.

(d) True Air Speed (TAS) (V). The TAS is obtained by dividing the EAS by the square
root of the relative air density.

36. The ASI can be calibrated to read correctly for only one density / altitude. All ASIs are
calibrated for ICAO atmosphere conditions. Under these conditions and at the standard sea level
density (ρ0), the EAS (Ve) is equal to the TAS (V). At any other altitude where the density is ρ then,
Ve = V √ (ρ / ρ0) = V √ σ. Thus, at 40,000 ft where the standard density is one quarter of the sea level
value, the EAS will be half the TAS.

37. From the preceding paras it can be seen that there are two important speeds. The TAS is
significant because it gives a measure of the speed of a body relative to the undisturbed air and the
EAS is significant because the aerodynamic forces acting on an aircraft are directly proportional to the
dynamic pressure and thus the EAS.

The Venturi Tube

38. Having talked much about airspeeds


let us see exactly which part of the aircraft
this indicated air speed is correct for.
Towards this let us consider a venturi and the
flow through it. As per Bernoulli’s theorem,
as the cross section area of the venturi
towards the throat reduces, the flow velocity
increases, considering the flow density
remaining constant. Therefore, if a complete
pitot - static tube is connected to the air
speed indicator and inserted at various parts
of the venturi, it will be clear that the air
Fig 2-2: The Venturi Tube
speed increases from mouth to throat and
27 The Atmosphere and Airspeeds

then decreases again to the exit. The air speed increases very nearly in the same proportion as the
area of cross-section of the venturi decreases (Fig 2-2).

39. Let us illustrate our points by taking a numerical example. Suppose the area of the throat of a
venturi tube is a quarter of the area of the mouth, and suppose the speed of airflow at the mouth is 60
m/s. If we neglect any change of density of the air as it flows through the venturi, the speed of flow at
the throat will be 240 m/s. Now if a complete pitot-static head is placed at the mouth, and is
connected to an air speed indicator, the reading will, of course, be 60 m/s, which is the air speed at
the mouth. If the complete head is placed at the throat, the reading will be 240 m/s.

40. But suppose a pitot tube is placed at the throat, where the speed is 240 m/s and the static
side of the air speed indicator is connected to a static tube at the mouth then what will the air speed
indicator read?

41. In order to determine this let us call:

The static pressure at the mouth, i.e. the atmospheric pressure, Pa


Velocity at mouth, Va
Static pressure at throat, Pb
Velocity at throat, Vb
The air density will be the same at both, i.e. ρa = ρb = ρ.

Then on the pitot tube at the throat the total pressure will be the sum of the static and
dynamic pressures.

i.e. Pb + ½ ρ Vb2

Which, by Bernoulli’s Theorem, is the same as Pa + ½ ρ Va2

On the static tube at mouth, the pressure will be just Pa

Therefore pressure on diaphragm in the air speed indicator will be the difference between
these two,

i.e. Pb + ½ ρ Vb2 – Pa
Or
Pa + ½ ρ Va2 – Pa which equals ½ ρVa2

Now when the net pressure on an air speed indicator is ½ ρVa2, ρ being the standard sea-
level density, the instrument reads Va.

Therefore, in this example, the reading will be 60 m/s.


FIS Book 1: Aerodynamics 28

Notice that the speed recorded is that of the airflow where the static is, and not where the
pitot is.

42. Let us reverse the positions of pitot and static and see what happens.

Place the pitot at the mouth were the air speed is 60 m/s and the Static at the throat where it
is 240 m/s.

Total pressure on pitot = Pa + ½ ρVa2


= Pb + ½ ρVb2 (By Bernoulli’s theorem)
Pressure on static = Pb

Therefore difference between the two = Pa + ½ ρ Va2 – Pb


= Pb + ½ ρVb2 – Pb
= ½ ρVb2

Therefore the indicator reads Vb i.e. 240 m/s, again a speed corresponding to the position of the static
and not the pitot.

43. The explanation is that since the total pressure in the pitot tube is the sum of the dynamic and
the static pressure it remains the same at any part of the air stream, in accordance to the Bernoulli’s
theorem. Therefore so long as the pitot tube is in the stream, it makes no difference at what part of
the stream it is. The static pressure, on the other hand, must be correct for the place where we wish
to measure the air speed.

44. It should be noted that the pitot tube must be in the smooth air stream whose velocity we wish
to measure. It is no good putting it behind an obstruction, nor in the slipstream from the propeller
where there will be greater energy, nor in any place where there is turbulence, because Bernoulli’s
Theorem is only true for streamline flow. But it can be in a position where the air is flowing at a higher
or lower speed than that of the aeroplane, and, provided the static gives the true atmospheric
pressure, the air speed indicator will read the correct speed of the aeroplane.

45. Mach Number. Small disturbances generated by a body passing through the atmosphere
are transmitted as pressure waves, which are in effect sound waves, whether audible or not. In
considering the motion of the body it is frequently found convenient to express its velocity relative to
the velocity of these pressure waves. This ratio is Mach number (M). M = V/a where V = TAS and a
= local speed of sound in air.

TAS Calculations

46. Given below are useful methods of converting RAS to TAS by simple mental processes.
29 The Atmosphere and Airspeeds

(a) TAS = RAS + RAS X altitude (in thousands of feet) (2.1)


60
or, more conveniently :

TAS = RAS (kmph) + RAS (km per min) X altitude (thousands of feet)
Usually, for pilot navigation, IAS may be substituted for RAS.

Example

RAS 240 kmph (4 km per min), altitude 15,000 ft.


TAS = 240 + (4 X 15) = 300 kmph

It should be noted that this method is valid only below 25,000 ft and is increasingly inaccurate
above 550 kmph (300 kt) TAS because of compressibility effects. (acceptably accurate upto
M 1.0).

(b) An alternative method gives the air speed correction as a percentage, or fraction of
RAS to be added to the RAS. The figures have been ‘rounded-out’ for convenience and are
based on ISA conditions only, but are sufficiently accurate for use in pilot navigation
(Table 2-3).

Height (ft) Percent (%) Aide de Memoir Fraction


5,000 7½ 3² 1/12
10,000 15 4² 1/6
15,000 25 5² 1/4
20,000 35 6² 1/3
25,000 50 7² 1/2
30,000 60 8² 3/5
35,000 75 9² 3/4
40,000 100 10² 1
45,000 120 11² 6/5 or 5/4
50,000 150 12² 1½

Table 2-3: RAS to TAS Conversion

These figures are also subject to increasing compressibility errors above 550 kmph (300 kt) TAS.
FIS Book 1: Aerodynamics 30
31

CHAPTER 3

THEORIES OF DEVELOPMENT OF AERODYNAMIC FORCES

1. Definitions

(a) Total Reaction (TR). The resultant of all the aerodynamic forces acting on the wing
or aerofoil section.

(b) Lift. That component of the TR, which is perpendicular to the flight path or Relative
Air Flow (RAF).

(c) Drag. That component of TR which is tangential to the flight path i.e. parallel to the
RAF.

(d) Chord Line. A straight line joining the centers of curvature of the leading and
trailing edges of an aerofoil.

(e) Chord (c). The distance between the leading and trailing edge measured along the
chord line. Mean chord is often used as a datum linear dimension in the same way that the
wing area (S) is used as a datum area e.g. for the LCA the chord is about 4.5 m and it is
about 5.6 m for the MiG 29 etc.

(f) Span (b). Span of a wing is the distance between the two wing tips and is usually
denoted as b, the examples are. 50.5 m for the IL 76, 11.36 m for the MiG 29 etc.

(g) Wing Area (S). Area of the wing projected on a plane perpendicular to the normal
axis e.g. about 40 m for the Mirage 2000 and 23 m2 for the MiG 21 etc.
2

(h) Mean Line or Camber Line. A line joining the leading and trailing edges of an
aerofoil equidistant from the upper and lower surfaces. Maximum camber is usually
expressed as a ratio of the maximum distance between the camber line and the chord line to
chord length. Where the camber line lies above the chord line, the aerofoil is said to have
positive camber.

(j) Angle of Attack (α). Angle between the chord line and the flight path or RAF. In
many old textbooks this was referred to as Incidence.

(k) (Rigger’s) Angle of Incidence. Angle at which an aerofoil is attached to the


fuselage. It is the angle between the mean chord line and the longitudinal fuselage datum.
FIS Book 1: Aerodynamics 32

The term is often used erroneously instead of Angle of Attack. Examples of angle of
0 0
incidences are 1 for the hawk, 2.5 for the HPT 32 etc.

(l) Thickness / Chord Ratio (t/c). Maximum thickness or depth of an aerofoil section
expressed as a percentage of chord length e.g. 5.5% for the SU 30 and 10% for the Hawk
aircraft.

(m) Centre of Pressure (CP). A point, usually on the chord line, through which the TR
may be considered to act.

(n) Streamline. It is the path traced by a particle in a steady fluid flow. It can also be
defined as an imaginary line drawn in the field of flow such that the velocity vector at any point
on the line is always tangential to the line. It follows from this definition that through every
point in the field there passes one, and only one, streamline. In general, no two streamlines
can intersect each other, since this would imply that at the point of intersection the velocity
vector points in two directions at once, and this is clearly inconsistent. There is one
exceptional case where a streamline may intersect itself i.e. at a point in the flow where the
local velocity is zero. Here there is no velocity vector, and so no inconsistency. Such a point is
called 'stagnation point'.

(p) Aspect Ratio (AR). = Span = b or ___Span2__ = b2


Chord c Wing Area S

(q) Wing Loading. The weight per unit area of the wing := __Weight__ = W
Wing Area S

(r) Load Factor (g or n). = Total Lift


Weight

(s) Particle Path. This is simply the path


through the field followed by an individual fluid particle.
In a steady flow, the velocity vector is clearly always
tangential to this path, so that in steady flow a particle
path and a streamline are the same.

(t) Stream Tube. Consider an imaginary closed


curve drawn in the field of flow, as in Fig 3-1. If all the
streamlines passing through the points on this curve
are drawn, they generate a tubular surface, which is
called a stream tube. Since the velocity vector is
Fig 3-1: Stream Tube
always tangential to the surface of such a tube, it
33 Theories of Development of AD Forces

follows that there is no flow into or out of the tube through the 'walls', but only along the tube.

(u) Stream Filament. This is a stream tube of infinitesimal cross-section.

Types of Flow

2. Steady Streamlined Flow. In a steady streamline flow the flow parameters (e.g. speed,
direction, pressure etc.) may vary from point to point in the flow but, at any point, are constant with
respect to time. This flow can be represented by streamlines and is the type of flow, which it is hoped,
will be found over the various components of an aircraft. Steady streamline flow may be divided into
two types:

(a) Classical Linear Flow. Fig


3-2 a illustrates the flow found over a
conventional aerofoil at low angle of
attack in which the streamlines, more
or less, follow the contour of the body
Fig 3-2 a: Classical Linear Flow
and there is no separation of the flow
from the surface.

(b) Controlled Separated Flow or Leading Edge Vortex Flow. This is a halfway
stage between steady streamline flow and unsteady flow described later. Due to boundary
layer effects, generally at a sharp leading edge, the flow separates from the surface, not
breaking down into a turbulent chaotic
condition but instead, forming a strong
vortex, which because of its stability and
predictability, can be controlled and made to
give a useful lift force. Such flows illustrated
in Fig 3-2 b are found in swept and delta
planforms, particularly at the higher angles of
attack, and are dealt with in more detail in
later chapters. Fig 3-2 b: Leading Edge Vortex Flow

3. Unsteady Flow. In this type of flow the flow parameters vary with time and the flow cannot
be represented by streamlines.

4. Two-dimensional Flow. If a wing is of infinite span, or, if it completely spans a wind tunnel
from wall to wall, then each section of the wing will have exactly the same flow pattern round it except
near the tunnel walls. This type of flow is called two-dimensional flow since the motion is confined to
a plane parallel to the free stream direction.
FIS Book 1: Aerodynamics 34

5. Three-dimensional Flow. The wing on an aircraft has a finite length and, therefore,
whenever it is producing lift the pressure differential tries to equalize around the wing tip. This
induces a span-wise drift of the air flowing over the wing, inwards on the upper surface and outwards
on the lower surface, producing a three-dimensional flow.

6. Vortices. Because the effect of the spilling at the wing tip is progressively less pronounced
from tip to root, the amount of transverse flow reduces towards the fuselage. As the upper and lower
airflows meet at the trailing edge they form vortices, small at the wing root and larger towards the tip.
These form one large vortex in the vicinity of the wing tip, rotating clockwise on the port wing and anti-
clockwise on the starboard wing, as viewed from the rear. Tip spillage means that an aircraft wing
can never produce the same amount of lift as an infinite span wing. If the wing has a constant section
and riggers angle of incidence from root to tip then the lift per unit span of the wing may be considered
to be virtually constant until about 1.2 chord distance of the wing tip.

7. Vortex Influences. The


overall size of the vortex at the
trailing edge will depend on the
amount of the transverse flow.
Therefore, the greater the force
(pressure difference), the larger it
will be. The familiar pictures of
wing-tip vortices showing them as
thin white streaks (Fig 3-3), only
show the low pressure central core
and it should be appreciated that the
influence on the airflow behind the
trailing edge is considerable. The
Fig 3-3: Wing Tip Vortices Visible as White Streaks
number of accidents following loss
of control by flying into wake vortex turbulence testifies to this. The vortex upsets the balance
between the upwash and downwash of two-dimensional flow, reinforcing the downwash and reducing
the effective angle of attack, also inclining the lift vector slightly backwards, as the effective relative
airflow is now inclined downwards. The component of TR, parallel to the line of flight is increased.
This increase is termed as the induced drag. The resolved lift vector perpendicular to the flight
direction is reduced.

8. The pattern of the airflow round an aircraft at low speeds depends mainly on the shape of the
aircraft and its attitude relative to the free stream flow. Other factors are the size of the aircraft,
density & viscosity of air and speed of the airflow. These factors are usually combined to form a
parameter known as Reynolds Number “RN” (discussed later in the chapter) and the airflow pattern is
then dependent only on the shape, attitude and Reynolds Number.
35 Theories of Development of AD Forces

9. Reynolds Number (i.e. size, density, viscosity and speed) and condition of surface determines
the characteristics of the boundary layer. This, in turn, modifies the pattern of the airflow and
distribution of pressure around the aircraft. The effect of boundary layer on the lift produced by the
wings may be considered insignificant throughout the normal operating range of angles of attack. In
later chapters it will be shown that the behavior of boundary layer has a profound effect on the lift
produced at high angles of attack.

10. When considering the velocity of the airflow it does not make any difference to the pattern
whether the aircraft is moving through the air or the air is flowing past the aircraft. It is the relative
velocity which is the important factor.

AERODYNAMIC THEORIES

General

11. Several methods or theories have been developed to predict the performance of a given
wing/aerofoil shape. These can be used to explain the subtle changes in shape necessary to produce
the required performance appropriate to the role of the aircraft. In practice, the appropriate wing
shape is calculated from the performance criteria.

12. The theories discussed in this volume are: Lift by Pressure Distribution, Momentum Theory,
Dimensional Analysis and Circulation Theory.

LIFT BY PRESSURE DISTRIBUTION

13. The most useful, non-mathematical method is an examination of the flow pattern and
pressure distribution on the surface of a wing in flight. This approach will reveal the most important
factors affecting the amount of lift produced, based on experimental (wind-tunnel) data.

14. Basic Aerodynamic Theory. The shape of the aircraft (and boundary layer) will determine
the velocity changes and consequently the airflow pattern and pressure distribution. For a simplified
explanation of why these changes occur it is necessary to consider:

(a) The Equation of Continuity.


(b) Bernoulli’s Theorem.

Equation of Continuity

15. The Equation of continuity states that mass can neither be created nor destroyed, or simply
stated, air mass flow remains constant.
FIS Book 1: Aerodynamics 36

16. Consider the steady flow of an inviscid fluid through


the stream filament shown in Fig 3-4. Let Al be the area of
cross-section at any station 1, A2 at any other station 2, both
small. Let ρ1, VI and ρ2, V2 be the fluid density and velocity at
these stations respectively, each effectively constant over the
given section since the area is, by definition, small. Since the
fluid is a continuous medium, and there is no flow across the
walls of the filament, and since the flow is steady, so that there
is no build up of fluid in the region bounded by stations 1 and
2, it follows that the rate of mass flow into this region at station
1 must be equal to the rate of mass flow out of the region at Fig 3-4: Steady Flow of Inviscid
station 2, i.e., Fluid Though a Stream Tube

ρI A1 VI = ρ2 A2 V2

Since stations 1 and 2 are arbitrarily chosen, it follows that the expression ρ A V has the
same value at any such station, i.e.,

ρ A V = Constant (along a stream filament) (3.1)

17. This equation is the simplest form of what is known as the Equation of Continuity. For
incompressible flow, ρ is a constant, so that the equation reduces to AV = constant. In compressible
flow theory it is convenient to assume that changes in fluid density will be insignificant at speeds
below about 0.4 M. This is because the pressure changes are small and have little effect on the
density. The equation of continuity may now be simplified to:

AV = constant (3.2)

This is the general equation of continuity and applies to both compressible and incompressible fluids.
From this it may be seen that a reduction in the tube’s cross-sectional area will result in an increase in
velocity and vice versa. This equation enables the velocity changes around a given shape to be
predicted mathematically.

18. These equations remain valid for a stream tube of finite size provided that the density and
velocity do not vary across the section. If they do vary, then the rate of mass flow across a section
may be written as

∫ ρ V. dA,
A

so that the equation of continuity becomes


∫ ρ V . dA = constant, for compressible flow
A

∫ V. dA = constant, for incompressible flow (3.3)


A
37 Theories of Development of AD Forces

BERNOULLI'S THEOREM FOR INCOMPRESSIBLE FLOW

Bernoulli’s Theorem

19. Consider a gas in steady motion. It possesses the following types of energy:

(a) Potential energy due to height.


(b) Heat (Kinetic) energy.
(c) Pressure (Potential) energy.
(d) Kinetic energy due to motion,

In addition work and heat may pass in or out of the system.

20. Daniel Bernoulli demonstrated that in the steady streamline flow of an ideal fluid, the sum
of the energies present remained constant. It is emphasized that the words in bold represent the
limitations of Bernoulli’s experiments. In low subsonic flow (< 0.4 M), it is convenient to regard air as
being incompressible and inviscid (i.e. ideal) and predictions of the pressure changes round a given
aerofoil section agree closely with measured values. Above 0.4 M, however, these simplifications
would cause large errors in predicted values and are no longer permissible.

21. For a mathematical derivation of Bernoulli’s


equation, consider again the flow of an
incompressible, inviscid fluid through a stream
filament, as shown in Fig 3-5. Consider in particular
the mass of fluid bounded by stations I and 2. Let ρ1,
A1, V1, ρ2, A 2, V2, have the same connotations as in
paragraph 16. In addition, let p1 be the fluid pressure
at station 1, which is located at height h1 above a fixed
reference level. Let p2, h2 have similar connotations at
station 2. Next consider the same mass of fluid a short
Fig 3-5: Flow of Ideal Fluid through a
time δt later. Station 1 will have moved through a Stream Tube
distance Vl δt along the tube, station 2 a distance V2 δt.

Since the fluid is incompressible, ρ1 = ρ2 = ρ (3.4)


And by continuity Al Vl = A2 V2 = AV (3.5)

The work done by the pressure forces at stations 1 and 2 will then be given by

W = p1 A1 V l δ t – p2 A2 V2 δt = A V (pl – p2) δt (3.6)

Since the fluid is inviscid, there are no tangential stresses at the wall of the tube, so that this is
the total work done on the given mass of fluid.
FIS Book 1: Aerodynamics 38

The gain in kinetic energy is

( ½ ρ 2 A 2 V 2 δt V22 ) – ( ½ ρ 1 A 1 V 1 δt V12 ) = { ½ ρ A V (V22 – V12 ) δt } (3.7)

The gain in potential energy is

(ρ2 V2 A2 δt g h2) – (ρ1 V1 A δt g h1 ) = { ρ A V δ t g (h2 – h1)} (3.8)

where g is the acceleration due to gravity. The principle of conservation of energy gives

work done = gain in kinetic energy + gain in potential energy.


A V (p 1 – p 2) δ t = {½ ρAV (V22 – V12 ) δt } + { g ρ V A (h2 – h1) δt }
pI – p 2 = { ½ ρ V22 } – { ½ ρ V12 } + { g ρ h2 } – { g ρ h1}
p 1 + ½ ρ V12 + ρ g h 1 = p 2 + ½ ρ V22 + ρ g h 2 (3.9)

22. Again, since stations 1 and 2 are arbitrarily chosen, it follows that the expression p+½ρ
2
V + ρ g h has the same value at all stations along the stream filament. If we now consider the limiting
case when the filament shrinks to a single streamline, we have:

p + ½ ρ V2 + ρgh = constant, along a streamline. (3.10)

23. This equation, which is an energy equation, is known as Bernoulli's Equation. When
considering the problem of an aircraft, or part of an aircraft moving through the atmosphere, the
changes in the value of ρgh within the field of flow affected by the presence of the aircraft are
generally very small compared with the changes in the values of the other two parameters in the
equation. It is therefore a valid and useful approximation, in such problems, to write

p + ½ ρ V2 = constant, along a streamline.

24. Further, at some distance upstream of the aircraft, the flow consists of a uniform stream. It
follows that on any given streamline in this region the value of p + ½ ρV2 is the same as it is on any
other streamline. Thus the value of this expression is constant, not simply along a given streamline,
but throughout the whole field of flow. Bernoulli's Equation is then written as

p + ½ ρ V2 = constant (3.11)
and applies throughout the field.

25. Another simplified way of arriving at the expression of equation (3.11) is by assuming that in
low subsonic flow, changes in potential energy and heat energy are insignificant and that there is no
transfer of heat or work. For practical purposes therefore, in the streamline flow of air round a wing at
low speed:

Pressure energy + kinetic energy = constant


39 Theories of Development of AD Forces

This simplified law can be expressed in terms of pressure, thus:

p + ½ ρV2 = constant, Where: p = Static pressure


ρ = Density and
V = Flow velocity

The significance of this law will be recognized if it is translated into words: static pressure + dynamic
pressure is a constant. This constant is referred to as Total Head Pressure, Stagnation Pressure or
Pitot Pressure.

26. One of the most interesting examples of Bernoulli’s theorem is provided by the venturi tube
(Fig 3-6). In the diagram, at point A, B & C the mass flow remains constant, or in other words

ρ 0 A0 V0 = ρ 1 A1 V1 = ρ 2 A2 V2
Or ρAV = K (3.12)

27. Before reaching the constriction, the


flow has a certain velocity and pressure. As
soon as the flow reaches the constriction or
the venturi, the velocity increases and
pressure decreases which can be marked
by the streamlines becoming closer, and
again past the venturi the flow pressure
increases and the velocity decreases
marked by the streamlines becoming further
apart. This phenomenon is of great
Fig 3-6: Flow Through a Venturi Tube
importance later on in the development of an
aerofoil.

28. It has already been stated that the flow velocity is governed by the shape of the aircraft. From
Bernoulli’s Theorem (simplified) it is evident that an increase in velocity causes a decrease in static
pressure and vice versa.

29. Velocity Indication. As the air flows round the aircraft its speed changes. In subsonic flow
a reduction in the velocity of the streamline flow is indicated by an increased spacing of the
streamlines whilst increasing velocity is indicated by decreased spacing of the streamlines.
Associated with the velocity changes there will be corresponding pressure changes.

30. Pressure Differential. As the air flows towards an aerofoil it will be turned towards the low
pressure (partial vacuum) at the upper surface, this being termed as ‘upwash’. After passing over the
aerofoil the airflow returns to its original position and state. This is termed as ‘downwash’ and is
FIS Book 1: Aerodynamics 40

shown in Fig 3-7. The reason


for the pressure and velocity
changes around an aerofoil is
explained by the Bernoulli’s
Theorem. The difference in
pressure between the upper
and lower surface of an
aerofoil is usually expressed
as relative pressure by ‘–’ and
‘+’. However, the pressure Fig 3-7: Two-dimensional Flow around an Aerofoil
above is usually a lot lower
than ambient pressure and the pressure below is usually slightly lower than ambient pressure (except
at high angles of attack), i.e. both negative.

Static, Total and Dynamic Pressure

31. It follows from Bernoulli's Equation that, in a given field of flow, any increase in velocity is
accompanied by a reduction in pressure, and vice versa. The pressure is lower at points in the flow
where the speed is higher. The greatest pressure is achieved when the speed is zero, i.e., at
stagnation points, and the pressure at such points is called the stagnation pressure, usually denoted
by po. The pressure and velocity at any given point in the flow are then related by the equation

p + ½ ρV2 = constant = po. (3.13)

32. This shows that the stagnation pressure is a constant, i.e., that the pressure is the same at all
stagnation points in a given flow, and further that the concept of stagnation pressure has significance
even in a flow where there are no actual stagnation points. For this reason, it is often referred to, not
as stagnation pressure but, as the total pressure associated with the flow. (This discussion
presupposes that the flow originates in a uniform stream as described in para 24. If it does not, then
the concept of total pressure remains valid, but its value may vary through the field, as it does also in
certain types of high-speed flow.)

33. Consider now the quantity po - p, i.e., the difference between the total pressure of the flow and
the local value of the pressure. This quantity has the dimensions of pressure and is a 'pressure'
associated with the motion of the fluid. It is therefore called the dynamic pressure. In the case of
incompressible flow considered here, the dynamic pressure is equal to ½ ρV2. To distinguish it from
the dynamic and total pressures, the actual pressure of the fluid, which is associated not with its
motion but with its state, is often referred to as the static pressure, but where the term pressure alone
is used it refers to this static pressure.
41 Theories of Development of AD Forces

34. The following observations, arising from the above definitions, should be noted:

(a) Total pressure = static pressure + dynamic pressure.


(b) Total pressure is constant throughout the field of flow.
(c) The dynamic pressure and static pressure generally vary through the field.
(d) At a stagnation point the total pressure = the static pressure, and the dynamic
pressure is zero.
(e) Nowhere can the static pressure (or the dynamic pressure) exceed the total pressure.

Pressure Distribution Around an Aerofoil

35. Although the whole aircraft contributes towards both lift and drag it may be assumed that the
wing is specifically designed to produce the necessary lift for the whole of the aircraft. Examination of
the distribution of pressure round the wing is the most convenient non-mathematical way to see how
the lift is produced.

36. Pressure distribution round an aerofoil is usually measured with a manometer. This consists
of a series of glass tubes filled with a coloured liquid and connected to small holes in the aerofoil
surface. As air flows over the aerofoil the variations in pressure are indicated by the differing levels of
fluid in the tubes. These measurements are in absolute pressure and it is useful to compare them to
ambient pressure in order to see the lifting effort produced by the aerofoil.

37. The pressure at a point on the aerofoil surface may be represented by a vector perpendicular
to the surface whose length is proportional to the difference between absolute pressure at that point
and free stream static pressure, i.e. proportional to (p - po). It is usual to convert this to a non-
dimensional quantity called the pressure coefficient (CP) by comparing it to free stream dynamic
pressure (q), thus: Cp = ( p – p0 ) / q

38. The convention for plotting these pressure coefficients is as follows:

(a) Measured Pressure Higher Than Ambient Pressure. At these points, (p - po) will
be positive, giving a positive CP and the vector is plotted towards the surface.

(b) Measured Pressure Lower Than Ambient Pressure. Here, (p - po) will be
negative, resulting in a negative CP which is plotted away from the surface.

39. It is useful to consider the value of the CP at the leading edge stagnation point where the air is
brought to rest. The absolute pressure will be total head pressure = free stream static pressure + free
stream dynamic pressure. The pressure coefficient will therefore be:

(p0 + q) – p0 and therefore CP = +1


q
FIS Book 1: Aerodynamics 42

40. Each of the pressure coefficient


vectors will have a component
perpendicular to the free stream flow
which is, by definition, a lift component.
Because the pressure is plotted in
coefficient form, the lift component will
also be a coefficient, and is the local lift
coefficient at that point. It is possible
therefore to obtain the total lift coefficient
from the pressure distribution, by
subtracting all the lift components pointing
down (relative to the free stream) from all
those pointing up (Fig 3-8).

41. Inspection of the pressure


distribution diagrams gives an indication of
the direction and magnitude of the total
reaction (TR) and the position of the Fig 3-8: Pressure Distribution around an Aerofoil
centre of pressure (CP). In particular, two
facts are immediately apparent which will
be of considerable use in later chapters:

(a) The lift coefficient


increases with an increase in
angle of attack.
(b) The CP moves forward
with an increase in angle of
attack.

42. A common way to illustrate the


pressure distribution round an aerofoil
section is to plot it in the form of a graph
(Fig 3-9). Notice that, whereas in the
previous diagrams the CP values are Fig 3-9: Graphical Representation of
Aerofoil Pressure Distribution
plotted perpendicular to the aerofoil (Angle of Attack Approx. 100)
surface, in the graph the CP values are
perpendicular to the chord line. Note also the convention of plotting negative values upwards to relate
it to lift in the natural sense.
43 Theories of Development of AD Forces

AERODYNAMIC THEORIES

43. Momentum Theory is another way of looking at Bernoulli’s Theorem, while Dimensional
Analysis is a mathematical derivation of aerodynamic force and takes into account the effect of
Reynolds Number and Mach Number.

Lift by Momentum Theory

44. If a vane or blade deflects a jet of fluid as in Fig 3-10


then, assuming frictionless flow past the vane, there will be
no loss of velocity in the jet.

45. From Newton’s second law:

F = ma = rate of change of momentum


= ( m . dv)
dt
Fig 3-10: Flow Past a Vane
If A is the cross-section area of the jet and ∆V is the
net change in velocity,

F = m X velocity change
Time
= _m_ X velocity change
time
= mass flow X ∆V

∴ F = ( ρAV ) X ∆V Fig 3-11: Vector Diagram

and from the vector diagram in Fig 3-11

∆V = √2 X V ≈ 1.4 V
∴ F = 1.4 ρAV2

and substituting q for ½ ρV2

F = 2.8 q A

46. Thus, by merely deflecting the stream a force of up to 2.8 q A may be produced. This principle
can be applied on an aerofoil, which can be said to produce an aerodynamic advantage by changing
the momentum of a stream of air. It can also be seen that the force produced depends on the attitude
in the stream (α) and on the curvature of the aerofoil.
FIS Book 1: Aerodynamics 44

Lift by Dimensional Analysis

47. It can be demonstrated experimentally that the force acting on geometrically similar objects
moving through the air is dependent on at least the following variables:

(a) Free stream velocity.


(b) Air density.
(c) The size of the object.
(d) The speed of sound, i.e. speed of propagation of small pressure waves.
(e) The viscosity of the air.
(f) Shape, attitude and smoothness of the object.

48. It is possible to derive explicit equations for lift and drag by the use of dimensional analysis.
This principle relies on the fact that in any physical equation the dimensions of the left-hand side
(LHS) must be the same as the dimensions of the right-hand side (RHS). The dimensions used are
those of mass, length and time (MLT) which are independent of specific units. The factors listed in
para 47 are repeated below together with their appropriate symbols and dimensions. There is no
convenient method of measuring shape, attitude or smoothness, so these variables are included in a
constant (k).

Dependent Variables Symbol Dimensions


Velocity V LT -1
Density ρ ML -3
Area (wing) S L2
Speed of Sound a LT -1
Viscosity µ ML -1T -1

49. The dependence of aerodynamic force on V, ρ, S, a and µ may be expressed as:

F = k (V . ρ . S . a . µ )

which theory allows to be written as

F = k (V a . ρ b . S c . a d . µ e ) (3.14)
Where k, a, b, c, d and e are unknown.

50. Equation (3.14) may now be rewritten in terms of the MLT system, remembering that

force = mass X acceleration


M L T-2 = k { (L T-1)a (M L-3)b (L2)c(L T-1)d (M L-1 T-1)e } (3.15)
45 Theories of Development of AD Forces

51. The condition for dimensional equivalence between the two sides of equation (3.13) is that
the sum of the indices of a given quantity on the RHS must equal the index of that quantity on the
LHS. Applying this rule we get:

for mass b+e = +1


for length a - 3b + 2c + d - e = +1
for time -a - d - e = -2

52. The three equations in para 51 contain five unknowns and so any three of these can be
calculated in terms of the other two. Considering a and µ of secondary importance, we can solve the
simultaneous equations above in terms of d and e and thus obtain:

a = 2 - d - e,
b = 1 - e,
c = 1–e
2

53. Substituting these values of a, b and c in equation (3.12) we get:

F = k { (V)2 - d – e. (ρ)1 – e. (S) 1 - e/2. (a)d. (µ)e }

and collecting terms of like indices we get:


d e
F = k ρ V2 S a __µ__ (3.16)
V ρ V S½

54. Dimensionally, √ S = √ (L2 ) = L and so the last term of the equation (3.16) is __ µ__
ρVL
which is , _______1_______ = __1__
Reynolds Number RN
also a = _______1______ = _1_
V Mach Number M

55. Equation (3.16) therefore becomes F = k (ρ V 2 S M - d RN - e ).

Noting that dynamic pressure = ½ ρ V2, we can introduce another constant, C, such that
C = 2 k and obtain:

F = C ½ ρ V2 S (M -d RN -e) (3.17)

56. It is significant that this solution includes unknown exponential functions of Mach number and
RN, a fact which is often ignored when the lift and drag equations are quoted. In practice, at the
sacrifice of an exact equation, the parameters M and RN are included in the constant C. C, therefore,
FIS Book 1: Aerodynamics 46

becomes a coefficient dependent on Mach number, Reynolds Number, shape, attitude and
smoothness.

Thus C = 2 k M -d RN -e

57. Equation (3.17) is a general equation for aerodynamic force and represents either the lift
component, the drag component or their vector sum (TR), i.e.

Lift = CL ½ρ V2 S
And Drag = CD ½ρ V2 S
Similarly, moment = CM ½ ρ V2 S, where CM is a function of M, RN, etc.

The Circulation Theory of Lift

58. A complete discussion of the circulation theory of lift demands the adoption of a formal
mathematical approach which is beyond the scope of this book. It is intended here to set out a
physical outline of the theory in order to provide a convincing account of the mechanism whereby lift is
created. Some results are quoted in mathematical form, and these are here to be taken on trust.

59. A basic assumption on which the analysis is founded is that air is an inviscid fluid. It will
ultimately be necessary to recognize that the viscosity, though very small, is non-zero, but this will
have little effect on the magnitude of the results quoted. The error in these results which arises from
the fact that air is slightly viscous is small. Threfore the assumption of inviscid flow is justifiable, in that
it gives rise to a simple, tenable and useful theory of lift.

60. A further assumption of the theory as presented here is that the flow is two-dimensional. By
this it is implied that the lift-producing body is an infinitely long cylinder of any given section, whose
axis is normal to the free stream velocity vector. There is then no component of velocity parallel to the
axis, and the flow in any one plane normal to this axis is identical to that in any other such plane.
There are thus only two degrees of freedom. It follows, of course, that the forces and moments acting
on such a body are generally infinite. They are therefore quoted in terms of the force or moment
acting on unit length of the cylinder, in which case they are finite and meaningful. The concept of two-
dimensional flow is widely used, but any results obtained must later be modified to allow for the three-
dimensional nature of practical problems.

Circular Cylinder without Circulation

61. A cylinder of circular section is not, of course, a practical shape for an aircraft wing,
nevertheless, the circulation theory is most easily understood as it applies to a circular cylinder, and
the results may then be extended to account for the lift produced by a two dimensional wing. Consider
first, therefore, the flow past an infinite circular cylinder placed in a uniform stream of air, without
47 Theories of Development of AD Forces

circulation. The streamline pattern which


results is depicted in Fig 3-12.

62. The line ABCD, together with the


circle itself, is a streamline, called the dividing
streamline, since it separates the fluid which
passes over the top of the cylinder from that
which passes below it. The points B and C are Fig 3-12: Flow Past a Circular Cylinder
stagnation points. The following facts should be noted:

(a) The streamlines are close together near the top and bottom of the cylinder, and far
apart near the front and rear. This implies, by continuity, high velocity at top and bottom, and
low velocity at front and rear.

(b) The stagnation points are symmetrically located at opposite ends of the horizontal
diameter.

(c) The streamline pattern is doubly symmetric, i.e., symmetrical about the two
perpendicular diameters BC and EF. This implies that the velocity distribution, and hence, by
Bernoulli's Theorem, the pressure distribution, are also doubly symmetric.

(d) It follows that there is no lift and no drag.

Circulation

63. Circulation must now be formally


defined. Consider any imaginary closed curve C
within the fluid, as shown in Fig 3-13. Let the
velocity at a point P on C be q, at an angle θ to
the tangent to C at P. The product of the
tangential component qs of velocity with an
element of length δs is then:

qs. δs = q. cos θ . δs
Fig 3-13: Circulation

The circulation round C, denoted by Г (Gamma,


referred to as K in some books) is then defined as the integral of this product, evaluated round the
contour C. Hence

Г = ∫qs . ds = ∫ q . cos θ . ds.


C C
FIS Book 1: Aerodynamics 48

64. A simple example of a flow with circulation is a


so-called free vortex. The streamlines of such a flow
are concentric circles, and the speed is everywhere
inversely proportional to the distance from the centre of
the circles, as shown in Fig 3-14.

Round any given circle, the circulation is then:

Г = k X 2πr = 2πk (3.18)


r
Fig 3-14: Circulation around any Circle

It shows that in this case the circulation is constant, i.e. it has the same value whichever circle is
chosen. It must be emphasized, however, that the existence of circulation as defined above does not
necessarily imply any rotation of fluid particles around a point or body. Circulation can exist in a flow in
which the streamlines are straight lines.

65. The two important properties of vortex flow are:

(a) Velocity. In the flow round a vortex the velocity is inversely proportional to the
radius, i.e. radius X velocity = constant.

(b) Pressure. Because the flow is steady, Bernoulli’s Theorem applies and the
pressure decreases as the velocity increases towards the centre of the vortex.

66. A strong vortex will have high rotational speeds associated with it and vice versa. The
constant K determines the size of the rotational speeds and it is therefore used to measure vortex
strength.

Circular Cylinder with Circulation

67. Consider now a circular


cylinder in a uniform stream, with
circulation Г (say, a free vortex of that
strength) superimposed. Suppose that
the circulation is in a clockwise sense.
The streamline pattern is then as
depicted in Fig 3-15. The total velocity at
B is now clearly in the upward sense, so
that B must lie above the dividing Fig 3-15: Circular Cylinder with Circulation
streamline, as must C. There is some
point B', below B, where the velocity associated with the flow of Fig 3-12 is equal and opposite to the
velocity associated with the circulation effect, i.e., with the vortex. This point is therefore the new
49 Theories of Development of AD Forces

stagnation point, and there is a second such point symmetrically placed at C'. The following should
then be noted as the effects of superimposing the circulation on the original flow:

(a) The stagnation points are moved downwards. The amount of downward movement
depends on the magnitude of the circulation in relation to the free stream speed, i.e., on the
ratio Г / V.

(b) The effect of the circulation is generally to increase the speed over the upper surface
of the cylinder and to reduce the speed over the lower surface. This effect is shown by the
spacing of the streamlines in Fig 3-15.

(c) From Bernoulli's Theorem, therefore, it follows that the pressure is generally reduced
on the upper surface and increased on the lower surface. As a result, there is a net force
vertically upwards. This is lift.

(d) The streamline pattern is still symmetrical about the diameter EF. It follows that there
is still no drag.

68. It can be shown that lift per unit length of the cylinder = ρVГ. (3.19)

In fact, the value is the same for a cylinder of any section, circular or otherwise, and in particular for
any two-dimensional wing. This result is known as the Kutta-Zhukowski relation.

69. If a vortex is in a stream of air and some


restraint is applied to the vortex to prevent it
being blown downstream, the streamlines of the
flow will be as shown in Fig 3-16.

70. The point S is of some interest. It is a


stagnation point in the stream, and the direction
Fig 3-16: Flow Round a Vortex Rotating in a
of flow passing through S is shown by the Uniform Stream
arrows. The air which passes through S splits
into two, some going over the top of the vortex and some below it. That part of the streamline through
S which passes over the vortex forms a closed loop and the air within this loop does not pass
downstream but circulates continually round the vortex. The velocity at any point is the vector sum of
the velocity of the uniform stream and the velocity due to the vortex in isolation. Thus it can be seen
that the air above the vortex is speeded up and that below the vortex is slowed down.

71. As a result of these velocity differences between the flow above and below the vortex, there
are pressure differences. These pressure differences cause a force on the vortex of magnitude ρVГ
FIS Book 1: Aerodynamics 50

per unit length which acts perpendicular to the direction of the free stream, i.e. it is a lift force by
definition.

72. Again, it can be shown that the drag of any two-dimensional body of any shape or size, when
placed in a stream of inviscid incompressible fluid with or without circulation is still zero. This result is
known as the Paradox of d' Alembert. It follows that, while lift can be accounted for and estimated as
above, assuming that the fluid is inviscid, it is necessary, in order to explain the existence of drag, to
take account of viscosity.

Flow Past a Two Dimensional Aerofoil

73. Consider now an infinite cylinder whose section is not circular but shaped like a section of an
aircraft wing. We shall call such a section an aerofoil section, or aerofoil. Suppose that such an
aerofoil is placed at a given attitude in a
uniform stream of inviscid fluid, without
circulation. The streamline pattern will be as
shown in Fig 3-17. There are again two
stagnation points. In this condition, there is
no lift or drag on the aerofoil, just as in the Fig 3-17: Streamline Pattern past an Aerofoil
case of the circular cylinder. Without Circulation

74. When this aerofoil is placed at some


positive angle of attack and is accelerated
from rest, the circulation around it, and
therefore lift, is not generated
instantaneously. At the instant of starting,
the flow does not have any circulation Fig 3-18: Streamline Pattern past an Aerofoil
around it and the streamlines are as shown in Fig 3-18, with the rear stagnation point occurring on the
rear upper surface. At the sharp trailing edge, the air is required to change direction suddenly.
However, because of viscosity, the large viscosity gradient produce large viscous forces, and the air is
unable to flow around the sharp trailing edge. Instead, a surface of discontinuity emanating from the
sharp trailing edge is rolled up into a vortex, which is called “starting vortex”. The stagnation point,
under the influence of this starting vortex, moves towards the trailing edge, as the circulation around
the aerofoil, and therefore lift increases progressively. The circulation around the aerofoil increases in
intensity until the flows from the upper and the lower surfaces join smoothly at the trailing edge (Fig 3-
20c). Thus, the generation of circulation around the wing and the resultant lift are necessarily
accompanied by a starting vortex, which results because of the effects of viscosity.

75. Lets consider another way in which the lift by circulation theory can be explained. Firstly once
again consider the aerofoil of Fig 3-17 generating no lift or drag. Suppose now that a circulation is
51 Theories of Development of AD Forces

superimposed on this aerofoil, while the free stream speed remains constant. Fig 3-18 then depicts
the streamline pattern that will result. Circulation is accompanied by downward movement of the
stagnation points. The rear stagnation point is on the upper surface for small values of Г, but moves
on to the lower surface when Г is large. The effect of angle of attack and / or camber on the flow
pattern round the aerofoil is to accelerate the air
over the top surface and decelerate the air along the
bottom surface. Thus the flow over an aerofoil can
be regarded as the combination of two different
types of flow, one of these is the normal flow round
the symmetrical fairing at zero angle of attack (Fig 3-
17) and the other is a flow in which air circulates
Fig 3-19: Circulation Imposed Around an
around the aerofoil, towards the trailing edge over
Aerofoil
the upper surface and towards the leading edge
over the lower surface (Fig 3-19).

76. Suppose now that various amounts of


circulation are superimposed, while the free stream
speed remains constant. Fig 3-20 then depicts the
streamline patterns that will result. Increasing
circulation is accompanied by downward movement
of the stagnation points. Since the rear stagnation
point is on the upper surface for small values of Г,
but moves on to the lower surface when Г is large.
There is clearly one intermediate value of Г for which
it is neither, and the flow leaves the trailing edge
smoothly, as indicated in Fig 3-20(c). It is this last
condition which is of the greatest importance, since
it is this one which actually occurs in the case of an
aerofoil in a real fluid. It is known as the Kutta
Condition.

77. In all the cases depicted in Fig 3-20, the


downward movement of the stagnation points
Fig 3-20: Airflow about an Aerofoil
indicates the existence of circulation: in the with Different Amounts of
clockwise sense. This implies increased speed, and Circulation Imposed
hence reduced pressure, on the upper surface, and
reduced speed, and hence increased pressure, on the lower surface, as in the case of the circular
cylinder. Thus lift is created, the amount depending on the amount of circulation. It is given, as already
noted above, by ρ V Г per unit length.
FIS Book 1: Aerodynamics 52

78. The lift of the aerofoil due to camber and angle of attack must therefore be associated with
the circulatory flow. A special type of flow which is the same as that of Fig 3-17 is that due to a
vortex, that is a mass of fluid rotating in concentric circles.

The Kutta Condition

79. In the above paragraphs, the effects of superimposing circulation on the flow past a circular
cylinder and two-dimensional aerofoil have been discussed without reference to the mechanism
whereby the circulation is created. It would appear that the performance of an aerofoil in a given
attitude may vary according to the amount of circulation. In fact, there is only one possible solution in
the case of an aerofoil in a real fluid, and the amount of circulation is uniquely determined as a result
of the viscosity of the fluid.

80. The aerofoil has a sharp trailing edge. In


the case of inviscid flow round a sharp corner,
the velocity at the corner is always infinite, in the
Fig 3-21: Formation of Vortex
case of a real flow, it would be very large. As a
result, because of the large velocity gradients, despite the very small coefficient of viscosity of air,
significant viscous stresses are set up, and the air, instead of following the contour of the aerofoil
round the corner, forms a vortex, as depicted in Fig 3-21. The effect of this vortex is to induce a
downward flow on the upper surface of the aerofoil, thus imparting a circulation effect, and causing
the stagnation point to move downwards. As long as the stagnation point is forward of the trailing
edge, the vortex persists, and so the stagnation point continues to move rearwards. Similarly if, in any
attitude, the stagnation point tends to lie on the lower surface, the vortex will be created in the
opposite sense, and will again cause the stagnation point to move towards the trailing edge.

81. Only when the flow leaves the trailing


edge smoothly, as in Fig 3-20(c), will a stable
condition be achieved. This is the Kutta
Condition. The vortex is then detached from the
trailing edge of the aerofoil and left behind
Fig 3-22: Cast Off Vortex
downstream, as in Fig 3-22. It is known as the
cast-off vortex, or starting vortex. In practice, of course, the vortex is cast off almost instantaneously. It
is quickly left far behind, and is eventually dissipated through the action of viscosity.

82. By this mechanism, then, a circulation is created which is always just sufficient to satisfy the
Kutta Condition, and which therefore determines the lift uniquely. It should be noted that:

(a) The streamline pattern is determined not just by Г but by the ratio Г / V. Thus, if the
speed is changed, the circulation will also be changed, and an additional cast-off vortex will
be created.
53 Theories of Development of AD Forces

(b) Since, given the Kutta Condition, with the aerofoil in a given attitude Г / V is
determined, then Г is proportional to V. Since the lift per unit length is ρVГ, it follows that the
lift is proportional to ρV2, i.e., to the dynamic pressure.

(c) If the attitude of the wing is changed, while the speed is kept constant, this will alter
the circulation required to satisfy the Kutta Condition. Hence another cast-off vortex will be
created, and the lift of the aerofoil will be changed.

83. The flow round an aerofoil section can be idealized into three parts:

(a) A uniform stream.


(b) A distortion of the stream due to aerofoil thickness.
(c) A flow similar to a vortex.

This is another way in which lift may be explained because earlier it was stated that a vortex in a
uniform stream experiences a lift force.

84. Assuming, for simplicity, that the lift is uniform along the span, and ignoring the effects of
thickness and viscosity, a wing
can be replaced by a vortex of
suitable strength along the locus
of the centres of pressure. This
vortex is known as line or bound
vortex, or just a lifting line, and is
illustrated in Fig 3-23. Fig 3-23: A Wing with Uniform Lift Replaced by
a Line Vortex

85. The arrangement in Fig 3-


23 is physically impossible
because a vortex cannot exist with
open ends due to the low
pressure in its core. A vortex
must therefore either be re-
entrant, or abut at each end on a
solid boundary. The simple line
Fig 3-24: The Horseshoe Vortex
vortex is therefore modified to
represent the uniformly loaded wing by a horseshoe vortex as illustrated in Fig 3-24. To make the
vortex re-entrant it is necessary to complete the fourth side of the rectangle by a suitable vortex.
When a wing is accelerated from rest, the circulation around it, and therefore lift, is not produced
instantaneously. As the circulation develops, a starting vortex is formed at the trailing edge and is left
behind the wing. This vortex is equal in strength and opposite in sense to the circulation around the
wing.
FIS Book 1: Aerodynamics 54

86. The starting vortex, or the cast off vortex, or initial eddy, can be demonstrated by simple
experiment. A flat piece of board to represent a wing is placed into water, cutting the surface at
moderate angle of attack. If the board is suddenly moved forward, an eddy will be seen to leave the
rear of the board.

87. The transformation of a wing into a vortex of suitable strength is a complex mathematical
operation. It can be seen, however, that when the value of Γ is obtained, the lift can be obtained from
the Kutta-Zhukovsky relationship. This theoretical model of an aircraft wing is most useful in
explaining three-dimensional effects such as induced drag, aspect ratio etc.

88. Magnus Effect. The creation of lift through circulation can easily be demonstrated by
rotating a circular cylinder with roughened surface in a stream of air. The rough surface of the
cylinder, through friction, causes the air in its vicinity to travel round with it, thus inducing circulation,
which gives rise to lift. This phenomenon is known as Magnus Effect, and it is largely responsible for
the swing of a sliced golf or tennis ball.

The Wake

89. In presenting the circulation theory, no account has been taken of the effects of viscosity,
except to say that it is the existence of some measure of viscosity, however small, which gives rise to
the Kutta Condition. This suggests that the amount of viscosity does not affect the value of the lift. In
fact, there is another effect which does modify this value.

90. The Kutta Condition implies that


the flow leaves the trailing edge smoothly.
Unless the trailing edge is cusped (which
in practice is never the case), it is clearly
not possible for the flow to leave both the
upper and lower surfaces smoothly. As a
Fig 3-25: Wake behind the Aerofoil
result, there is a thin turbulent region
established behind the trailing edge, extending some distance behind the wing, as shown in Fig 3-25.
This region, in which viscous effects are not negligible, is called the wake. The effect of its presence is
to reduce the circulation to a value somewhat below that predicted by the Kutta Condition, thus
establishing a rear stagnation point on the upper surface just forward of the trailing edge, and
reducing the lift of the wing. The size of the wake, and hence the magnitude of its effect on lift,
depends on a number of considerations, including the detail design of the wing section, especially on
the trailing edge angle. In practice, under normal operating conditions, it generally reduces the lift by
about 10% of the value predicted by circulation theory.
55 Theories of Development of AD Forces

REYNOLDS NUMBER

91. During the 19th Century a physicist named Osborne Reynolds was involved in experiments
with the flow of fluids in pipes and he made the important discovery that the flow changed from
streamlined to turbulent when the velocity reached a value which was inversely proportional to the
diameter of the pipe. The larger the diameter of the pipe, the lower the velocity at which the flow
became turbulent, e.g. if the critical velocity in a pipe of 2.5 cm diameter was 6 m/s then 3 m/s would
be the critical velocity in a pipe of 5 cm diameter. He also discovered that the rule applied to the flow
past any body placed within the stream. For example, if two spheres of different sizes were placed
within a flow, then the transition to turbulence would occur when the velocity reached a value which
was inversely proportional to the diameter of each sphere. Turbulence would therefore occur at a
lower speed of flow over the larger sphere than over the smaller, and furthermore, the transition point
would be at the point of maximum thickness of the body, relative to the flow. He observed that the
density, viscosity and the velocity of fluid and the diameter of the tube, all played a part in determining
whether the flow would be laminar or turbulent. He realized that he could combine the influence of all
these factors in one non-dimentional parameter viz. ρ V l which subsequently became known as
µ
the Reynolds Number, denoted by RN.

RN = ρ V l Where: ρ = density in kg per M3 (1.2250 for air at sea level).


µ V = velocity of the test in metres per second.
l = a dimension of the body (for aerofoil the chord
length is used),
µ = viscosity of the fluid.

Considering the units involved it is not surprising to see test results quoted at :
RN = 4 X 106, or even 12 X 10 6 (12,000,000).

92. If RN were relatively small (less than ½ Million) the flow was always laminar, if RN were
relatively large (more than 10 million) the flow was always turbulent. If RN had some intermediate
value (between one and five million), the flow might be either laminar or turbulent, according to the
other conditions of the experiment. The effect of RN on the boundary layer and drag characteristics
are explained in the next chapter.
FIS Book 1: Aerodynamics 56
57

CHAPTER 4

VISCOSITY AND BOUNDARY LAYER

Viscosity

1. Viscosity in fluids is related to friction in solids. It is a property inherent in all liquids and gases
and manifests itself as a tendency to retard the relative motion between adjacent layers of the fluid. In
the case of gases, viscosity increases with increase in temperature, whereas in the case of liquids,
viscosity decreases with increase in temperature.

2. Viscosity exists in the form of tangential


stress (shearing stress or force) distributed
within the fluid wherever relative motion, and
hence a velocity gradient exists in particular,
where fluid moves over the surface of an
immersed body. Consider two adjacent layers of
fluid as given in the Fig 4-1. F is the force
Fig 4-1: Viscosity
exerted per unit area (stress) between two
layers of fluid flow at velocities V1 and V2 respectively (where V1 > V2). ‘t’ is the thickness of the fluid
flow.

3. The measure of viscosity is co-efficient of viscosity. It is denoted by ‘µ’ (mu) and is defined as
the ratio of viscous stress to the velocity gradient. In the diagram above (Fig 4-1), the co-efficient of
viscosity is: -

µ = _F / Area_ X t (4.1)
V1 – V2

4. The units of coefficient of viscosity are kg/m/sec. These can also be found by ‘Dimensional
Analysis’ i.e. by substituting M, L and T values in equation (4.1). The equation will then become:

µ = M L –1 T –2 X L = M L –1 T –1 (4.2)
L T –1

The co-efficient of viscosity for air in ISA conditions is 1.79 x 10-5 kg/m/sec.

5. As already mentioned, the value of µ for a given gas depends purely upon its temperature

and it increases with temperature. In the atmosphere µ is directly proportional to T ¾ (Kelvin)


approximately. Thus the value decreases with height up to troposphere where it is constant at
1.42 x 10-5 kg/m/sec. In the stratosphere its value remains constant.
FIS Book 1: Aerodynamics 58

6. Kinematic Viscosity. A more useful measure of viscosity of a gas is its kinematic viscosity.
It is defined as the ratio of co-efficient of viscosity to density and is denoted by ν (nu)

ν = _µ_
ρ

-5
The standard sea level value of ν is 1.46 x 10 m 2/sec. In the troposphere µ and ρ both reduce with
increasing altitude, but ρ does so more rapidly. In the stratosphere µ remains constant (temperature
constant) whereas ρ continues to reduce. Thus ν increases with increase in altitude in both regions.

The Boundary Layer

7. Definition. The boundary layer is defined as that region of flow, adjacent to the surface, in
which the speed is less than 99% of the free stream flow. It can also be defined as the layer of air
extending from the surface to the point where no dragging effect is discernible or the layer of air over
the surface, which shows local retardation of airflow due to viscosity. It usually exists in two forms,
laminar and turbulent. The nature of the boundary layer affects the CL max , the stalling AOA, the value
of form drag and to some extent the high-speed characteristics of an aerofoil.

8. Because the air has viscosity, air will encounter resistance to flow over a surface. The viscous
nature of airflow reduces the local velocities on a surface and accounts for the skin friction drag. The
retardation of air particles due to viscosity is
greatest immediately adjacent to the surface.
At the very surface of an object, the air
particles are slowed to a relative velocity of
near zero. Above this area other particles
experience successively smaller retardation,
until finally, at some distance above surface,
the local velocity reaches the full value of the
air stream above the surface. This layer of
air over the surface, which shows local LAMINAR LAYER TURBULENT LAYER
FLOW IS SMOOTH & STEADY UNSTEADY & NOT SMOOTH
retardation of airflow from viscosity, is LAYER IS VERY THIN EDDIES & LAYER IS THICKER
MODERATE VEL GRADIENT LARGER VEL GRADIENT
termed the ‘boundary layer’. The LOW SKIN FRICTION HIGH SKIN FRICTION
characteristics of this boundary layer are
Fig 4-2: Development of Boundary Layer on a
illustrated in Fig 4-2 with the flow of air over
Smooth Flat Plate
a smooth flat plate.

9. The initial flow on a smooth surface gives evidence of a very thin boundary layer with the flow
occurring in smooth laminations. The boundary layer flow near the leading edge is similar to layers or
laminations of air sliding smoothly over one another and the obvious term for this type of flow is the
59 Viscosity and Boundary Layer

‘laminar boundary layer’. This smooth laminar flow exists without the air particles moving from an
elevation.

10. As the flow continues back from the leading edge, friction forces in the boundary layer
continue to dissipate energy of the air stream, and the laminar boundary layer increases in thickness
with distance from the leading edge. After some distance back from the leading edge, the laminar
boundary layer begins an oscillatory disturbance, which is unstable. A waviness occurs in the laminar
boundary layer, which ultimately grows larger and more severe and destroys the smooth laminar flow.
Thus a transition takes place in which the laminar boundary layer decays into a ‘turbulent’ boundary
layer. The same sort of transition can be noticed in the smoke from a cigarette in still air. At first, the
smoke ribbon is smooth and laminar, then develops a definite waviness, and decays into a random
turbulent smoke pattern.

11. As soon as the transition to the turbulent boundary layer takes place, the boundary layer
thickness grows at a more rapid rate. (The small-scale turbulent flow within the boundary layer should
not be confused with the large-scale turbulence associated with airflow separation). The flow in the
turbulent boundary layer allows the air particles to travel from one layer to another producing an
energy exchange. However some small laminar flow continues to exist in the very lower levels of the
turbulent layer and is referred to as the ‘laminar sub layer’. The turbulence, which exists in the
turbulent boundary layer, allows determination of the point of transition by several means. Since the
turbulent boundary layer transfers heat more easily than the laminar layer, frost, water and oil films
will be removed more rapidly from the area aft of the transition point. Also, a small probe may be
attached to a stethoscope and positioned at various points along the surface. When the probe is in the
laminar area, a low ‘hiss’ will be heard and
when the probe is in the turbulent area, a
sharp ‘crackling’ will be audible.

12. In order to compare the characteristics


of the laminar and turbulent boundary layers,
the velocity profiles (the variation of boundary
layer velocity with height above the surface)
should be compared under conditions, which
could produce either laminar or turbulent flow.
The typical laminar and turbulent profiles are
shown in Fig 4-3. The velocity profile of the
turbulent boundary layer shows a much
sharper initial change of velocity but a greater
height (or boundary layer thickness) required
to reach the free stream velocity. As a result of Fig 4-3: Boundary Layer Characteristics
these differences, a comparison will show:
FIS Book 1: Aerodynamics 60

(a) The turbulent boundary layer has a fuller velocity profile and has higher local
velocities immediately adjacent to the surface. The turbulent boundary layer has higher kinetic
energy in the airflow next to the surface.

(b) At the surface, the laminar boundary layer has the less rapid change of velocity with
distance above the plate. Since shearing stress is proportional to the velocity gradient, the
lower velocity gradient of the laminar boundary layer is evidence of a lower friction drag on
the surface. If the conditions of flow were such that either a turbulent or a laminar boundary
layer could exist, the laminar skin friction would be about one third that for turbulent flow.

13. The low surface friction drag of the laminar boundary layer makes it quite desirable. However,
the transition tends to take place in a natural fashion and limit the extensive development of the
laminar boundary layer.

Reynolds Number

14. The classical experiments of Osborne Reynolds involved flow of fluids in pipes and he made
the important discovery that the flow changed from streamlined to turbulent, when the velocity
reached a value, which was inversely proportional to the diameter of the pipe. The broader the pipe,
lower the velocity at which the flow became turbulent. He identified the governing parameters as:

ρVL (4.3)
µ
Where : ρ = density of fluid
V = mean velocity of fluid
L = diameter of the tube
µ = co-efficient of viscosity of the fluid

This number has since been known as the Reynolds Number. It is dimensionless.

Proof. Through dimensional analysis:

ρ = M . L –3 (Kg / m3)
V = L . T –1 (m / s)
L = L (m)
µ = M L –1 T –1 (From equation 4.2)
Numerator ( ρ . V . L) = (M . L –3 ) (L . T –1 ) (L)
= M L –1 T –1
Denominator (µ) = M L –1 T –1

The dimensions of both are the same. Therefore RN is dimensionless.


61 Viscosity and Boundary Layer

Derivation of Reynolds Number

15. Any particular flow experiment on geometrically similar bodies is completely determined once
we settle the speed (V), the length of the body (L), the density of the fluid (ρ) and the viscosity (µ), that
is, the Reynolds Number. Furthermore, RN is a measure of the ratio between the inertial forces and
the viscous forces of the fluid.

16. The inertial force acting on a typical fluid particle is measured by the product of its mass and
its acceleration.

∵ Mass per unit volume of the fluid is, by definition, the density (ρ).

And the volume is proportional to the cube of the characteristic length (L3).

∴ The mass is proportional to ρL3.

∵ The acceleration is the rate of change of velocity or the change in velocity divided by the
time during which the change occurs.

The change in the velocity, as the fluid accelerates and decelerates over the body, is
proportional to (V).

The time is proportional to the time taken by a fluid particle to travel the length of the body at
the speed (V), so the time is L
V
The acceleration is therefore proportional to

__V__ i.e. V2
L/V L

The inertial force, which is mass X acceleration, can now be expressed as:

Inertial Force ∝ ρ L3 X V2
L
Inertial Force ∝ ρ L2 V2 (4.4)

17. The viscous forces are determined by the product of the viscous shear stress and the surface
area over which it acts.

The area is proportional to the square of the characteristic length, that is, to L2.

The viscous shear stress is proportional to viscosity (µ), and to the rate of change of speed
with distance, that is V / L
FIS Book 1: Aerodynamics 62

∴ Viscous force ∝ L2 X µ X V
L

Viscous force ∝ LµV (4.5)

18. Comparing inertial forces with viscous forces we have:

Inertial Forces = ρ L2 V2 = ρVL


Viscous Forces LµV µ

which is the formula for Reynolds’ Number. µ/ρ is the kinematic viscosity of the fluid and when this is
constant the only variables are V and L.

19. Through further investigation Reynolds discovered that the rule applied to the flow past any
body placed within the stream. Over the years, as the speed of flight has increased phenomenally, the
design of aircraft became more sophisticated, hence wind tunnels were developed for the purpose of
testing scale models. The application of RN lies in the fact that the model aircraft and the full sized
aircraft should be compared at the same Reynolds number to determine the values of lift, drag and
study the behaviour of the boundary layer. It represents physically the ratio of the inertial and viscous
forces existing in the flow. When the RN is low, viscous or friction forces predominate and when the
RN is high, dynamic or inertial forces predominate. Thus the flow around two geometrically similar
bodies in the same relative attitude to the flow will have similar flow patterns when the RN is the
same.

Effect of RN on Boundary Layer

20. At low RN the flow is normally laminar and at high RN the flow is turbulent. In general flow
problems RN is defined as ρVL / µ where L is any representative length of the body. In the flow past
an aerofoil there are two such concepts to take account of:

(a) Overall RN defined as:

RN = ρVC Where C is the Chord length and is used in place of L.


µ

(b) Local RN (Rx) at a point, a distance x behind the leading edge denoted as:

Rx = ρ V x
µ

21. The relationship between RN and the nature of the boundary layer explains why in the flow
past an aerofoil, the boundary layer is usually laminar to begin with but turbulent further downstream.
Near the leading edge Rx is relatively small which implies laminar flow. Further downstream x
increases until eventually Rx is so large that the flow must become turbulent. At RN less than half a
63 Viscosity and Boundary Layer

million the boundary layer will be entirely laminar unless there is extreme surface roughness or
turbulence induced in the air stream. RN between one and five million produce boundary layer flow
which is partly laminar and partly turbulent. At RN above ten million the boundary layer characteristics
are predominantly turbulent.

22. From the known variation of boundary layer characteristics with RN certain general effects
may be anticipated. In the first graph of Fig 4-4, the effect of RN on the CL curve shows that at a high
value of RN the stall is delayed to a higher angle of attack and that the CL max is increased. This
merely means that if the pilot flies his aircraft to the high speed stall the wings will produce more lift
(and, incidentally, less drag) because, if the aircraft is flown at a given height the density and
temperature (which affects viscosity) are constant, and the only variable is speed. The simple reason
is that the turbulent boundary layer has greater inertial force, or greater kinetic energy, the effect of
which is to delay boundary layer separation. This results in a higher stalling angle, higher CL max, and
to complete the picture, a lower value of section drag coefficient (Fig 4-4).

Fig 4-4: Effect of RN on Section Characteristics

Critical Reynolds Number

23. We have seen that, in general, turbulent boundary layers result in much more drag than
laminar ones, and it may seem to be an advantage to ensure laminar flow if at all possible. However,
there is one respect in which the reverse may well be the case. If the flow separates, the resulting
increase in drag is many times greater than the increase resulting from transition, and there are
additional adverse effects. Also, a laminar boundary layer tends to separate much more readily than a
turbulent one. Thus, if one has a flow with a laminar boundary layer which separates early to give high
drag, transition to turbulent flow with the resultant delay in separation may well result in a substantial
reduction in drag.

24. This is well illustrated by the example of a circular cylinder (or sphere) in a real fluid, as
illustrated in Fig 4-5. The diagram on top shows the case of a flow at low Reynolds number. The
FIS Book 1: Aerodynamics 64

boundary layer is laminar, and early


separation takes place, giving a large
dead air region (known as a Von Karman
vortex street) and a high drag coefficient.
At higher Reynolds numbers, transition
takes place at a point upstream of the
separation point. As a result, the turbulent
boundary layer remains attached until a
point much further round is reached. The
size of the dead air region is much
reduced, as shown in the lower diagram of
Fig 4-5: Circular Cylinder in a Real Fluid
Fig 4-5 and so is the drag coefficient. This
phenomenon occurs quite suddenly as the
Reynolds number is increased through a
certain value which is known as the critical
Reynolds number or transition RN. Figure
4-6 illustrates the way in which the drag
coefficient of a sphere varies with
Reynolds number. In the same way,
transition from laminar to turbulent flow in
the flow past an aerofoil may often delay
the onset of stall. The affect of large RN
can be generated by artificially inducing Fig 4-6: Variation of CD at RCRIT

turbulence, like what is caused by the dimples on a golf ball, where these cause early transition
resulting in a delayed separation and thus lesser drag and a longer shot.

25. The swing of a cricket ball is


another example of this kind of effect.
Consider the ball with its seam at an angle
to the flow direction, as depicted in Fig 4-
7. The boundary layer at the front on the
side away from the seam is laminar, early
separation occurs on that side of the ball.
On the other side, the roughness of the
seam causes early transition so that
Fig 4-7: Swing of a Cricket Ball
separation is delayed. The result of this
asymmetry in the flow is a side force which produces sideways movement of the ball. The speed, and
the angle of the seam, is crucial. Also, the ball must be fairly new because, if the surface is too rough,
there will be turbulent boundary layers on both sides, without the aid of the seam.
65 Viscosity and Boundary Layer

26. Variation of Airflow over an Aerofoil with Changing RN. The characteristics depicted by
Fig 4-4 are for a typical ‘conventional’ aerofoil section. The lift curve shows a steady increase in CL max
with increasing RN. However, note that a smaller change in CL max occurs between Reynolds numbers
of 6.0 and 9.0 million than occurs between 2.6 and 6 million. In other words, greater changes in CL max
occur in the range of Reynolds number where the laminar (low energy) boundary layer predominates.
The drag curve for the section show essentially the same feature i.e. the greatest variations occur at
very low Reynolds numbers. Typical full scale Reynolds numbers for aircraft in flight may be 3 to 500
million where the boundary layer is predominantly turbulent. Scale model tests may involve Reynolds
number of 0.1 to 5 million where the boundary layer will be predominantly laminar. Hence, the ‘scale’
corrections are very necessary to correlate the principal aerodynamic characteristics.

27. The most direct use of Reynolds


number is the indexing or correlating the
skin friction drag of a surface. Fig 4-8
illustrates the variation of the surface
friction drag of a smooth flat plate with
Reynolds’s number, which is based on
the length or chord of the plate. The
graph shows separate lines of drag
coefficients if the flow should be entirely
laminar or entirely turbulent. The two
curves for laminar and turbulent friction
drag illustrate the relative magnitude of
friction drag coefficient if either type of
boundary layer could exist. The drag
coefficients for either laminar or the
turbulent flow decrease with increasing
Reynolds’s number since the velocity
gradient decreases as the boundary
layer thickens.

28. In order to obtain low drag


sections, the transition from laminar to
Fig 4-8: Skin Friction Drag
turbulent must be delayed so that a
greater portion of the surface will be influenced by the laminar boundary layer. The conventional, low
speed aerofoil shapes are characterized by minimum pressure points very close to the leading edge.
Since high local velocities enhance early transition, very little surface is covered by the laminar
boundary layer. A comparison of two 9 percent thick symmetrical aerofoils and their drag
characteristics are presented in Fig 4-8. One section is the conventional NACA 0009 section which
has a minimum pressure point at approximately 10 percent chord at zero lift. The other section is the
FIS Book 1: Aerodynamics 66

NACA 66-009 which has a minimum pressure point at approximately 60 percent chord at zero lift. The
lower local velocities at the leading edge and the favorable pressure gradient of the NACA 66-009
delay the transition to some point further aft on the chord. The subsequent reduction in friction drag at
low angles of attack is called “Drag Bucket”.

Scale Effect

29. Variation of aerodynamic characteristics with RN is called Scale Effect and is a very important
consideration in wind tunnel testing. The two most important section characteristics affected by scale
effects are drag and maximum lift – the effect on pitching moments usually being negligible. All
experiments are carried out with models. Models have three drawbacks. These are:

(a) The smaller the model, the more difficult it is to make it accurate.

(b) Larger models would be constrained by wind tunnel dimensions- larger wind tunnels
would be required.

(c) Consider a 1/10th Scale model. All linear dimensions are 1/10th, but the areas are
1/100th and the mass (assuming same material for construction) is 1/1000th. So the model is
to scale in some respects but not in others. This is one of the difficulties in trying to learn from
flying models of aircraft, and unless the adjustment of weights is very carefully handled, the
results of tests in manoeuvre and spinning may be completely false.

30. Geometric and Dynamic Similarity. If we consider two bodies of identical shape but
different sizes placed at the same attitude in a uniform stream (possibly of different fluids, under
different conditions and at different speeds), then we say that the two flows are geometrically similar.
If however the flows, though incompressible, are not inviscid, the presence of a boundary layer must
be taken into account. Even if the flows are geometrically similar, the boundary layer may be laminar
in one case and largely turbulent in the other, so that the drag coefficients will be different. The way in
which RN influences aerodynamic characteristics is, as mentioned before, referred to as scale effect,
since scale largely determines the value of RN. If however, transition takes place at corresponding
points on the two bodies (say one-third of the way back from the leading edge) then the two flows will
be dynamically similar. Assuming that the transition RN is the same in both the cases, this implies that
the overall RN was also the same. Thus the condition for dynamic similarity is geometric similarity
plus identity of RN. In this case the aerodynamic force coefficients are the same for the two bodies
and there is no scale effect.

31. If the flow patterns over a full scale aircraft are not the same as being produced by the model,
then fundamental laws of aerodynamics (L= CL ½ ρ V 2 S and D = CD ½ ρ V 2 S) will not hold good. In
order to make these laws hold good, certain conditions need to be applied. These conditions are
based on Reynolds’s experiments, i.e. product of velocity (V) and linear dimension (L) must remain
67 Viscosity and Boundary Layer

constant in order to produce a similarity of flow. This is also referred to as the VL law. If it were
necessary to carry out a test on a 1/10th scale model to determine what would happen on the full size
aircraft at 400 kmph, the wind-tunnel speed would have to be 4,000 kmph. Furthermore, the wind-
tunnel would have to be very large, to prevent interference between the tunnel walls and the flow over
the model, especially at such high Mach numbers. In addition, the area of the model would be 1/100th
that of the aircraft and the wing of the model would have to support forces equal to those on a full size
aircraft.

32. Fortunately, Reynold also discovered that if different fluids were used, the type of flow was
affected by the density and viscosity of the fluid. In fact, he established that similarity of flow pattern
would be achieved if the value of density X velocity X size was constant.
viscosity

33. Clearly, it is not very practical to fill a wind-tunnel with oil or water, and accelerate it to 300 or
400 kmph to simulate higher flight speeds, but it is quite possible to increase the density of the air by
using a high pressure tunnel. Increasing the pressure has little or no effect upon the viscosity, so that
with increased density it is possible to reduce the velocity and/or size and still maintain aerodynamic
similarity. For example, if the air is compressed to 25 atmospheres - and this is possible - then the
density factor is 25, with no corresponding increase in viscosity (µ). It is then possible to test the
model referred to in para 31, at

4000 kmph i.e. at 80 kmph


25

34. Aerodynamic Forces. Since one of the major problems in using models was the effect of
the large aerodynamic force felt on a small model, it is useful to look at the effect of using high
pressure tunnels. For example, taking again a 1/10th model, a speed of 160 kmph and 25
atmospheres, to test for a full-scale flight at 400 kmph, the forces will be:

__1__ (Scale factor) X _(160)2_ (Speed factor) X 25 (Density factor)


100 (400)2
__1__ X _25600_ X 25 = _1_
100 160000 25

Thus dynamic similarity is achieved, with forces that are 1/25th of the full scale forces and even that is
quite high. For every wind-tunnel test there is one Reynolds Number (RN), and it is always published
with the results of the test.
FIS Book 1: Aerodynamics 68
69

CHAPTER 5

AEROFOILS

General

1. A thin flat plate is unsatisfactory as an aeroplane wing for two reasons. Firstly, it is necessary
from structural considerations to have a certain depth of the wing and secondly, the plate is far from
being the most efficient type of lifting surface. By curving the plate it is possible to increase the lift,
while by increasing its thickness and giving it a streamlined shape, it is possible to reduce its drag.
This is roughly the manner in which an aerofoil is designed. The amount of curving and thickness is
determined by the particular requirement of the aeroplane for which it is intended.

2. Aerofoil Terminology. Since the shape of an aerofoil and the inclination to the air stream
are so important in determining the pressure distribution, it is necessary to properly define the aerofoil
terminology. Fig 5-1 shows a typical aerofoil and illustrates the various items of aerofoil terminology.

(a) Chord line is a straight line connecting the leading and trailing edges of the aerofoil.

(b) Chord length is the characteristic dimension of the aerofoil.

(c) The mean camber line is a line drawn halfway between the upper and lower surfaces.
Actually, the chord line connects the ends of the mean-camber line.

Fig 5-1: Typical Aerofoil Terminology

(d) The shape of the mean-camber line is very important in determining the aerodynamic
characteristics of an aerofoil section. Maximum camber (displacement of the mean line from
the chord line) and the location of the maximum camber help to define the shape of the
FIS Book 1: Aerodynamics 70

mean-camber line. These quantities are expressed as fractions or percent of the basic chord
dimension. A typical low speed aerofoil may have a maximum camber of 4 percent, located
40 percent aft of the leading edge.

(e) The maximum thickness and thickness distribution of the profile are important
properties of a section. The maximum thickness and location of maximum thickness are
expressed as fractions or percent of the chord. A typical low speed aerofoil may have a
maximum thickness of 12 percent, located 30 percent aft of the leading edge.

(f) The leading edge radius of an aerofoil is the radius of curvature giving the leading
edge shape. It is the radius of the circle centred on a line tangent to the leading edge camber
and connecting tangency points of upper and lower surfaces with the leading edge. Typical
leading edge radii are zero (knife edge) to 1 or 2 percent.

(g) The lift produced by


an aerofoil is the net force
produced perpendicular to
the relative airflow (Fig 5-2).

(h) The drag incurred by


an aerofoil is the net force
produced parallel to the
relative airflow. Fig 5-2: Aerodynamic Forces

(j) The angle of attack is the angle between the chord line and the relative airflow. Angle
of attack is given the shorthand notation (alpha). It is important to differentiate between pitch
attitude angle and angle of attack. Regardless of the condition of flight, the instantaneous flight
path of the surface determines the direction of the oncoming relative airflow and the angle of
attack is the angle between the instantaneous relative airflow and the chord line. To
understand the definition of angle of attack, visualize the flight path of the aircraft during a loop
and appreciate that the relative airflow is defined by the flight path at any point during the
manoeuvre.

Type of Aerofoils

3. The performance of an aerofoil is governed by its contour. Generally, aerofoils can be divided
into three classes:

(a) High lift aerofoils.


(b) General purpose conventional aerofoils.
(c) High speed aerofoils.
71 Aerofoils

Typical examples of each are illustrated in Fig 5-3.

Fig 5-3: Aerofoil Section

4. High Lift Aerofoils. A typical high lift aerofoil section is shown in Fig 5-3(A) and its general
properties are:

(a) High lift sections employ a high t/c ratio, a pronounced camber, and a well-rounded
leading edge. Their maximum thickness is at about 25% - 30% of the chord aft of the leading
edge.

(b) The greater the camber, i.e., the amount of curvature of the mean camber line, the
greater the shift of centre of pressure for a given change in the angle of attack. The range of
movement of the CP is therefore large on a high lift section. This movement can be greatly
decreased by reflexing the trailing edge of the wing upwards, but some lift is lost as a result.

(c) Sections of this type are used mainly on gliders and other aircraft where a high CL is
all-important and speed is a secondary consideration.
FIS Book 1: Aerodynamics 72

5. General Purpose Conventional Aerofoils. Typical general-purpose aerofoil sections are


shown in Fig 5-3(B) and their general properties are:

(a) General purpose sections employ a lower t/c ratio, less camber and sharper leading
edge than those of the high lift type, but their maximum thickness is still at about 25%-30% of
the chord aft of the leading edge. The lower t/c ratio results in less drag and a lower CL than
those of a high lift aerofoil.

(b) Sections of this type are used on aircraft whose duties require speeds which, although
higher than those mentioned in the previous type, are not high enough to subject the aerofoil to
the effects of compressibility.

6. High Speed Aerofoils. Typical high-speed aerofoil sections are shown in Fig 5-3(C) and
their general properties are:

(a) High speed sections employ a very low t/c ratio, low camber and a sharper leading
edge. Their maximum thickness is at about the 50% chord point.

(b) Most of these sections lie in the 5% -10% t / c ratio band, but even thinner sections
have been used on research aircraft. The reason for this is the overriding requirement for low
drag. Naturally, the thinner sections have low maximum lift co-efficient.

(c) High speed aerofoils are usually symmetrical about the chord line. Some sections are
wedge shaped whilst others consist of arcs of a circle placed symmetrically about the chord
line, called the symmetrical bi-convex aerofoil. The behaviour and aerodynamics of these
sections at supersonic speeds are dealt with in detail in the chapter on Design for High-Speed
Flight.

7. Laminar Flow Aerofoils. There is another


type of aerofoil in common use on modern
aeroplanes. It is a fairly recent development and is
known as the laminar flow aerofoil. Laminar flow
aerofoils were originally developed for the purpose
of making an aeroplane fly faster. The laminar flow
wing is usually thinner than the conventional wing,
the leading edge is more pointed and its upper and
lower surfaces are nearly symmetrical. The major
and most important difference between the two Fig 5-4: Laminar Aerofoil
types of aerofoil is that the thickest part of a laminar wing occurs at 50% chord while in the
conventional design the thickest part is at 25% chord.
73 Aerofoils

8. The effect achieved by this design of the wing is to maintain the laminar flow of air through a
greater percentage of the chord of the wing and to control the transition point. Drag is, therefore,
considerably reduced since the laminar aerofoil takes less energy to slide through the air. The pressure
distribution on the laminar flow wing is much more even since the camber of the wing from the leading
edge to the point of maximum camber is more gradual.

9. Performance. The performance of all aerofoils is sensitive to small changes in contour.


Increasing or decreasing the thickness by as little as 1% of the chord, or moving the point of maximum
camber an inch or so in either direction, will alter all the characteristics. In particular, changes in the
shape of the leading edge have a marked effect on the maximum lift and drag obtained, and the
behaviour at stall. A sharp leading edge stalls more readily than one that is well rounded. Also it is
important that the finish of wing surfaces be carefully preserved if the aircraft is expected to attain its
maximum performance. Any dents or scratches in the surface bring about deterioration in the general
performance. These points are of particular importance on high performance aircraft where a poor
finish can result in a drastic reduction not only in performance, but also in control at high Mach
numbers.

Pressure Distribution around an Aerofoil

10. When air flows past an aerofoil, there are changes in velocity on the upper and lower surfaces
and associated pressure changes. The static pressure (absolute pressure) distribution around an
aerofoil may be measured with a manometer. This consists of a series of glass tubes mounted on the
aerofoil surface and all connected to a reservoir containing coloured fluid. As air flows over the aerofoil,
variation in absolute pressure is indicated by differing levels of fluids in the glass tubes.

11. Although most low speed aerofoils are similar in shape, each section is intended to give
certain specific aerodynamic characteristics. Therefore, there can be no such thing as a typical
aerofoil section or a typical aerofoil pressure distribution and it is only possible to discuss pressure
distributions around aerofoils in the broadest of general terms. So, in general, at conventional angles
of attack, compared with the free stream static pressure, there is a pressure decrease over much of
the upper surface, a lesser decrease over much of the lower surface so that the greatest contribution
to overall lift comes from the upper surface.

12. The aerofoil profile presented to the airflow determines the distribution of velocity and hence
the distribution of pressure over the surface. This profile is determined by the aerofoil geometry, i.e.
thickness distribution and camber, and by the angle of attack. The greatest positive pressures occur
at stagnation points where the flow is brought to rest, at the trailing edge, and somewhere near the
leading edge, depending on the angle of attack. At the front stagnation point the flow divides to pass
over and under the section. At this point there must be some initial acceleration of the flow at the
surface otherwise there could be no real velocity anywhere at the aerofoil surface, therefore there
FIS Book 1: Aerodynamics 74

must be some initial reduction of pressure below the stagnation value. If the profile is such as to
produce a continuous acceleration there will be a continuous pressure reduction and vice versa.
Some parts of the contour will produce the first effect, other parts the latter, bearing in mind always
that a smooth contour will produce a smoothly changing pressure distribution which must finish with
the stagnation value at the trailing edge.

13. Fig 5-5 shows the pressure distribution


around a particular aerofoil section at varying
angles of attack. The flow over the section
accelerates rapidly around the nose and over
the leading portion of the surface, the rate of
acceleration increasing with increase in angle of
attack. The pressure reduces continuously from
the stagnation value through the free stream
value to a position when a peak negative value
is reached. From there onwards the flow is
continuously retarded, increasing the pressure
through the free stream value to a small positive
value towards the trailing edge. The flow under
the section is accelerated much less rapidly than
that over the section, reducing the pressure
much more slowly through the free stream value
to some small negative value, with subsequent
deceleration and increase in pressure through
free stream value to a small positive value
toward the trailing edge. If the slight concavity
on the lower surface towards the trailing edge
was carried a little further forward, it might be
possible to sustain a positive pressure over the
whole of the lower surface at the higher angles
of attack (reflexed cambered aerofoil).
However, although this would increase the lifting
Fig 5-5: Pressure Distribution
properties of the section, it might also produce
Around an Aerofoil
undesirable changes in the drag and pitching
characteristics. Therefore, it can be seen that any pressure distribution around an aerofoil must
clearly take account of the particular aerofoil contour.

14. From the typical pressure distribution around an aerofoil (Fig 5-5), the following points can be
noted:
75 Aerofoils

(a) At a small negative angle of attack (-40), for this aerofoil, the decrease in pressure
above and below the section would be equal and the section would give no lift.

(b) At higher angles of attack (+80), lift is partly due to the decreased pressure on the
upper surface and partly due to increased pressure on the lower surface.

(c) At small angles of attack (+40), the lift arises from a difference between the pressure
reduction on the upper and the lower surfaces.

(d) At positive angles of attack (+140), the increase in pressure over the lower surface
remains almost constant. The greatest contribution to lift comes from the increasing reduction
of pressure on the upper surface as the AOA is increased.

(e) Beyond the stalling AOA (+16°), there is a sudden flattening of the pressure envelope
above the upper surface. This causes a drastic loss of lift and is referred to as stall. Whatever
lift remains, does so primarily due to the increased pressure on the lower surface.

(f) The zero lift angle of attack for a positively cambered aerofoil is at about –4o. For a
symmetrical aerofoil, the zero lift AOA would be zero degrees and for a reflexed camber
aerofoil it will be some positive value.

CENTRE OF PRESSURE (CP)

15. The sum of all aerodynamic forces on an aerofoil may be represented by a single vector
acting at a particular point on the chord line, called the centre of pressure. This vector is referred to
as the Total Reaction. The TR is always perpendicular to the aerofoil chord and may be divided into
Lift and Drag (Fig 5-6). It can also be seen from Fig 5-5 that as the angle of attack is increased, the
height and the size of the upper
surface suction peak increases and it
moves further forward. This indicates
that the total reaction increases in
magnitude and the centre of pressure
(through which the TR seems to act)
moves forward towards the leading
edge with increasing angle of attack
up to the point of stall. Beyond stalling
angle of attack, the magnitude of the
vector reduces rapidly and it
simultaneously moves backwards
Fig 5-6: Aerodynamic Forces on Aerofoil
towards the trailing edge.
FIS Book 1: Aerodynamics 76

16. Movement of Centre of Pressure.


The location of the CP is a function of camber
and section lift co-efficient, both the resultant
force and its position varying with angle of
attack (Fig 5-7). As the angle of attack is
increased the magnitude of the force increases
and the CP moves forward. When the stall is
reached the force decreases abruptly and the
CP generally moves back along the chord. With
a cambered aerofoil, the CP movement over the
normal working range of angles of attack is
between 20% - 30% of the chord aft of the
leading edge. With a symmetrical aerofoil, there
is virtually no CP movement over the working Fig 5-7: CP Movement
range of angles of attack at subsonic speeds.

17. CP Movement on a Flat Plate. On a flat plate the CP lies well behind the leading edge and
any increase in angle of attack tends to move the CP further back. This creates a nose-down pitching
tendency with increasing angle of attack on a flat plate. Thus, CP movement on a flat plate in subsonic
flow is stable.

18. On a cambered aerofoil, on the other hand, the CP moves forward with increasing AOA, further
accentuating the ‘nose up’ pitching tendency. Thus, CP movement on an aerofoil (up to stalling AOA) is
unstable.

19. Reduction in CP Movement. As movement of CP depends upon camber, any reduction in


camber will reduce movement of CP. Two common methods of reducing camber and thereby reducing
CP movement are:

(a) By providing convexity to the under surface of the aerofoil.


(b) By reflexing the trailing edge of the aerofoil upwards. This, however has the
disadvantage of reducing lift and increasing drag.

AERODYNAMIC CENTRE

20. An aircraft pitches about the lateral axis which passes through the centre of gravity. The wing
pitching moment is the product of lift and the distance between the CG and the CP of the wing.
Unfortunately, the CP moves when the angle of attack is altered so calculation of the pitching moment
becomes complicated.
77 Aerofoils

21. Pitching Movement Characteristics. In


Fig 5-8, we consider moments about a point A which
is ahead of an aerofoil and another point B which is
behind the aerofoil. At large negative angles of attack
the value of CL is negative as evidenced from the CL
Vs. α graph. For negative values of lift, the pitching
moment about point A will be nose up and for point B
will be nose down. As α is increased, the value of
negative lift reduces and so will the values of pitching
moments about points A (nose up) and B (nose Fig 5-8: Pitching Moment
down). As the α is increased beyond the zero lift
angle of attack, the aerofoil starts producing positive
lift and the pitching moment about point A changes to
nose down while that about point B changes to nose
up. As the angle of attack is increased further, the lift
force increases and the CP also moves forward.
Despite the reduced moment arm about point A, the
nose down pitching moment increases because the
rate of increase of lift force is greater than the
reduction in the moment arm. The pitching moments
about point B will increase likewise, the effect being Fig 5-9: Pitching Moments Vs. AOA
accentuated because both the moment arm and the
lift component are increasing (Fig 5-9).

22. The pitching moment of a wing can be measured experimentally by direct measurement on a
balance or by pressure plotting. The pitching moment coefficient (CM) is calculated as follows:

CM = Pitching Moment Where: c = Mean Aerodynamic Chord.


q.S.c q = Dynamic Pressure.
S = Wing Area.

The value of pitching moment and therefore CM will depend upon the point about which the moment is
calculated and will vary because the lift force and the position of the CP changes with angle of attack.

23. If pitching moments are measured at various points along the chord for several values of CL
one particular point is found where the CM is constant. This point occurs where the change in CL with
angle of attack is offset by the change in distance between CP and CG. This is the aerodynamic
centre (AC). For a flat or curved plate in inviscid, incompressible flow the aerodynamic centre is at
approximately 25% of the chord from the leading edge. Thickness of the section and viscosity tend to
move it forward and compressibility moves it rearwards. Some modern low drag aerofoils have the
AC a little further forward at approximately 23% chord.
FIS Book 1: Aerodynamics 78

24. From the property of constant


moment at the AC, whatever the angle of
attack, it may be argued that the lift force
changes resulting from angle of attack
changes may be assumed to act at the AC. In
other words, provided we always insert the
zero lift moment Mo, we may assume that the
Fig 5-10: Two Ways of Presenting the
lift force acts, not at the centre of pressure (a Lift Force
varying point) but at the aerodynamic centre,
a fixed point. The two ways of presenting the
lift force and pitching moment shown in Fig 5-
10 are equivalent. Sometimes the
convenience of having the lift force anchored
to a fixed point outweighs the slight
complication of inserting a zero lift moment
MO.

25. The rate of change of CM with respect


to CL or angle of attack is constant through
most of the angle of attack range and the
value of CM / CL depends on the point on the
aerofoil at which CM is measured. Curves of Fig 5-11: CM Against CL
CM versus CL are shown in Fig 5-11. It can be
seen that a residual pitching moment is
present at zero lift. This is because an aerofoil
with positive camber has a distribution of
pressure as illustrated in Fig 5-12. It should
be noted that the pressure on the upper
surface towards the leading edge is higher
than ambient and towards the trailing edge
the pressure is lower than ambient. This
results in a nose down (negative) pitching
moment even though there is no net lift at this
angle of attack. The CM at zero lift angle of
attack is called CMo and, since the pitching
moment about the AC is constant with CL by
definition, its value is equal to CMo. The value
Fig 5-12: Pressure Distribution at –ve AOA
of CM is determined by camber and is usually
negative but is zero for a symmetrical aerofoil and can be positive when there is reflex curvature at
the trailing edge. When CM is measured at point A (Fig 5-10) an increase in CL will cause an increase
79 Aerofoils

in lift and a larger negative (nose down) moment. When measured at point B an increase in CL will
cause the moment to become less negative and eventually positive (nose up).

26. Approaching CL max the CM / CL graph


departs from the straight line (Fig 5-13).
The CL decreases and the CP moves aft.
If at this stage the CM becomes more
negative it tends to unstall the wing and is
stable. If the CM becomes less negative
(positive) the pitch up aggravates the stall
and is unstable. This is known as pitch-
up and is associated with highly swept
wings. At stall, the aerodynamic centre
Fig 5-13: CM in the Region of CL MAX
will also move back from the ¼ chord
position. The AC also moves back to 50% when the aircraft goes supersonic. This causes an increase
in trim drag.

DESIGN AND NOMENCLATURE OF AEROFOIL SECTIONS

27. In the early days, in fact until late 1930s, very few aerofoil shapes were suggested by theory.
The usual method was to sketch out a shape by hand, give it a thorough test and then try to improve
on it by slight modifications. As a result, we had a mass of experimental data obtained under varying
conditions in various wind tunnels. The results were interpreted in different ways, and several systems
of units and symbols were used, so it was difficult to make use of the data available.

28. It is true that this hit-and-miss method of aerofoil design produced a few excellent sections but
it was gradually replaced by more systematic methods. The first step in this direction was to design and
test a 'family' of aerofoils by taking a standard symmetrical section and altering the curvature, or
camber, of its centre line. An early example of this system in the UK was the series beginning with RAF
30, a symmetrical section from which RAF 31, 32, 33, etc. were evolved by curving the centre line in
various ways and according to a definite plan. Then RAF 40 was used as the basic section of a new
family, and so on. RAF, in ‘RAF 40’, referred to the Royal Aircraft Factory, now the Defence Research
Agency (DRA). In Germany, similar investigations were made under a series named after the
‘Gottingen Laboratory’, and in America with the ‘Clark Y’ series. Later sections in these series have
been based on theoretical calculations and wind tunnel testing.

29. The naming and numbering of sections has also been rather haphazard. At first the actual
number, such as RAF 15, meant nothing except perhaps that it was the 15th section to be tried. But the
National Advisory Committee for Aeronautics in USA soon attempted to devise a system whereby the
letters and numbers denoting the aerofoil section served as a guide to its main features. This meant
FIS Book 1: Aerodynamics 80

that we could get a good idea of what the section was like simply from its number. Unfortunately the
system has been changed from time to time, and this has caused confusion over the years. However
since NACA sections (or slight modifications of them), are now used by nearly every country in the
world, we may be interested in getting some idea of its nomenclature.

30. The geometric features that have most effect on the qualities of an aerofoil section are:

(a) The camber of the centre line.


(b) The position of maximum camber.
(c) The maximum thickness, and variation of thickness along the chord.
(d) The radius of curvature of the leading edge.
(e) Whether the centre line is straight, or reflexed near the trailing edge; and the angle
between the upper and lower surfaces at the trailing edge.

31. The NACA sections designed for comparatively low speed aircraft are based on either the four
or five digit system. Laminar flow sections for high subsonic speeds on the 6, 7 or 8 systems (the 6, 7
or 8 being the first figure and not the number of digits). In each system there are complicated formulae
for the thickness distribution, the radius of the leading edge and the shape of the centre line, but we
need not worry about these. What is easier to understand is the meaning of the digits or integers. For
instance, in the four-digit system:

(a) The first digit gives the maximum camber as a percentage of the chord.
(b) The second digit gives the position of the maximum camber, i.e. distance from the
leading edge, in tenths of the chord.
(c) The third and, fourth digits indicate the maximum thickness as a percentage of the
chord.

32. Thus NACA 4412 has a maximum camber of 4 per cent of the chord, the position of this
maximum camber is 40 per cent aft of the chord leading edge, and the maximum thickness is 12 per
cent of the chord. In a symmetrical section there is no camber, so the first two digits will be zero. Thus
NACA 0009 is a symmetrical section of 9 per cent thickness.

33. Notice that these are all geometric features of the section, but in later systems attempts are
made to indicate some of the aerodynamic characteristics also. For instance, in the five-digit system

(a) The 'design lift coefficient’ is three-halves divided by 10, of the first digit.
(b) The second and third digits together indicate twice the distance back of the maximum
camber, as a percentage of the chord.
(c) The last two, once again, indicate the maximum thickness as a percentage of the
chord.
81 Aerofoils

34. The 'design lift coefficient' is the lift coefficient at the angle of attack for normal level flight,
usually at about 2° or 3°. Most of these sections have a 2 per cent camber, and there is a relationship
between the design lift coefficient and the maximum camber which has sometimes led to confusion
about the meaning of the first digit. The point of maximum camber is at 15 per cent, 20 per cent or 25
per cent of the chord (which accounts for the doubling of the second and third digits to 30, 40 or 50
accordingly). In fact the most successful, and so the most common of these sections, begins with the
digits 230, followed by the last two indicating the thickness. Thus NACA 23012, as used on the Britten-
Norman Islander, has a design lift coefficient of 0.3 (it also has 2 per cent camber), the maximum
camber is at 15 per cent of the chord, while the maximum thickness is 12 per cent.

35. The forward position of the maximum camber in the five-digit sections results in low drag and
but poor stalling characteristics, which explains why, when these sections are used near the root of a
wing, they are often changed to a four-digit one (which gives a smooth stall) near the tip.

36. It should be noted that the position of maximum thickness (not indicated in either of these
systems) is not necessarily the same as that of maximum camber, and in one British system eight
digits were used so that this too could be
illustrated. Two pairs of digits gave the
thickness and its position, two other pairs,
the maximum camber and its position. Fig
5-14 illustrates 1240/0658 based on this
system. For a symmetrical section, the
last four figures are omitted since they
Fig 5-14: NACA Section 1240 / 0658
would all be zero.

37. In the NACA 6, 7 and 8 series, as in nearly all the NACA series, the last two digits again
indicate the percentage thickness, but the other figures, letters, suffixes, dashes and brackets become
so complicated that it is necessary to refer to tables. Most of these sections are particularly good for
high subsonic speeds.
FIS Book 1: Aerodynamics 82
83

CHAPTER 6

LIFT

Introduction

1. A boy flying a kite can be considered to be a pilot in charge of an aeroplane! Ponder the idea
a moment and it may not appear quite as absurd as it seems at first glance. A kite is an inclined plane
(material surface inclined at an acute angle to horizon), the weight of which is supported in the air by
the reaction of the wind flowing against it. If we substitute for the string, which holds the kite against
the wind, the engine and propeller of an aeroplane, which move the wings forward against the airflow,
we will see that the analogy of the kite is not without some validity. The wings of an aeroplane are so
designed that when moved through the air horizontally, the force exerted on them produces a reaction
as nearly vertical as possible. It is this reaction that lifts the weight of the aeroplane.

2. By definition lift is that component of total aerodynamic reaction which is perpendicular to the
flight path of the aircraft. The fact that air is invisible makes flight difficult to understand. When a ship
passes through water we can see the “bow
wave,” the wash behind, and all the eddies
and whirlpools which are formed, whereas
when an aeroplane makes its way through
the atmosphere nothing appears to
happen, yet in reality, even more
commotion has been caused than by the
ship (Fig 6-1). If only we could see this
commotion many of the phenomena of
flight would need much less explanation.
One way to see this commotion is by
introducing smoke into the air or by
watching the flow of water, which exhibits
many characteristics similar to those of the
Fig 6-1: Effect of Aircraft on Airstream
air (fluid= liquid or gas).

3. Another property of air which is apt to give us misleading ideas when we first begin to study
flight is its low density. The air feels thin, it is difficult for us to obtain any grip upon it, and if it has any
weight at all we usually consider it as negligible for all practical purposes. Ask anyone what is the
weight of air in any ordinary room and you will probably receive answers varying from “almost nothing”
up to “about ten kilograms.” Yet the real answer will be nearer 150 Kg, and in a large hall may be over
one ton! Again, most of us who have tried to dive into water have experienced the sensation of
coming down “flat” on to the surface of water. Since then we have treated water with respect,
FIS Book 1: Aerodynamics 84

realising that it has substance and that it can exert forces which have to be reckoned with (I bet you
can’t forget the 10 meter board). We have probably had no such experience with air. It is true that the
density of air is low compared with water (at ground level the weight of a cubic meter of water weighs
nearly 800 times air), yet it is this very property of air, its density, which makes all airborne flight
possible (this does not apply to missiles and satellites). All airborne objects are supported in the air by
forces which are entirely dependent on the density of air. The lesser the density, the more difficult
does flight become, and for all of them flight becomes impossible in a vacuum. So let us next realise
the fact that, however thin the air may seem to be, it possesses, like other substances, the properties
of weight, of mass and of density.

4. It will now be easy to understand that air must possess, in common with other substances,
the property of inertia and the tendency to obey the laws of mechanics. In the earlier chapters we
discussed various aerodynamic theories associated with the production of lift. Let us now examine
these theories with reference to Newton’s laws of motion. According to Newton’s First Law, when in
motion the air tends to remain in motion. The introduction of an aerofoil into the streamlined flow alters
the uniform flow of air. Newton’s Second law states that a force must be applied to alter the state of
uniform motion of a body. The aerofoil is the force that acts on the body (in this case, the air) to
produce a change of direction. The application of such a force causes an equal and opposite reaction
(Newton’s Third Law) called, in this case, lift. It is this force that supports the weight of the aircraft in
air. Thus it can be said that lift is an aerodynamic force (force caused due to relative motion of air)
which is perpendicular to the flight path of the aircraft. One might be tempted to ask at this stage that
if lift is responsible for keeping the aircraft airborne then what does the thrust do? It would be sufficient
to know at this stage that it is thrust which provides the desired relative motion between the air and
the aircraft at which lifting force becomes powerful enough to lift the aircraft in air.

5. It can be demonstrated experimentally that the total aerodynamic reaction, and therefore the
lift acting on a wing moving through air, is dependent upon at least the following variables:

(a) Free stream velocity (V²).


(b) Air density (ρ).
(c) Wing area (S).
(d) Wing shape in section and in planform.
(e) Angle of attack (α).
(f) Condition of the surface.
(g) Viscosity of the air (µ).
(h) The speed of sound “a”, i.e. the speed of propagation of small pressure waves.

6. In the chapter on Theories of Aerodynamic Forces and the one on the Aerofoils, it was seen
that lift increased when the angle of attack of a given aerofoil section was increased, and that the
increase in lift was achieved mechanically by greater acceleration of the airflow over the section, with
85 Lift

an appropriate decrease in pressure. The general and simplified equation for aerodynamic force is
½ ρ V² S X a coefficient, and the coefficient indicates the change in force which occurs when the
angle of attack is altered.

7. The equation for lift is CL½ρV²S, and CL for a given aerofoil section and planform allows for
angle of attack and all the unknown quantities which are not represented in the force formula. Three
proofs for the lift equation are given in the chapter of Theories of Aerodynamic Forces.

Coefficient of Lift (CL)

8. The coefficient of lift is obtained experimentally at a quoted RN from the equation:

L = CL ½ ρ V2 S
CL = ____L___
½ ρ V2 S
CL = (L/S)__
½ ρ V2

These values of CL are usually plotted against angle of attack and it is then possible to consider the
factors affecting lift in terms of CL and to show these effects on the CL curve.

Factors Affecting CL

9. The coefficient of lift is dependent upon the following factors:

(a) Angle of attack.


(b) Shape of the wing section and planform.
(c) Condition of the wing surface.
(d) Reynolds Number.
(e) Speed of sound (Mach number).

10. Angle of Attack. A typical lift curve is


shown in Fig 6-2 for a wing of 13%
thickness/chord (t/c) ratio and 2% camber (the
zero lift angle of attack (αL0), is negative, its
magnitude is often roughly equal, in degrees to
the percentage camber, e.g. an aerofoil with 2%
camber will have αL0 ≈ -2º (For a symmetrical
aerofoil, αL0 = 0). The greater part of the curve is
linear and the airflow follows the design contour
Fig 6-2: Typical Lift Curve for a Moderately
of the aerofoil almost to the trailing edge before Thick, Cambered Section
FIS Book 1: Aerodynamics 86

separation. At higher angles of attack the curve begins to lean over slightly, indicating a loss of lifting
effectiveness. From the point of maximum thickness to the trailing edge of the aerofoil, the flow
outside the boundary layer is decelerating, accompanied by a pressure rise (Bernoulli’s theorem).
This adverse pressure gradient thickens the existing boundary layer. In the boundary layer, the
airflow’s kinetic energy has been reduced by friction, the energy loss appearing as heat. The
weakened flow, encountering the thickened layer, slows still further. With increasing angle of attack,
the boundary layer separation point (explained in the chapter on drag) moves rapidly forward, the
detached flow causing a substantial reduction of CL. The aerofoil may be considered to have changed
from a streamlined body to a bluff one, with the separation point moving rapidly forwards from the
region of the trailing edge. The desirable progressive stall of an actual wing is achieved by washout at
the tips or change of aerofoil section along the span, or a combination of both.

11. Effect of Shape. Changes in the shape of a wing may be considered under the following
headings:

(a) Leading Edge Radius. The


shape of the leading edge, and the
condition of its surface largely
determines the stalling characteristics
of a wing. In general, a blunt leading
edge with a large radius will result in a
well-rounded peak of the CL curve. A
small radius, on the other hand,
invariably produces an abrupt stall
(Fig 6-3) but this may be modified
considerably by surface roughness
which is discussed later. Fig 6-3: Effect of Leading Edge Radius

(b) Camber. The effect of


camber is illustrated in Fig 6-4. Line
(a) represents the curve for a
symmetrical section. Lines (b) and (c)
are for sections of increasing camber.
A symmetrical aerofoil at zero angle of
attack will have the same pressure
distribution on its upper and lower
surfaces, therefore it will not produce
lift. As the angle of attack is increased,
the stagnation point moves from the
Fig 6-4: Effect of Camber on Lift Curve
chord line to a point below, moving
87 Lift

slightly further backwards with increase in


angle of attack. This effectively lengthens
the path of the flow over the top surface and
reduces it on the lower, thus changing the
symmetrical section (Fig 6-5 upper figure) to
an apparent cambered one as in Fig 6-5
(lower figure). A positively cambered wing
will produce lift at zero angle of attack
because the airflow attains a higher velocity
over the upper surface creating a pressure
differential and lift. This gives it a lead over
Fig 6-5: Stagnation Point on the
the symmetrical section at all normal angles Symmetrical Aerofoil
of attack but pays the penalty of an earlier
stalling angle as shown by the CL versus α curve which shifts up and left (Fig 6-4) as the
camber is increased. The angle of attack at which the CL is zero is known as the zero-lift
angle of attack (αL0) and a typical value is -3° for a cambered section.

(c) Aspect Ratio. Fig 6-6


shows the downward component of
airflow at the rear of the wing,
caused by trailing edge vortices and
known as induced downwash (ω).
The induced downwash causes the
flow over the wing to be inclined
slightly downwards from the
direction of the undisturbed stream
(V) by the angle α1. This reduces
the effective angle of attack, which
determines the airflow and the lift
and drag forces acting on the wing.
The effect on the CL by change of
aspect ratio (AR) will depend on
how the effective angle of attack is
influenced by change in AR. It has
been shown in the chapter on
theories of aerodynamic forces that
a wing of infinite span has no Fig 6-6: Effect of Aspect Ratio on Induced
Downwash and on the Lift Curve
induced downwash. It can be
demonstrated that the nearer one gets to that ideal, i.e. high AR, the less effect the vortices
will have on the relative airflow along the semispan and therefore the least deviation from the
FIS Book 1: Aerodynamics 88

shape of the CL curve of the wing with infinite AR. It can be seen from Fig 6-6 that at any α,
apart from the αL0, the increase in CL with changes in α of the finite wing is lesser than the
infinite wing, the lag increasing with reducing AR due to increasing α1. Theoretically the CL
peak values should not be affected, but experimental results show a slight reduction of CL max
as the aspect ratio is lowered.

(d) Sweepback. If an aircraft’s wings are swept and the wing area remains the same,
then by definition the aspect ratio (span2/area) must be less than the AR of the equivalent
straight wing. The shape of the CL vs. angle of attack curve for a swept wing, compared to a
straight wing, is similar to the comparison between a low and a high aspect ratio wing.
However, this does not explain the marked reduction in CL max at sweep angles in excess of
40-45°, which is mainly due to earlier flow separation from the upper surface. An alternative
explanation is to resolve the airflow over a swept wing into two components. The component
parallel to the leading edge produces no lift. Only the component normal to the leading edge
is considered to be producing lift. As this
component is always less than the free
stream flow at all angles of sweep, a
swept wing will always produce less lift
than a straight wing (Fig 6-7). For e.g.
the MiG 27 aircraft has a CL max of
approximately 1.25 with its wings swept
forward (Sweep 160) while the critical
angle of attack is 230 to 260. The figures
when its wings are swept back (Sweep
720) are the CL max of approximately 1.1
to 1.15 and the stalling angle of attack of
Fig 6- 7: Effect of Sweepback on Lift
320 to 340. Coefficient

12. Effect of Surface Condition. It has


long been known that surface roughness,
especially near the leading edge, has a
considerable effect on the characteristics of wing
sections. The maximum lift coefficient, in
particular, is sensitive to the leading edge
roughness. Fig 6-8 illustrates the effect of a
roughened leading edge compared to a smooth
surface. In general, the maximum lift coefficient
decreases progressively with increasing
roughness of the leading edge. Roughness of
the surface further downstream than about 20%
chord from the leading edge has little effect on Fig 6- 8: Effect of Leading Edge Roughness
89 Lift

CL max or the lift-curve slope. The standard roughness illustrated is more severe than that caused by
usual manufacturing irregularities or deterioration in service, but is considerably less severe than that
likely to be encountered in service as a result of the accumulation of ice, mud or combat damage.
Under test, the leading edge of a model wing is artificially roughened by applying carborundum grains
to the surface over a length of 8% from the leading edge of both surfaces.

13. Effect of Reynolds Number. The formula for Reynolds Number is:

RN = ρ V L
µ

that is density X velocity X a mean chord length, divided by viscosity (A fuller explanation of Reynolds
Number is given in the chapter on Theories of Aerodynamic Forces). If we consider an aircraft
operating at a given altitude, L is constant, ρ is constant, and at a given temperature the viscosity is
constant. Therefore the only variable is V. For all practical purposes the graph in Fig 6-9 shows the
effect on CL of increasing velocity on a general purpose aerofoil section. It should be remembered
than an increase in Reynolds Number, for any reason, will produce the same effect. Fig 6-9 shows
that with increasing velocity both the maximum value of CL and the stalling angle of attack is
increased. An increase in the velocity of the
airflow over a wing will produce earlier
transition and an increase in the kinetic energy
of the turbulent boundary layer due to mixing,
the result is delayed separation. An increase
in density or a reduction in viscosity will have
the same effect on stall. The effect shown in
Fig 6-9 is generally least for thin sections (t/c
<12%) and greatest for thick, well-cambered
sections. Note that smaller changes in CL max

occur at higher RN 6 & 8.9 than at 2.6 & 6.


This would be further discussed in the chapter Fig 6-9: Effect of RN on CL Max
on Drag.

14. Effect of Mach number. The effect of Mach number is discussed later in the chapter on
Transonic and Supersonic Aerodynamics.
FIS Book 1: Aerodynamics 90
91

CHAPTER 7

DRAG

Introduction

1. Each part of an aircraft in flight produces an aerodynamic force. Total drag is the sum of all
the components of the aerodynamic forces, which act parallel and opposite to the direction of flight.
Each part of total drag represents a value of resistance of the aircraft’s movement, that is, lost energy.

Components of Total Drag

2. Some text books still break down Total Drag into the old terms Profile Drag and Induced
Drag. These latter terms are now more widely known as Zero Lift Drag and Lift Dependent Drag
respectively and are described later. There are three points to be borne in mind when considering
total drag:

(a) The causes of subsonic drag have changed very little over the years but the balance
of values has changed, e.g. parasite drag is such a small part of the whole that it is no longer
considered separately, except when describing helicopter power requirements.

(b) An aircraft in flight will have drag even when it is not producing lift.

(c) In producing lift the whole aircraft produces additional drag and some of this will be
increment in those components which make up zero lift drag.

3. Zero Lift Drag. When an aircraft is flying at zero lift angle of attack the resultant of all the
aerodynamic forces acts parallel and opposite to the direction of flight. This is known as Zero Lift
Drag (referred to as Profile Drag or Boundary Layer Drag in some textbooks) and is composed of:

(a) Surface friction drag.


(b) Form drag (boundary layer normal pressure drag).
(c) Interference drag.

4. Lift Dependent Drag. In producing lift the whole aircraft will produce additional drag
composed of:

(a) Induced drag (vortex drag).


(b) Increments of:
(i) Surface friction drag.
(ii) Form drag.
(iii) Interference drag.
FIS Book 1: Aerodynamics 92

ZERO LIFT DRAG

The Boundary Layer

5. Although it is convenient to ignore the effects of viscosity whenever possible, certain aspects
of aerodynamics cannot be explained if viscosity is disregarded.

6. Because air is viscous, any object moving through it collects a group of air particles which it
pulls along. A particle directly adjacent to the object’s surface will, because of viscous adhesion, be
pulled along at approximately the speed of the object. A particle slightly further away from the surface
will also be pulled along, however its velocity will be slightly less than the object’s velocity. As we
move further and further away from the surface, the particles of air are affected less and less, until a
point is reached where the movement of the body does not cause any parallel motion of air particles
whatsoever.

7. The layer of air extending from the surface to the point where no dragging effect is
discernable is known as the boundary layer. In flight, the nature of the boundary layer determines the
maximum lift coefficient, stalling characteristics of a wing, value of form drag, and to some extent the
high speed characteristics of an aircraft.

8. The viscous drag force within the boundary layer is not sensitive to pressure and density
variations and is therefore unaffected by the variations in pressures at right angles (normal) to the
surface of the body. The coefficient of viscosity of air is directly proportional to temperature and
therefore decreases with altitude.

Surface Friction Drag

9. The surface friction drag is determined by:

(a) The total surface area of the aircraft.


(b) The coefficient of viscosity of air.
(c) The rate of change of velocity across the flow.

10. Surface Area. The whole surface area of the aircraft has a boundary layer and therefore
has surface friction drag.

11. Coefficient of Viscosity. The absolute coefficient of viscosity (µ) is a direct measure of the
viscosity of a fluid, i.e. it is the means by which a value can be allotted to this property of a fluid. The
greater the viscosity of air, the greater the dragging effect on the aircraft’s surface.
93 Drag

12. Rate of Change of Velocity.


Consider the flow of air moving across a
thin flat plate, as in Fig 7-1. The
boundary layer is normally defined as
that region of flow in which the speed is
less than 99% of the free stream flow,
and usually exists in two forms, laminar
and turbulent. In general, the flow at the
front of a body is laminar and becomes
turbulent at a point some distance along
Fig 7-1: Boundary Layer
the surface, known as the transition
point. From Fig 7-1 it can be seen that the rate of change of velocity is greater at the surface in the
turbulent flow than in the laminar. This higher rate of change of velocity results in greater surface
friction drag. The velocity profile for the turbulent layer shows the effect of mixing with the faster
moving air above the boundary layer. This is an important characteristic of the turbulent flow since it
indicates a higher level of kinetic energy. Even when the boundary layer is turbulent, a very thin layer
exists immediately adjacent to the surface in which random velocities are smoothed out. This very
thin layer (perhaps 1% of the total thickness of the turbulent layer) remains in the laminar state and is
called the laminar sub-layer. Though extremely thin, the presence of this layer is important when
considering the surface friction drag of a body and the reduction that can be obtained in the drag of a
body as a result of smoothing the surface.

Transition to Turbulence

13. It follows from para 12 that forward movement of the transition point increases the surface
friction drag. The position of the transition point depends upon:

(a) Surface condition.


(b) Speed of the flow.
(c) Size of the object.
(d) Adverse pressure gradient.

14. Surface Condition. Both the laminar and turbulent boundary layers thicken downstream,
the average thickness varying from approximately 2 mm at a point 1 m downstream of the leading
edge, to 20 mm at a point 1 m downstream of the transition point. Generally, the turbulent layer is
about ten times thicker than the laminar layer but exact values vary from surface to surface. The thin
laminar layer is extremely sensitive to surface irregularities. Any roughness which can be felt by the
hand, on the skin of the aircraft, will cause transition to turbulence at that point, and the thickening
boundary layer will spread out fanwise down-stream causing a marked increase in surface friction
drag.
FIS Book 1: Aerodynamics 94

15. Speed and Size. The nineteenth century physicist, Reynolds, discovered that, in a fluid flow
of given density and viscosity, the flow changed from streamline to turbulent when the velocity
reached a value that was inversely proportional to the thickness of a body in the flow. That is, thicker
the body, lower is the speed at which transition occurred. Applied to an aerofoil of given thickness it
follows that an increase of flow velocity will cause the transition point to move forward towards the
leading edge. Earlier transition means that a greater part of the surface is covered by a turbulent
boundary layer, creating greater surface friction drag. However, it should be remembered that the
turbulent layer has greater kinetic energy than the laminar, the effect of which is to delay separation,
thereby increasing the maximum value of CL, as explained in the chapter on Lift.

16. Adverse Pressure Gradient. It


has been found that a laminar boundary
layer cannot be maintained, without
mechanical assistance, when the pressure
is rising in the direction of flow, i.e. in an
adverse pressure gradient. Thus on the
curved surfaces of an aircraft the transition
point is usually beneath, or near to, the
point of minimum pressure and this is
normally found to be at the point of
maximum thickness. Fig 7-2 illustrates the
comparison between a flat plate and a
curved surface. Fig 7-2: Effect of Adverse Pressure Gradient

Form Drag (Boundary Layer Normal Pressure Drag)

17. The difference between surface friction and form drag can be easily appreciated if a flat plate
is considered in two attitudes, first at zero angle of attack when all the drag is friction drag, and
second at 90° angle of attack when all the drag is form drag due to the separation.

18. Separation Point. The


effect of surface friction is to reduce
the velocity, and therefore the kinetic
energy of the air within the boundary
layer. On a curved surface the effect
of the adverse pressure gradient is to
reduce further the kinetic energy of the
boundary layer. Eventually, at a point
close to the trailing edge of the Fig 7-3: Boundary Layer Separation
surface, a finite amount of the
95 Drag

boundary layer stops moving, resulting in eddies within the turbulent wake. Fig 7-3 shows the
separation point and the flow reversal, which occurs behind that point. Aft of the transition point, the
faster moving air above mixes with the turbulent boundary layer and therefore has greater kinetic
energy than the laminar layer. The turbulent layer will now separate as readily under the influence of
the adverse pressure as would the laminar layer.

19. Streamlining. When a boundary layer separates some distance forward of the trailing
edge, the pressure existing at the separation point will be something less than at the forward
stagnation point. Each part of the aircraft will therefore be subject to a drag force due to the
differences in pressure between its fore and aft surface areas. This pressure drag (boundary layer
normal pressure drag) can be a large part of the total drag and it is therefore necessary to delay
separation for as long as possible. The streamlining of any object is a means of increasing the
fineness ratio to reduce the curvature of the surfaces and thus the adverse pressure gradient
(fineness ratio of an aerofoil is Chord / Thickness).

20. Interference Drag. On a complete aircraft the total drag is greater than the sum of the
values of drag for the separate parts of the aircraft. The additional drag is the result of flow
interference at wing / fuselage, wing / nacelle and other such junctions leading to modification of the
boundary layers. Further turbulence in the wake causes a greater pressure difference between fore
and aft surface areas and therefore additional resistance to movement. For subsonic flight this
component of total drag can be reduced by the addition of fairings at the junctions, e.g. at the trailing
edge wing roots.

21. Effect of Configuration, Altitude and Speed. Inferring from the above explained factors,
changes in configuration, altitude and speed will cause the Zero Lift Drag to vary as follows:

(a) Effect of Configuration. The zero lift drag is unaffected by the lift but is affected by
the dynamic pressure and the effective frontal area. With changes in configuration the
effective frontal area changes which causes the form drag and the interference drag to
change in the same ratio as the change in effective frontal area.

(b) Effect of Altitude. For a given aircraft, flying at the same TAS and configuration,
an increase in altitude will reduce the dynamic pressure acting upon it and thus the zero lift
drag will reduce. The reduction in dynamic pressure will be directly proportional to the
change in relative density. However, if the ac is flying at the same IAS, the dynamic pressure
and therefore the zero lift drag will not change.

(c) Effect of Speed. The effect of speed alone on the zero lift drag is the most
important. If all the other factors like altitude, configuration, weight etc. are kept constant, the
zero lift drag is directly proportional to the square of EAS.
FIS Book 1: Aerodynamics 96

LIFT DEPENDENT DRAG

General

22. All of the drag which arises because the aircraft is producing lift is called lift dependent drag.
It comprises mainly of induced drag, but also contains increments of the types of drag which make up
zero lift drag. The latter are more apparent at high angles of attack.

23. Wing Tip Vortices. If a finite


rectangular wing at a positive angle of
attack is considered, the spanwise
pressure distribution will be as shown in
Fig 7-4. On the underside of the wing,
the pressure is higher than that of the
surrounding atmosphere so the air spills
around the wing tips, causing an
outward airflow towards them. On the
upper surface, the pressure is low, and
the air flows inwards. This pattern of
airflow results in a twisting motion in the
air as it leaves the trailing edge. Viewed
from just downstream of the wing, the Fig 7-4: Spanwise Flow

air rotates and forms a series of vortices


along the trailing edge, and near the
wing tips the air forms into a
concentrated vortex. Further
downstream all of the vortices collect
into trailing vortices as seen in Fig 7-5.
The wing tip vortices intensify under
high lift conditions, e.g. during
manoeuvre, and the drop in pressure at
the core may be sufficient to cause
vapour trails to form (Fig 3-3). Fig 7-5: Trailing Vortices

Induced Downwash

24. The main consequence of these


vortices is that the air in the immediate
vicinity of the wing, and behind it, Fig 7-6: Downwash
acquires a downward velocity
97 Drag

component. This phenomenon is known as


induced downwash, though the adjective
‘induced’ is often omitted. It may be
measured in terms of either downwash
velocity, usually denoted by “w”, or
downwash angel, denoted by “ε”. These two Fig 7-7: Induced Downwash Angle
parameters are, of course, related, as shown
by the velocity triangle in Fig 7-7. It is clear from the diagram that tan ε = w / V. But w is always
small compared with V, so that ε is a small angle, and tan ε = ε. We therefore use the relationship

ε = w
V

Where ε is measured in radians.

25. In general, the downwash angle varies across the span. However, there are some cases in
which the downwash is constant across the span, and this is commonly assumed to be approximately
true for most wings of conventional planform, i.e., straight and moderately tapered. The downwash
also varies in the stream-wise direction. It reaches its ultimate value little more than a chord length
behind the trailing edge and its mean value at the wing itself can be shown to be one half of this
ultimate value. The downwash is particularly important for two reasons:

(a) It reduces the effective angle of attack of the wing. This affects both the lift and drag
characteristics of the wing adversely.

(b) In a conventional aircraft design, downwash affects the flow over the tailplane. This
has important consequences in connection with the stability of the aircraft.

Lift and Downwash

26. The lift produced by a wing is imparted to it through the variations in pressure over its surface.
This lift force has its reaction in the downward momentum, which is imparted to the air, as it flows over
the wing. Thus the lift of the wing is equal to the rate of transport of downward momentum of this air.

27. This downward momentum is measured in terms of the induced downwash described above.
A simple analysis can be made on the basis of the rather sweeping assumption that the air affected,
i.e., deflected downwards by the wing, is that contained within an imaginary cylinder whose diameter
is the span if the wing. This assumption is, of course, a very crude one, nevertheless it leads to
results which are same as derived from more complex mathematical analysis which is beyond the
scope of this chapter.
FIS Book 1: Aerodynamics 98

28. Consider then, this cylinder of air, as


illustrated in Fig 7-8. The area of cross-section
of the cylinder (π r2) of affected air is ¼ π b2.
The rate of mass flow of affected air past the
wing (ρAV) is therefore ρ ¼ π b2 V. The rate of
transport of downward momentum (MV) is
therefore ρ ¼ π b2 V w, and this must equal
Fig 7-8: Airflow Influenced by the Aerofoil
the lift, L. Thus:

CL ½ ρ V 2 S = ρ ¼ π b2 V w
w = 2CL . S = 2CL
V π b2 π.A

Where A is the aspect ratio. This gives the ultimate value of the downwash angle for downstream of
the wing. At the wing, the mean downwash is only half of this value, so that it is given by.

ε = __CL__ (7.1)
π.A

29. On the basis of this assumption, it is clear that only a finite mass of air is affected by a finite
wing. If we consider unit span of an infinite wing, the air above this unit span forms part of a cylinder
of infinite radius, and its mass is therefore infinite. Since the lift of this unit span is finite, it follows that
the downward momentum imparted to this air in unit time is finite, and since the mass of the air is
infinite the induced downwash velocity must be zero. This is already clear, since there are no trailing
vortices to produce downwash, in the case of a two- dimensional wing.

30. This conclusion is also easily reached from equation (7.1). For a two-dimensional wing, A is
infinite, so that ε is zero for all values of CL. Equation (7.1) also shows that the greater the aspect
ratio the lesser is the downwash at any given CL. Finally, it also shows that when the lift is zero so
also is the downwash. A result which could be inferred simply from the fact that the downward
momentum only arises as the reaction to lift.

Induced Drag

31. The trailing vortices which are shed from near the tips of a finite wing contain energy
associated with the rotational velocities. This energy is abstracted from the airflow, so that some
power must be provided to maintain the airflow at a given velocity. This power must equal the rate of
flow of energy associated with the trailing vortices. This is equivalent to saying that there is now a
further drag force of the wing, to be added to its profile drag. This is known as induced drag, and it
forms a very important part of the total drag of a finite wing.
99 Drag

32. We shall determine an expression for induced drag by two methods. The first consists of
finding an expression for the energy abstracted in the trailing vortices. This is, in effect, the additional
kinetic energy acquired by the air due to its downwash velocity. In Fig 7-7, V’ represents the resultant
of the free stream velocity V, and the downwash velocity w, so that:

V’2 = V2 + w2
V’2 - V2 = w2

Now the rate of mass flow of affected air past the wing is ρ ¼ π b2 V, (see para 28) so that the
rate of increase of kinetic energy (½mv2 ) in flowing past the wing, due to downwash, is:

½ . ¼ . ρ . π . b2 . V . (V’2 - V2) = ρ . π . b2 . V .w2

This must equal the power required to overcome the induced drag, and this power is the
product of the induced drag with velocity. Thus:

D i . V = 1/8 ρ . π . b2 . V . w2

Where D i, is the induced drag. The induced drag coefficient is then given by

CDi ½ ρ V 2 S . V = 1/8 ρ . π . b2 . V . w2
CDi = ¼ . π . (b2 / s) . (w / V)2

As we have seen, far downstream of the downstream of the wing.

W = ε = 2CL and b2 / S = A So that


V πA

CDi = CL 2 (7.2)
πA

33. This relation is of utmost importance. A more exact analysis shows that the expression for
induced drag coefficient is in fact a function of the spanwise load distribution. The above value can be
shown to be the best possible value, i.e., the one that gives the smallest induced drag. It has been
obtained here by assuming that the downwash angle is constant across the span, and this is true for
an elliptical wing, or, more precisely, a wing with elliptic loading. For any other load distribution, the
induced drag is given by

CDi = K . CL2 (7.3)


πA

Diagrammatic Explanation

34. Another way of understanding induced drag is by the diagrammatic explanation. The trailing
vortices modify the whole flow pattern. In particular, they alter the flow direction and speed in the
FIS Book 1: Aerodynamics 100

vicinity of the wing and tail surfaces. The trailing vortices therefore have a strong influence on the lift,
drag and handling qualities of an aircraft. The airflow behind the wing is drawn downwards therefore
this effect is known as downwash which also influences the flow over the wing itself, with important
consequences. Firstly the angle of attack relative to the modified total airstream direction is reduced.
This in effect is a reduction in angle of attack, and means that less lift will be generated unless the
angle of attack is increased. The second and more important consequence is that what was
previously the lift force vector is now tilted backward relative to the free stream flow. There is
therefore a rearward component of the force which is induced drag or vortex drag. A further serious
consequence of downwash is that the airflow approaching the tailplane is deflected downwards so
that the effective angle of attack of the tailplane is reduced.

35. Consider a section of a wing of


infinite span which is producing lift but has
no trailing edge vortices. Fig 7-9 a shows
a total aerodynamic reaction (TR), which
is divided into lift (L) and drag (D). The lift
component, being equal and opposite to
weight, is at right angles to the direction of
flight, and the drag component is parallel
and opposite to the direction of flight. The
angle at which the total reaction lies to the
relative airflow is determined only by the
angle of attack of the aerofoil as the total
reaction is considered to be perpendicular
to the RAF.

36. Fig 7-9 b shows the same section,


but of a wing of finite span and therefore
having trailing edge vortices. The effect of
the induced downwash, due to the
vortices, is to tilt downwards the effective
relative airflow, thereby reducing the
effective angle of attack. To regain the
consequent loss of lift the aerofoil must be
raised until the original lift value is restored
(Fig 7-9 c). The total reaction now lies at
Fig 7-9 a, b & c: Induced Drag
the original angle, but relative to the
effective airflow, the component parallel to the direction of flight is longer. This additional value of the
drag, resulting from the presence of wing vortices, is known as induced drag or vortex drag.
101 Drag

Factors Affecting Induced Drag

37. The main factors affecting vortex formation and therefore induced drag are:

(a) Planform.
(b) Aspect ratio.
(c) Lift and weight.
(d) Speed.

For an elliptical planform, which gives the minimum induced drag at any aspect ratio:

CDi = CL2 Where CDi = Coefficient of induced drag


πA A = Aspect ratio

A correction factor k is required for other planforms. Only for elliptical planforms the value of k = 1,
for any other planform it is greater than one.

38. Planform. Induced drag is greatest where the vortices are greatest, that is, at the wing tips
and so, to reduce the induced drag the aim must be to achieve an even spanwise pressure
distribution. An elliptical planform has unique properties due to elliptic spanwise lift distribution, viz:

(a) The downwash is constant along the span.


(b) For a given lift, span and velocity, this planform creates the minimum induced drag.

An elliptical wing poses manufacturing difficulties, fortunately a careful combination of taper and
washout, or section change at the tips can approximate to the elliptic ideal.

39. Aspect Ratio (AR). It has been previously stated that if the AR is infinite then the induced
drag is zero. The nearer we can get to this impossible configuration, the lesser induced drag is
produced. The wing tip vortices are aggravated by the increased tip spillage, caused by the
transverse flow over the longer chord of a low AR wing and the enhanced induced downwash affects
a greater proportion of the shorter span. Induced drag is inversely proportional to the AR, e.g. if the
AR is doubled, then the induced drag is halved.

40. Effect of Lift and Weight. The induced downwash angle and therefore the induced drag
depends upon the difference in pressure between the upper and lower surfaces of the wing, and this
pressure difference is the lift produced by the wing. It follows then that an increase in CL (e.g. during
manoeuvres or with increased weight) will increase the induced drag at that speed. In fact, the
induced drag varies as CL2 or the lift2 and therefore as weight2, at a given speed.

41. Effect of Speed and Altitude. If, while maintaining level flight, the speed is reduced to say,
half the original, then the dynamic pressure producing the lift (½ρV2) is reduced four times. To restore
the lift to its original value, the value of CL must be increased four-fold. The increased angle of attack
FIS Book 1: Aerodynamics 102

necessary to do this inclines the lift vector to the rear, increasing the contribution to the induced drag.
In addition, the vortices are affected, since the top and bottom aerofoil pressures are altered with the
angle of attack. If an ac is flown at a constant TAS, keeping all other factors like weight, surface area
and aspect ratio etc. constant, then with a change in altitude, for the same TAS, the IAS will reduce,
which will therefore cause the induced drag to increase. However if the aircraft is flying at the same
IAS, the induced drag at the two levels will remain unchanged.

42. High Angles of Attack. Induced drag at high angles of attack, such as that occurs at take-
off, can account for nearly three-quarters of the total drag, falling to an almost insignificant figure at
high speed.

Increments of Zero Lift Drag Resulting from Lift Production

43. Effect of Lift. Remembering that two of the factors affecting zero lift drag are transition
point and adverse pressure gradient, consider the effect of increasing lift from zero to the maximum
for manoeuvre. Forward movement of the peak of the low pressure envelope will cause earlier
transition of the boundary layer to turbulent flow, and the increasing adverse pressure gradient will
cause earlier separation. Earlier transition increases the surface friction drag and earlier separation
increases the form drag.

44. Frontal Area. As an aircraft changes its angle of attack, either because of a speed change
or due to manoeuvre, the frontal area presented to the airflow is changed and consequently the
amount of form drag is changed.

45. Interference Drag. Interference drag arises from the mixing of the boundary layers at
junctions on the airframe. When the aircraft is producing lift, the boundary layers are thicker and more
turbulent and therefore create greater energy losses where they mix.

46. The increments of surface friction drag, form drag and interference drag arise because the
aircraft is producing lift, and these values of additional drag may be included in lift dependent drag. It
follows that the greater the lift, the greater the drag increments, and in fact, this additional drag is only
really noticeable at high angles of attack.

47. Variation of Drag with Angle of Attack. Fig 7-10 shows that total drag varies steadily with
change of angle of attack, being least at small negative angles and increasing on either side. The
rate of increase becomes marked at angles of attack above about 8° and after the stall it increases at
a greater rate. The sudden rise at the stall is caused by the turbulence resulting from the breakdown
of steady flow.
103 Drag

Fig 7-10: Typical Drag Curve for a Fig 7-11: Variation of L/D Ratio with
Cambered Section Angle of Attack

48. Variation of Lift / Drag Ratio with Angle of Attack. It is apparent that for a given amount
of lift it is desirable to have the least possible drag from the aerofoil. Typically, the greatest lifting
effort is obtained at an angle of attack of about 16° and least drag occurs at an angle of attack of
about –2°. Neither of these angles is satisfactory, as the ratio of lift to drag at these extreme figures is
low. What is required is the maximum lifting effort compared with the drag at the same angle, i.e. the
highest lift/drag ratio (L/D ratio).

49. The L/D ratio for an aerofoil at any selected angle of attack can be calculated by dividing the
CL at that angle of attack by the corresponding CD. In practice the same result is obtained irrespective
of whether the lift and drag or their coefficients are used for the calculation:

L = CL ½ ρ V2 S = CL
W CD ½ ρ V2 S CD

50. Fig 7-11 shows that the lift/drag ratio increases rapidly up to an angle of attack of about 4° at
which point the lift may be between 12 to 25 times the drag, the exact figure depending on the aerofoil
used. At larger angles the L/D ratio decreases steadily, because even though the lift itself is still
increasing the proportion of drag is rising at a faster rate. One important feature of this graph is the
indication of the angle of attack for the highest L/D ratio. This angle is one at which the aerofoil gives
its best all-round performance. At a higher angle the required lift is obtained at a lower, and hence,
uneconomical speed where as at a lower angle it is obtained at a higher, and also uneconomical,
speed.
FIS Book 1: Aerodynamics 104

TOTAL DRAG

General

51. It was stated in para 2


that total drag can be
considered in terms of zero lift
drag and lift dependent drag
and Fig 7-12 is a graph showing
variations of total drag with
EAS. This curve is valid for one
particular aircraft for one weight
only in level flight and is the
most important factor in
performance theory. Speeds of
particular interest are indicated
on the graph Fig 7-12: Total Drag Curve

(a) Minimum Power Required Speed (VMP). On the drag curve in Fig 7-12 let us
consider a point A and construct ordinates to the two axes. The area of the rectangle so
formed is the product of drag and speed.

Drag X Speed = Force X Velocity = Work done in unit time = Power.

Therefore the area under A is a measure of the power required to fly the aircraft in level flight,
at that speed. The smallest possible area under the graph will be found to the left of the
minimum drag speed and will therefore represent the minimum value of drag x speed for
level flight. The shaded area is the minimum product of drag and velocity and is therefore the
minimum power required speed (VMP).

(b) Minimum Drag Speed (VIMD). The graph shows the variation of total drag with
speed in level flight, and lift is therefore constant. The maximum value of L/D is at the lowest
drag speed and it coincides with where the drag is minimum, i.e. at VIMD.

(c) Maximum EAS / Drag Ratio Speed (1.32 VIMD). If in the figure 7-12, a line is drawn
from the origin to the curve, the angle subtended by the line with the horizontal axis is a
measure of the ratio of EAS/Drag. In particular, cot θ = EAS/Drag. The ratio of EAS / Drag is
maximum where the angle θ is minimum. The only point which satisfies this condition is C on
the tangent to the drag curve, drawn through the origin. This is the main aerodynamic
consideration for best range and is usually expressed as a fraction of min drag speed. It has
the value of 1.32 VIMD.
105 Drag

52. Effect of Weight. It has been


emphasized that the drag curves
illustrated in Fig 7-12 is valid for only one
stated weight (or lift). Fig 7-13 shows that
when the weight is changed and therefore
the CL, to maintain level flight, there is a
corresponding change in the lift
dependent drag. Since this component of
total drag varies as CL2 (or Weight2), it
follows that the difference in total drag will
be greatest at high angles of attacks and
least at the lower angles. Furthermore, a
change in weight will alter the point at Fig 7-13: Effect of Weight on Drag Curves
which the lift dependent and zero lift drag
curves cross, and so change the minimum
drag speed, e.g., an increase in weight will
increase the VIMD as well as the total drag.

53. Effect of High Drag Devices.


It is sometimes necessary to decrease the
VIMD. This is achieved by deliberately
increasing the zero lift drag. This is done
by using airbrakes, barn-door flaps, drag
parachutes etc., which have the effect of
increasing the zero lift drag (Fig 7-14) and
will also increase the total drag.
Fig 7-14: Effect of Increasing Zero Lift Drag

Coefficient of Drag

54. The drag coefficient of a wing is found experimentally, at a quoted Reynolds number, from the
equation:

Drag = CD ½ ρ V 2 S = CD q S

CD = Drag
qS

55. Coefficient of total drag is dependent upon the following factors:

(a) Angle of attack.


(b) Shape i.e. section and planform.
(c) Surface condition.
FIS Book 1: Aerodynamics 106

(d) Reynolds Number.


(e) Speed of sound (Mach number).

56. As the total drag is consisting of two components i.e. zero lift drag and the lift dependent drag,
the total drag coefficient for the whole aircraft has two components i.e.

CD (Total) = CD0 + CDi where CDi = KCL2


πA

57. The variation of total drag coefficient


with angle of attack is illustrated in Fig 7-15. It
is usually more common to find CD plotted
against CL. Also it is more difficult to draw a
typical drag coefficient curve than a CL curve
because the value of Zero Lift Drag varies Fig 7-15: Variation of CD with
considerably from section to section. Angle of Attack

58. Effect of Angle of Attack. At low angles of attack, and therefore CL, the drag is mostly zero
lift drag. At high angles of attack the drag is mostly induced drag plus the increases in the zero lift
drag.

Effect of Shape

59. The drag coefficient of an aerofoil can be affected by many subtle changes in section, only
the more important and obvious are discussed here. The value of minimum CD is affected by t/c ratio
and camber, that is, a thin high speed aerofoil
having little or no camber will have a low value
of minimum CD. However, the most
interesting change is produced by rearward
positioning of the point of maximum thickness.

60. Most low drag aerofoils have the


maximum thickness between 40% and 50%
aft of the leading edge. This ensures that a
large part of the wing surface is covered by a
laminar boundary layer, hence the reduced
drag. Such aerofoils are known as laminar
wings and Fig 7-16 shows the effect, known
as laminar buckets, on the CD curve for a
Fig 7-16: Drag Characteristics of a
cambered and a symmetrical section. Laminar Section
107 Drag

61. The width and depth of the favourable range is determined by the shape of the symmetrical
fairing. The central value of the lift co-efficient is known as the optimum or ideal lift co-efficient (CLOPT
or CLI) and its value is decided by the shape of the mean camber line and amount of camber. By
suitable design the favourable range may be placed to cover the most common range of lift co-
efficient for a particular aircraft throughout most of its flight, with obvious benefits in performance and
economy.

62. Effect of Roughness. The


effect of surface roughness on the drag
coefficient is illustrated in Fig 7-17. The
roughened leading edge causes
immediate transition to turbulent flow, thus
increasing the surface friction drag.

63. Effect of Reynolds Number.


As Reynolds Number is increased, the
boundary layer becomes turbulent over
more of its length, and more energetic.
The separation point moves aft resulting in Fig 7-17: Effect of Roughened Leading
a reduction in pressure drag. The Edge on CD

increased turbulent boundary layer also


produces an increase in surface friction
drag, and the overall result when
considering only the effects on an aerofoil
in a wind-tunnel may be a reduction of
boundary layer drag with increasing
Reynolds Number as illustrated in Fig 7-
18. However, when considering the
aircraft as a whole the additional
interference drag resulting from the
thickened and lengthened boundary layer
invariably results in an overall increase in
total drag with increasing Reynolds Fig 7-18: Effect of Reynolds Number on the CD
Number.

64. Effects of Mach Number. The effect of Mach Number is discussed in the chapter on High
Speed Aerodynamics.

65. High Speed Drag. The effects of compressibility covering wave drag, trim drag and
interference drag are discussed later in the book.
FIS Book 1: Aerodynamics 108

ANGLE OF ATTACK AND AIRFRAME EFFICIENCY

Minimum Drag

66. In the earlier paragraphs it was seen that in level flight lift is constant and therefore, the L/D
ratio is maximum when drag is least. To find the optimum angle of attack for any given wing it is
necessary to consider how the L/D ratio varies with angle of attack. At any speed in level flight:

L = CL q S = CL
D CD q S CD

Since the relationship between CL, CD and angle of attack is known, the CL/CD ratio can be plotted.
Minimum drag is therefore obtained at the angle of attack appropriate to the minimum value of the
CL/CD ratio.

Minimum Power Required

67. Power = force X velocity = drag X TAS. To show how power required varies with angle of
attack it is necessary to derive an expression involving coefficients, thus in level flight:

D = D W = CD W
L CL

In level flight; W = L = CL ½ ρ V2 S,
½
therefore, V = W_W
CL½ρS

thus power required = DXV


½
∝ CD W W .
CL CL½ρS
½
∝ CD W3 .
CL3/2 ½ρS

68. From the above equation it is seen that the power required in level flight is dependent on the
following factors:

3/2
(a) Weight (power ∝ W ).
(b) Altitude (power ∝ 1/√ρ)
(c) _CD_ which varies with angle of attack. In particular, minimum power required occurs
CL3/2
When _CD_ is a minimum or when CL3/2 is a maximum.
CL3/2 CD
109 Drag

Like the CL/CD ratio, this only occurs at one particular angle of attack as illustrated in Fig 7-19.

Maximum EAS / Drag Ratio

69. In later chapters it will be seen that the range of a jet aircraft is dependent on the ratio of
speed to drag. As before, the following identities can be used to produce a non-dimensional
parameter:

½
D = CD .W and V = __W__
CL CL½ρS
½
from which V ∝ __W__ X CL X 1
D CL½ρS CD W
½
Thus V ∝ CL ___1___
D CD W½ρS

The range of a jet aircraft is therefore dependent on:

(a) Weight (range ∝ 1 / √ W ).


(b) √CL / CD which varies with angle of
attack

In particular the angle of attack necessary for


maximum jet range is that appropriate to the
maximum value of √CL / CD . The graphs of CL/CD,
√CL / CD and CL3/2/ CD are calculated from the basic
lift/drag data obtained from calculations or wind-
tunnel tests. The values shown are representative
of the whole aircraft, i.e. they include the drag of the
fuselage, nacelles and tail unit.
Fig 7-19: Performance Parameters of an
Aircraft
70. Significance of the Three Points on the
Drag Curve.

(a) Minimum drag speed (VIMD) gives the following:


(i) Endurance speed for turbo jets.
(ii) Range speed for propeller aircraft.
(iii) Maximum angle of climb for turbo jets.
(iv) Maximum gliding range for both types.

(b) Minimum power speed (VIMP) gives the following:


(i) Endurance speed for propeller aircraft.
FIS Book 1: Aerodynamics 110

(ii) Maximum gliding endurance for both types.

(c) Maximum TAS / Drag speed gives best range speed for turbo jets.

Summary

71. The drag created when an aircraft is not producing lift, e.g. in a truly vertical flight path, is
called zero lift drag and it comprises of:

(a) Surface friction drag.


(b) Form drag (boundary layer normal pressure drag).
(c) Interference drag.

72. Surface friction drag is dependent upon:

(a) Total wetted area.


(b) Viscosity of air.
(c) Rate of change of velocity across the flow which is dependent upon:
(i) Transition point.
(ii) Surface condition.
(iii) Speed and size.
(iv) Adverse pressure gradient.

73. Form drag is dependent upon:

(a) Separation point:


(i) Transition point.
(ii) Adverse pressure gradient.
(b) Streamlining.

74. Interference drag is caused by the mixing of airflows at airframe junctions.

75. Zero lift drag varies as the square of the equivalent air speed (EAS).

76. Lift dependent drag comprises:

(a) Induced drag (vortex drag).


(b) Increments of:
(i) Surface friction drag.
(ii) Form drag.
(iii) Interference drag.
111 Drag

77. Induced drag varies as:

(a) CL2
(b) _ 1_
V2
(c) Weight2.
(d) 1____
Aspect ratio

REQUIREMENTS OF AN IDEAL AEROFOIL

78. Having studied the factors affecting lift and drag of an aerofoil let us see what are the
requirements of an ideal aerofoil:

(a) High CL max. As this factor governs the landing speeds, higher the CL max, lower will
be the landing speed and nothing contributes more towards the safety of an aircraft than low
landing speeds.

(b) Low CD min. This indicates that as far as the wings are concerned the aircraft will
experience a low resistance or drag and will therefore be able to attain higher speed. It is
important that the CD min should not be low only at a certain angle of attack, but it should
remain small over a large range of angles.

(c) Good L/D Ratio. As mentioned above CL max and CD min effect the extreme ends of
speed range, whereas L/D concerns more the speeds and angle of attack of normal flight.
High L/D ratio spells efficiency, good weight carrying capacity, less fuel consumed for
distance covered and less expense.

(d) High maximum value of CL3/2/ CD. Since power required is proportional to this ratio,
the greater the value of this ratio, lesser the power required and this is especially important
from the point of view of endurance.

(e) Small and Stable Movement of Centre of Pressure. If we can restrict this
movement, then we can rely upon the greatest pressures on the wing remaining in one fixed
position, and thus reduce the weight of the structure required to carry these pressures.
Stable movements imply that with increase in angle of attack, the nose down moment about
the leading edge should increase.
FIS Book 1: Aerodynamics 112
113

CHAPTER 8

STALLING

Introduction

1. We have learnt earlier that the nature of the boundary layer determines the stalling
characteristics of a wing. In particular the phenomenon of boundary layer separation is extremely
important. This chapter will first discuss what happens when a wing stalls, it will then look at the
aerodynamic symptoms and the variations in the basic stalling speed and finally consider autorotation.

2. Boundary Layer Separation. Boundary layer separation is produced as a result of the


adverse pressure gradient developed around the body. The low energy air close to the surface is
unable to move in the opposite direction to the pressure gradient and the flow close to the surface
starts to flows in the reverse direction to the free stream. The development of separation is shown in
Fig 8-1. A typical velocity profile is shown corresponding to point A, while a little further down the
surface, at point B, the adverse pressure gradient will have modified the velocity profile as shown. At
point C the velocity profile has been modified to such an extent that at the surface flow has ceased.
Further down the surface, at point
D, the flow close to the surface has
reversed and the flow is said to
have separated. Point C is defined
as the separation point. In the
reversed flow region, aft of this
point, the flow is eddying and
turbulent, with a mean velocity of Fig 8-1: Velocity Profile
motion in the opposite direction to
the free stream.

3. Trailing Edge Separation. On a normal subsonic section virtually no flow separation


occurs before the trailing edge at low angles of attack. The flow remains attached over the rear part of
the surface in the form of a turbulent boundary
layer. As the angle of attack increases, the
adverse pressure gradient is also increased
and the boundary layer will begin to separate
from the surface near the trailing edge of the
wing (Fig 8-2). As the angle of attack is further
increased the separation point will move
forward along the surface of the wing towards Fig 8-2: Trailing Edge Separation
the leading edge. As the separation point
FIS Book 1: Aerodynamics 114

moves forward the slope of the lift vs. angles of attack curve decreases and eventually an angle of
attack is reached at which the wing is said to stall. The flow over the upper surface of the wing is then
completely broken down and the lift produced by the wing decreases. This type of turbulent trailing
edge separation is usually found on conventional low speed wing sections.

4. Leading Edge Separation. Another type of separation sometimes occur on thin wings with
sharp leading edges. This is laminar flow separation, the flow separating at the thin nose of the
aerofoil before becoming turbulent. Transition may then occur in the separated boundary layer and
the layer may then reattach further down
the body in the form of a turbulent
boundary layer. Underneath the
separated layer a stationary vortex is
formed and this is often termed a bubble
(Fig 8-3). The size of these bubbles
depend on the aerofoil shape and has
been found to vary from a very small Fig 8-3: Leading Edge Separation

fraction of the chord (short bubbles) to a length comparable to that of the chord itself (long bubbles).
The long bubbles affect the pressure distribution even at small angles of attack, reducing the lift-curve
slope. The eventual stall on this type of wing is more gradual. However, the short bubbles have little
effect on the pressure distribution, and hence on the lift curve slope, but when the bubble eventually
bursts the corresponding stall is an abrupt one.

5. Critical Angle of Attack. The marked


reduction in the lift coefficient, which accompanies
the breakdown of airflow over the wing, occurs at
the critical angle of attack for a particular wing.
Weight, bank angle, load factor, density altitude and
airspeed have no direct effect on the stalling angle
of attack. In subsonic flight an aircraft will always
stall at the same critical angle of attack except at
high Reynolds Numbers, as explained in the chapter
on Viscosity and Boundary Layer. A typical lift curve
showing the critical angle of attack is shown in Fig 8-
Fig 8-4: Lift Curve
4. It should be noted that not all lift is lost at the
critical angle of attack. In fact, the aerofoil will give a certain amount of lift up to 90°.

6. Aerodynamic Symptoms of Stall. The most consistent symptom of stall warning arises
from the separated flow behind the wing passing over the tail surfaces. The turbulent wake causes
buffeting of the control surfaces, which can usually be felt at the control column and rudder pedals.
As the separation point starts to move forward, to within a few degrees of the critical angle of attack,
115 Stalling

the buffeting will usually give adequate warning of the stall. On some aircraft separation may also
occur over the cockpit canopy to give additional audible warning. The amount of pre-stall buffet
depends on the position of the tail surfaces with respect to the turbulent wake. When the trailing edge
flaps are lowered the increased downwash angle behind the (inboard) flaps might reduce the amount
of buffet warning of the stall. (This is the reason of reduced pre stall buffet felt while carrying out stall
with flaps fully down on some aircraft like the Kiran.)

7. Pitching Moments. As the angle of attack is increased through the critical angle of attack
the wing pitching moment changes. Changes in the downwash angle behind the wing also cause the
tail pitching moment to change. The overall effect varies with aircraft type and may be masked by the
rate at which the elevator is deflected to increase the angle of attack. Most aircraft, however, are
designed to produce a nose-down pitching moment at the critical angle of attack.

Factors Affecting Aircraft Behaviour during Stall

8. The Effect of Wing Section. We have seen


that stalling is due to the effects of flow separation,
and is characterized by loss of lift as well as
increase in drag. Thus, if stalling occurs in flight, the
aircraft will lose height, unless some action is taken
to prevent it. Additionally there are aspects of aircraft
behaviour and handling at and near the stall which
depend on the design of the wing. The shape of the
wing section affects the behaviour of the ac near stall.
With some sections, the stall occurs very suddenly and
the reduction in lift is very marked (Fig 8-5). With others,
the approach to the stall is more gradual, and the Fig 8-5: Stalling Characteristics

reduction in lift is less pronounced.

9. It is desirable that the stall does not occur too suddenly, and the pilot should have adequate
warning, in terms of the handling qualities of the aircraft, of the approach of stall. This warning
generally takes the form of buffeting and lack of response to control inputs. If a particular wing is such that
it stalls too suddenly then it may be necessary to provide some artificial pre-stall warning device.

10. Features of wing section design, which affect the behaviour near stall, are:

(a) Leading edge radius of curvature.


(b) Thickness/chord ratio.
(c) Camber, and particularly the amount of camber near the leading edge.
(d) Chord wise location of the points of maximum thickness and maximum camber.
FIS Book 1: Aerodynamics 116

11. Generally, sharper the nose, thinner the wing, or further aft the positions of maximum
thickness and camber, the more sudden will be the stall.

12. The Effect of Protuberances. Any protuberances on the wing, or indeed elsewhere on the
aircraft, may significantly affect the stalling pattern by causing local flow separations. The main cause
of such affects is the positioning of the engines, externally carried weapon stores, fuel tanks or
configuration changes, since separation may occur near engine nacelles, or spring from the intakes of
jet engines etc.

13. Other protuberances may be deliberately placed on the wing with a view to control stalling
behaviour. These include:

(a) Vortex Generators. Vortex generators can either take the form of metal projections
from the wing surface or of small jets of air issuing normal to the surface. Both types work on
the same principle of creating vortices which entrain the faster moving air near the top of the
boundary layer down into the more stagnant layer near the surface thus transferring
momentum which keeps the boundary layer attached further back on the wing and therefore
delay stall. There are many shapes for the metal projections, such as plain rectangular plates
or aerofoil sections, the exact selection and positioning of which depends on the particular
requirement of the designer. An advantage of the air jet type is that they can be switched off
for those stages of flight when they are not required and thus avoid the drag penalty.

(b) Boundary Layer Fences. These are simply small plates intended to prevent the
outward drift of the boundary layer, which is a factor in causing the tip stall. Notches in the
wing leading edges tend to produce a similar effect.

14. Tip stalling. The desirable stall pattern of any wing is a stall which begins on the root
section of the wing first. Therefore the wings of an aircraft are designed to stall progressively from the
root to the tip. The reasons for this are threefold:

(a) To induce early buffet symptoms over the tail surface.


(b) To retain aileron effectiveness up to the critical angle of attack.
(c) To avoid a large rolling moment, which would arise if the tip of one wing stalled before
the other (wing drop).

15. A rectangular straight wing will usually stall from the root because of the reduction in effective
angle of attack at the tips caused by the wing tip vortex. If washout is incorporated to reduce vortex
drag, it also assists in delaying tip stall. A tapered wing on the other hand will aggravate the tip stall
due to the lower Reynolds Number (smaller chord) at the wing tip.
117 Stalling

16. The most common features designed to prevent wing tip stalling are:

(a) Washout. A progressive reduction of angle of incidence from wing root to the wing
tip, called the washout, will result in the wing root reaching its critical angle of attack before
the wing tip.

(b) Root Spoilers. By making the leading edge of the root sharper, the airflow has
more difficulty in following the contour of the leading edge and an early stall at the wing root is
induced.

(c) Change of Section. An aerofoil section with more gradual stalling characteristics
may be employed towards the wing tips (increased camber).

(d) Slats and Slots. The use of slats and/or slots on the outer portion of the wing
increases the stalling angle of that part of the wing.

STALLING SPEED

General

17. In level flight the weight of the aircraft is balanced by the lift, and from the lift formula it can be
seen that lift is reduced whenever any of the other factors in the formula are reduced. For all practical
purposes density (ρ) and wing area (S) can be considered constant (for a particular altitude and
configuration). If the engine is throttled back the drag will reduce the speed and from the formula lift
will be reduced. To keep the lift constant and so maintain level flight, the only factor that is readily
variable is the lift coefficient (CL).

18. As has been shown, the CL can be made larger by increasing the angle of attack, and by
doing so the lift can be restored to its original value so that level flight is maintained at the reduced
speed. Any further reduction in speed necessitates a further increase in the angle of attack, each
succeeding lower IAS corresponding to each succeeding higher angle of attack. Eventually, at a
certain IAS, the wing reaches its stalling angle, beyond which point any further increase in angle of
attack, in an attempt to maintain the lift, will precipitate a stall.

19. The speed corresponding to a given angle of attack is obtained by transposing the lift
formula, thus:

L = CL ½ ρo VI2 S
∴ V I2 = ___L___
CL ½ ρo S
Therefore, V I = _____√L____
√ C L ½ ρo S
FIS Book 1: Aerodynamics 118

If it is required to know the speed corresponding to the critical angle of attack, the value of CL
in the above formula is CL max and:

VI stall = ______√ L_____ (8.1)


√ CL max ½ ρo S

Inspection of this formula shows that the only two variables (clean aircraft) are V, and L, i.e.
VI stall ∝ √ Lift.

20. This relationship is readily demonstrated by the following examples:

(a) Steep Dive. When pulling out of a dive, if the angle of attack is increased to the
critical angle, separation will occur and buffet will be felt at a higher air speed.

(b) Vertical Climb. In true vertical flight lift is zero and no buffet symptoms will be
produced even at zero IAS.

21. There has always been difficulty in deciding upon an exact definition of stall or stalling
speed. The stall occurs because the smooth airflow over the wing becomes turbulent but this is a
gradual process. At quite small angles of attack there is some turbulence near the trailing edge. As
the angle of attack increases, the turbulence spreads forward. What is even more important is that
it also spreads span wise, usually from tip to root on highly tapered wings, and from root to tip on
rectangular wings. If we define the stall as being the break up of the airflow, when did it occur? There
may be buffeting of the tail plane or main planes, but this too may be from slight and unimportant to
fairly violent. As a result of the change from smooth to turbulent airflow the curve of lift coefficient
reaches a maximum and then starts to fall. Therefore the stalling angle is the angle at which the lift
coefficient is a maximum. But how does the pilot know that his aircraft is at its maximum value of
CL? In any case, how does he know what angle of attack is he flying at ? All, the pilot knows, is that if he
tries to fly below a certain speed he gets into difficulties.

22. Basic Stalling Speed. The most useful stalling speed to remember is the stalling speed
corresponding to the critical angle of attack in straight and level flight. Basic stalling speed may be
defined as the speed below which a clean aircraft of stated weight, with the engines throttled back,
can no longer maintain straight and level flight. This speed is listed in the Aircrew Manual for different
weights.

23. Applying these qualifications to equation (8.1), it can be seen that level flight requires a
particular value of lift, i.e. L = W and:

VB = _____√W_____ where VB = basic stalling speed. (8.2)


√CL max ½ ρo S
119 Stalling

24. If the conditions in para 22 are not met, the stalling speed will differ from the basic stalling
speed. The factors, which change VB, therefore are:

(a) Change in weight.


(b) Manoeuvre (load factor).
(c) Configuration (changes in CL max).
(d) Power and slipstream.

25. Weight Change. The relationship between the basic stalling speeds at two different weights
can be obtained from equation (8.2), i.e. the ratio of VB1 : VB2

VB1 : VB2 = ___√ W1___ : ___√ W2____


CL max ½ ρo S √CL max ½ ρo S

As the two denominators on the right hand side are identical:

VB1 : VB2 = √W1 : √W2 or


VB2_ = _√ W2_
VB1 √ W1

From which, we have

VB2 = VB1 X √ W2_ (8.3)


√ W1

where VB1 and VB2 are the basic stalling speeds at weights W1 and W2 respectively.

26. This relationship is true for any given angle of attack provided that the appropriate value of CL
is not affected by speed. The reason is that, to maintain a given angle of attack in level flight, it is
necessary to reduce the dynamic pressure (IAS) if the weight is reduced.

27. Manoeuvre. The relationship between the basic stalling speed and the stalling speed in any
other manoeuvre (VM) can be obtained in a similar way by comparing equation (8.1) with (8.2) i.e. the
ratio VM : VB as:

VM : VB = ____√ L______ : _____√ W______


√ (CL max ½ ρo S) √ (CL max ½ ρo S)

Again the denominators on the right-hand side are identical, so:

VM : VB = √ L : √ W or __VM_ = __√ L__


VB √W

from which, VM = VB _√ L_ (8.4)


√W
FIS Book 1: Aerodynamics 120

The relationship (L / W) is the load factor, n, and is indicated on the accelerometer (if fitted).
Thus VM = VB√n and in a 4g manoeuvre, the stalling speed is twice the basic stalling speed.

28. In the absence of an accelerometer the


artificial horizon may be used as a guide to the
increased stalling speed in a level turn. From Fig
8-6 It can be seen that the component of lift which is
acting in the vertical plane is balancing the weight of
the aircraft. Hence we have:

L Cos ϕ = W, therefore

_L_ = __1__ Fig 8-6: Forces on an Aircraft in Level Turn


W Cos ϕ

substituting this value of L/W in equation (8.4) We get:

VM = VB ___1___ (8.5)
√Cos ϕ

29. From above we can see that 1 / Cos ϕ = n, and VM = VB √ n. As each angle of bank
(ϕ) corresponds to a stated value of load factor (n), stalling speed in a manoeuvre becomes solely a
function of bank or load factor, increasing as the square root of the load factor. Typical values of load
factor determined by this relationship are:

φ 0° 15° 30° 45° 60° 75.5°


n 1.00 1.035 1.154 1.414 2.0 4.0

30. It can be seen that the stalling speed is increased to 1.2 times its basic value at 45° bank and
1.4 times at 60° bank, and at 75.5° bank, the stalling speed doubles. Therefore, significant increase in
stalling speed occurs as bank crosses 45°.

31. Configuration. From equation (8.2), it can be seen that the stalling speed is inversely
proportional to CL max i.e. in level flight,

VB ∝ __1___
√ CL max

32. Any change in CL max due to the operation of high lift devices or compressibility effects will
affect the stalling speed. In particular, lowering of flaps or extending slats will result in a new basic
stalling speed as stated in the next paragraph.
121 Stalling

33 Effects of High Lift Devices. The primary purpose of high lift devices (flaps, slots, slats,
etc.) is to increase the CL max of the aeroplane and reduce the stalling speed. The take-off and landing
speeds are consequently reduced. The effects of a typical high lift device is summarized here:

Configuration CL max Stalling AOA


Clean (flaps up). 1.5 20°
Flaps down. 2.0 18.5°

34. The principal effect of the extension of flaps is to increase the CL max and reduce the angle of
attack for any given lift coefficient. (As explained later in the chapter of High Lift Devices) The
increase in CL max provided by flap deflection reduces the stalling speed in a certain proportion, the
effect being described by the equation:

VF = VB √ (CL max)
√ (CL MF)
Where: VF = Stalling speed with flaps down.
VB = Basic stalling speed (without flaps)
CL max = Maximum lift co-efficient in clean configuration.
CL MF = Maximum lift co-efficient with flaps down.

For example, assume the airplane whose CL values are stated above has a stalling speed of 200
kmph at the landing weight in the clean configuration. If the flaps are lowered the stall speed is
reduced to 173 kmph.

35. Because of the stated variation of stall speed with CL max, large changes in CL max are
necessary to produce significant changes in stalling speed. This effect is illustrated by the typical
values shown below:

Percent increase in CL max 2 10 50 100 300


Percent reduction in stalling speed 1 5 18 29 50

36. The contribution of high lift devices must be considerable to cause large reduction in stalling
speed. The most elaborate combination of flaps, slots, slats and boundary layer control throughout the
span of the wing would be required to increase CL max by 300%. A common case is that of a propeller
driven transport aircraft which experiences appox 70% increase in CL max by full flap deflection. A
single engine jet fighter with a thin sweptback wing obtains about 20% increase in CL max by full flap
deflection. Thin aerofoil sections with sweepback impose distinct limitations on the effectiveness of
flaps, which shall be discussed in detail in the subsequent chapters.
FIS Book 1: Aerodynamics 122

37. So we can summarize that any change in CL max due to operation of high lift devices or due to
compressibility effects will affect the stalling speed. In particular, the lowering of flaps or extending the
slats will result in a new stalling speed. These changes are usually listed in the Aircrew Manual.

38. Power. At the basic stalling speed the engines are throttled back and it is assumed that the
weight of the aircraft is entirely supported by the wings. If power is applied at the stall the high nose
up attitude produces a vertical
component of power, which
assists in supporting the
weight and less force is
required from the wings. This
reduction in lift is achieved at
the same angle of attack by
reducing the dynamic
pressure (IAS) and results in Fig 8-7: Comparison of Power On and Power Off Stall
a lower stalling speed as
shown in Fig 8-7.

39. From Fig 8-7 it can be seen that L = W – T sin φ which is less than the power-off case and, as
VM ∝ √ L, VM with power on is less than VB.

40. It should be noted that, for simplicity, the load on the tailplane has been ignored and the
engine thrust line assumed parallel to the wing chord line.

41. Slipstream Effect. For a propeller-powered


aircraft the velocity of the slipstream behind the
propellers causes an additional effect as shown in Fig
8-8 and enumerated below.

(a) Vector Change of RAF. The vector


addition of the free stream and slipstream
velocities results in a change in the RAF over
that part of the wing affected by the propellers.

(b) Increase in Velocity. The local


increase in dynamic pressure will result in more
lift behind the propellers relative to the power-
Fig 8-8: Slipstream Effect
off case. Thus, at stalling angle of attack, a
lower IAS is required to support the weight, i.e. the stalling speed is reduced.
123 Stalling

(c) Decrease in Angle of Attack. The reduction in angle of attack partially offsets the
increase in dynamic pressure but does not materially alter the stalling speed unless CL max is
increased. The only effect will be to increase the attitude at which the stall occurs. Unless
the propeller slipstream covers the entire wing, probability of a wing drop will increase. This
is because the wings will stall progressively from the tips, which are at a higher angle of
attack than the wings behind the propellers.

Effect of Altitude on Stalling Speed

42. As we know that in level flight L = W = CL ½ ρV2S. To consider the effect of density we will
assume that the CL max remains constant and for a given aircraft the surface area S is constant.
As altitude increases air density (ρ) will be less, and this means that in order to keep CL ½ ρV2S
equal to the weight, the stalling speed (TAS) has to be greater than at ground level. This fact is
not, however, of great importance, because, although the stalling speed (TAS) is in reality
greater, the air speed indicator, which is in itself worked by the effect of the air density, will record
the same indicated speed when the aeroplane stalls as it did at sea level. In other words, the
indicated stalling speeds will remain the same at all altitudes considering only the effect of changes in
density with altitude.

43. At high altitude airfields, e.g. Leh, Kargil etc., the true landing speed of an aeroplane
will be appreciably higher than on sea level airfields, and in tropical countries the air density is
decreased owing to the high temperatures, and the true landing speed is consequently increased.
The take-off speed, and the run required, is also increased in both these instances, and this is
perhaps an even more important consideration.

44. An increase in altitude will also alter compressibility and viscosity effects and, generally
speaking, cause the indicated stalling speed to increase. This happens primarily due to reduction
in RN with increase in altitude. The reduction in RN causes the CL max to reduce and hence the
basic stalling speed is increased. However, this particular consideration is usually significant only
above altitudes of 20,000 ft.

45. Recovery from Stall. Two factors common to operating at max lift conditions are the angle
of attack and the pressure distribution. The maximum lift coefficient of a particular wing configuration
is obtained at one angle of attack and pressure distribution. The factors of weight, load factors, bank
angle, power and density altitude have no bearing on the critical angle of attack. These factors only
affect the basic stalling speed of the aircraft. During any condition of flight and at any speed and
attitude the aircraft will stall if the critical angle of attack is exceeded.

46. From the above, it is evident that recovery from the stall involves a very simple concept. Since
stall is precipitated by an excessive angle of attack, the angle of attack must be decreased. This is a
FIS Book 1: Aerodynamics 124

fundamental concept common to all aircraft. The reduction of angle of attack is achieved by moving
the stick centrally forward, that is a nose down control input. A recovery from the stall will be indicated
by a steady build up of speed.

47. Having studied the basics of stall and aircraft behaviour in the stalled state of flight, now we
will touch upon a few types of stalls, which are peculiar to a specific condition of flight and/or design of
aircraft. You will notice that in a few of the stalls discussed below, the aircraft is actually well below the
critical angle of attack, but still behaves in a stalled manner. The intent here is to make a mention of
these, the details of which shall be discussed in the subsequent chapters.

48. Deep Stall. The reduced effective angle of attack of very low aspect ratio wings can
delay the stall considerably. Some delta wings have no measurable stalling angle upto 40° of
angle of attack. At this sort of angle, the drag is so high that the flight path is usually inclined
downwards at a steep angle to the horizontal. Apart from a rapid rate of descent, and possible
loss of stability and control, such aircraft may have a fairly high nose up attitude with
respect to the horizon and this can be deceptive. The condition is called the super stall or deep
stall, although the wing may be far from a true stall and still be generating appreciable lift.

49. Some of the characteristics of swept wings i.e. tip stall and pitch-up increase the possibility
of entering the deep stall region. Recovery from a stall is dependent on the pitching moment
from the wing and the pitching moment from the tailplane. In the case of a straight wing aircraft
at the stall, the CP tends to move rearwards, giving a nose-down pitching moment. On a swept
wing, however, the tendency is for the CP to remain forward possibly giving a nose-up pitching
moment. All now depends on the tailplane, which, if set low relative to the mainplane, will be
working clear of the separated flow from the mainplane and will therefore provide a nose-down
pitching moment and give a
conventional response to the stall.
If, however, the tail plane is set high
(i.e. T-tail like the IL - 76), then
increasing the angle of attack
reduces the vertical separation
between mainplane and tail plane
and, at the stall, the tailplane is
immersed in the separated
mainplanes wake and therefore
suffers a reduction in effectiveness
(Fig 8-9). The nose-up pitching
moment of the mainplane may be
sufficient to hold the aircraft in its
Fig 8-9: Deep Stall
stalled condition. The rate of
125 Stalling

descent, which develops, aggravates the condition by increasing the angle of attack even further
and may make subsequent recovery impossible. While being unsuitable from the deep stall and pitch-
up aspects, the T-tail configuration is dictated by the mounting of engines on the rear fuselage. Aircraft
in which the T-tail design is used, an artificial stall reaction is usually designed into the control system
in the form of a stick shaker, stick pusher or both. Both are sensitive to pitching rates and angle of
attack. The stick shaker provides the stall warning (sometimes accompanied by an audible warning)
and then, if this is ignored by the pilot, at a slightly higher angle of attack the stick pusher
simulates the conventional nose-down pitching moment at the stall by pushing the stick forward.
Some aircraft, particularly those involved in the investigation of stall characteristics, are equipped
with the drag parachute or rocket in the tail. When used in a deep stall, the chute or rocket provides
the essential nose-down pitch moment to enable recovery to be made.

50. Shock Stall. A shock wave constitutes a very large adverse pressure gradient, and thus
produces a thickening of the boundary layer, and often flow separation. In transonic range the
upper surface shock wave becomes strong enough to cause the flow to separate. When this happens,
the lift coefficient begins to decrease, and the drag rises rapidly. This phenomenon is known as the
shock stall. It differs from the conventional stall in that it may occur at low angle of attack, though at
higher angles of attack it will occur at lower Mach number. Associated with the shock-induced
separation, there may also be intense buffeting, just
as there is with a high angle of attack stall at low
speeds. With subsonic aircraft, it is generally this
buffeting which sets the upper limit to its operating
speed range, just as the high angle of attack stall
sets the lower limit. The shock stall will be studied
in detail in the chapter on High Speed Flight.

Autorotation

51. The auto rotational properties of a wing are


due to the negative slope of the CL vs. α curve when
α is greater than the stalling angle of attack.

52. With reference to the stalled aircraft (Fig 8-


10) if the aircraft starts to roll, there will be a
component of flow induced tending to increase the
angle of attack of the down going wing and
decrease the angle of attack of the up going wing.
The cause of the roll may be either accidental (wing
Fig 8-10: Vector Addition of Roll
drop), deliberate (further effects of applied rudder) or
use of aileron at the stall.
FIS Book 1: Aerodynamics 126

53. The effect of this


change in angle of attack on
the CL and CD (Fig 8-11) is that
the “damping in roll” effect
normally produced at low α is
now reversed. The increase in
angle of attack of the down-
going wing decreases the CL
and increases the CD by a
larger amount. Conversely, the
decrease in angle of attack of
the up-going wing slightly
decreases the CL and CD. The Fig 8-11: Autorotation
difference in lift produces a rolling moment towards the down-going wing, tending to increase the
angular velocity. This angular acceleration is further increased by the roll induced by the yawing
motion due to the large difference in drag.

54. The cycle is automatic in the sense that the increasing rolling velocity sustains or even
increases the difference in angle of attack. It should be noted, however, that at higher angles, the
slope of the CL curve might recover again to zero. This may impose a limit on the α at which
autorotation is possible.

55. It is emphasized that autorotation, like the stall, is an aerodynamic event which is dependent
on angle of attack. It is therefore possible to autorotate the aircraft in any attitude and at speeds
higher than the basic stalling speed. This principle is the basis of many of the more advanced
aerobatic manoeuvres.

Summary

56. Boundary layer separation is produced as a result of the adverse pressure gradient
developed round the body.

57. In subsonic flight an aircraft will always stall at the same critical angle of attack.

58. The wing of an aircraft is designed to stall progressively from the root to the tip. The reasons
for this are:

(a) To induce early buffet symptoms over the tail surface.


(b) To retain aileron effectiveness up to the critical angle of attack.
127 Stalling

(c) To avoid a large rolling moment, which would arise if the tip of one wing stalled before
the other.

59. The most common design features for preventing tip stalling are:

(a) Washout.
(b) Root spoilers.
(c) Change of section.
(d) Slats and slots.

60. Basic stalling speed is the speed below which a clean aircraft of stated weight, with engines
throttled back, can no longer maintain straight and level flight.

61. The formula for the basic stalling speed is : VB = ____√W______


√CL max ½ ρo S
Factors, which affect VB are:

(a) Change in weight: VB2 = VB1 __√ W2__


√ W1
(b) Manoeuvre (load factor): VM = VB √n or VM =VB ___√ 1__
√ Cos ϕ
(c) Configuration (changes in CL max): VB ∝ __1___
√ CL max
(d) Power and Slip-stream: VM with power on < VB
FIS Book 1: Aerodynamics 128
129

CHAPTER 9

SPINNING

Introduction

1. Spinning is a complicated subject to analyze in detail. It is also a subject about which it is


difficult to make generalizations which are true for all aircraft. One type of aircraft may behave in a
certain manner in a spin whereas another type will behave completely differently under the same
conditions. This chapter is based on a deliberately induced, erect spin to the right although inverted
and oscillatory spins are discussed in later paragraphs.

2. The accepted sign conventions applicable to this chapter are given in the Table and Fig 9-1
below.

Sign Conventions used in this Chapter

AXIS LONGITUDINAL LATERAL NORMAL


(Symbol) (x) (y) (z)
Positive Direction Forwards To right Downwards
Angular Velocity Roll Pitch Yaw
Symbol p q r
Moment of Inertia A B C
Inertial Moments L M N

Positive Direction Rolling moment to Pitching moment Yawing moment to


right nose-up right

Fig 9-1: Aircraft Reference Axis


FIS Book 1: Aerodynamics 130

3. Phases of the Spin, The spin manoeuvre can be divided into three phases:

(a) The incipient spin.


(b) The fully developed spin.
(c) The recovery.

4. The Incipient Spin. A necessary ingredient of a spin is the aerodynamic phenomenon


known as autorotation. This leads to an unsteady manoeuvre which is a combination of:

(a) The ballistic path of the aircraft, which is itself dependent on the entry attitude.
(b) Increasing angular velocity generated by the autorotative rolling moment and drag-
induced yawing moment.

5. The Steady Spin. The incipient stage may continue for some 2-6 turns after which the
aircraft will settle into a steady stable spin. There will be some sideslip and the aircraft will be rotating
about all three axes. For simplicity, but without suggesting that it is possible for all aircraft to achieve
this stable condition, the steady spin is qualified by a steady rate of rotation and a steady rate of
descent.

6. The Recovery. The pilot initiates the recovery by actions aimed at first opposing the
autorotation and then reducing the angle of attack (α) so as to unstall the wings. The aircraft may
then be recovered from the ensuing steep dive.

THE STEADY ERECT SPIN

7. While rotating, the aircraft will describe some sort of ballistic trajectory dependent on the entry
manoeuvre. To the pilot this will appear as an unsteady, oscillatory phase until the aircraft settles
down into a stable spin with steady rate of descent and rotation about the spin axis. This will occur if
the aerodynamic and inertia forces and moments can achieve a state of equilibrium. The attitude of
the aircraft at this stage will depend on the aerodynamic shape of the aircraft, the position of the
controls and the distribution of mass throughout the aircraft.

Motion of the Aircraft

8. The motion of the centre of gravity in a spin has two components:

(a) A vertical linear velocity (rate of descent = V).

(b) An angular velocity (Ω radians per sec) about a vertical axis, called the spin axis. The
distance between the CG and the spin axis is the radius of the spin (R) and is normally small
(about half wing span).
131 Spinning

The combination of these


motions result in a vertical spiral
or helix. The helix angle is small,
usually less than 10°. Fig 9-2
shows the motion of the aircraft in
spin.

9. As the aircraft always


presents the same face to the
spin axis, it follows that it must be
rotating about a vertical axis
passing through the centre of
gravity at the same rate as the
CG about the spin axis. This
angular velocity may be resolved
into components of roll, pitch and
yaw with respect to the aircraft
body axes. In the spin illustrated
in Fig 9-2 b the aircraft is rolling
right, pitching up and yawing
right. For convenience the
direction of the spin is defined by
the direction of yaw.

10. In order to understand


the relationship between these
angular velocities and aircraft
Fig 9-2: The Motion of an Aircraft in Erect
attitude it is useful to consider Spin to the Right
three limiting cases:

(a) Longitudinal Axis Vertical. When the longitudinal axis is vertical the angular
motion will be all roll.

(b) Lateral Axis Vertical. For the aircraft to present the same face (pilot’s head) to the
spin axis, the aircraft must rotate about the lateral axis. The angular motion is all pitch.

(c) Normal Axis Vertical. For the aircraft to present the same face (inner wing tip) to
the axis of rotation, the aircraft must rotate about the normal axis at the same rate as the
aircraft rotates about the axis of rotation. Thus the angular motion is all yaw.
FIS Book 1: Aerodynamics 132

11. Although these are hypothetical examples which may not be possible in practice, they
illustrate the relationship between aircraft attitude and angular velocities. Between the extremes
quoted in the previous paragraph, the motion will be a combination of roll, pitch and yaw, and
depends upon:

(a) The rate of rotation of the aircraft about the spin axis.

(b) The attitude of the aircraft, which is usually defined in terms of the pitch angle and the
wing tilt angle. Wing tilt angle, often confused with bank angle, involves displacement about
the normal and the longitudinal axes.

12. The aircraft’s attitude in spin also has an important effect on the sideslip present (Fig 9-2 c).
If the wings are level, there will be outward sideslip, that is, the relative airflow will be from the
direction of the outside wing (to port in the diagram). If the attitude of the aircraft is changed such that
the outer wing is raised relative to the horizontal, the sideslip is reduced. This attitude change can
only be due to a rotation of the aircraft about the normal axis. The angle through which the aircraft is
rotated, in the plane containing the lateral and longitudinal axes, is known as the wing tilt angle and is
positive with the outer wing up. If the wing tilt can be increased sufficiently to reduce the sideslip
significantly, the pro-spin aerodynamic rolling moment will be reduced.

Balance of Forces in Spin

13. Only two forces are acting on the centre of gravity while it is moving along its helical path (Fig
9-2 a).

(a) Weight (W).


(b) The aerodynamic force (N) coming mainly from the wings.

The resultant of these two forces is the centripetal force necessary to produce the angular motion.

14. Since the weight and centripetal force act in a vertical plane containing the spin axis and the
CG, the aerodynamic force must also act in this plane, i.e. it passes through the spin axis. It can be
shown that, when the wing is stalled, the resultant aerodynamic force acts approximately
perpendicular to the wing. For this reason it is sometimes called the wing normal force.

15. If the wings are level (lateral axis horizontal), from the balance of forces in Fig 9-2a:

(a) Weight = Drag = CD½ρV2S


½
V = __W___ (9.1)
CD½ρS
133 Spinning

(b) Lift = Centripetal force

CL ½ρV2S = W Ω2 R
G

R = g CL ½ ρ V2 S (9.2)
W Ω2
Where: R = Spin radius
S = Area
V = Rate of descent
W = Weight

If the wings are not level, it has been seen that the departure from the level condition can be regarded
as a rotation of the aircraft about the longitudinal and normal axes. Usually this angle, the wing tilt
angle, is small and does not affect the following reasoning.

Effect of Attitude on Spin Radius

16. If for some reason the angle of attack is increased by a nose-up change in the aircraft’s
attitude, the vertical rate of descent (V) will decrease because of the higher CD (equation 9.1). The
increased angle of attack on the other hand, will decrease CL, which, together with the lower rate of
descent, results in a decrease in spin radius, (equation 9.2). It can also be shown that an increase in
pitch increases the rate of spin, which will decrease R still further.

17. The two extremes of aircraft attitude possible in the spin are shown in Fig 9-3. The actual
attitude adopted by an aircraft will depend on the balance of moments.

Fig 9-3: Effect of pitch attitude

18. The effects of pitch attitude are summarized below. An increase in pitch (e.g. flat spin) will:

(a) Decrease the rate of descent.


(b) Decrease the spin radius.
(c) Increase the spin rate.

It can also be shown that an increase in pitch will decrease the helix angle.
FIS Book 1: Aerodynamics 134

19. In a steady spin, equilibrium is achieved by a balance of aerodynamic and inertia moments.
The inertia moments result from a change in angular momentum due to the inertia cross coupling
between the three axes.

20. The angular momentum about an axis depends on the distribution of mass and the rate of
rotation. It is important to get a clear understanding of the significance the spinning characteristics of
different aircraft and the effect of controls in recovering from the spin.

Moment of Inertia (I)

21. A concept necessary to predict the


behaviour of a rotating system is that of moment
of inertia. This quantity not only expresses the
amount of mass but also its distribution about
the axis of rotation. It is used in the same way
that mass is used in linear motion. For example,
the product of mass and linear velocity
measures the momentum or resistance to
acceleration of a body moving in a straight line.
Similarly, the product of moment of inertia (mass
Fig 9-4: Two Rotors of Same Weight and
distribution) and angular velocity measures the Angular Velocity
angular momentum of a rotating body (Fig 9-4).

22. The concept of moment of inertia may be applied to an aircraft by measuring the distribution
of mass about each of the body axes in the following ways:

(a) Longitudinal Axis. The distribution of mass about the longitudinal axis determines
the moment of inertia in the rolling plane, which is denoted by A. An aircraft with fuel stored in
the wings and in external tanks will have a large value of A, particularly if the tanks are close
to the wing tips. The tendency in modern high speed aircraft towards thinner wings has
necessitated the stowage of fuel elsewhere and this, combined with lower aspect ratios, has
resulted in a reduction in the value of A for those modern high performance fighter and
training aircraft.

(b) Lateral Axis. The distribution of mass about the lateral axis determines the
moment of inertia in the pitching plane which is denoted by B. The increasing complexity of
modern aircraft has resulted in an increase in the density of the fuselage with the mass being
distributed along the whole length of the fuselage and a consequent increase in the value of
B.
135 Spinning

(c) Normal Axis. The distribution of mass about the normal axis determines the
moment of inertia in the yawing plane, which is denoted by C. This quantity will be
approximately equal to the sum of the moments of inertia in the rolling and pitching planes.
C, therefore, will always be larger than A or B.

23. These moments of inertia measure the mass distribution about the body axes and are
decided by the design of the aircraft. It will be seen that the values of A, B and C for a particular
aircraft may be changed by altering the disposition of equipment, freight and fuel.

Inertia Moments in a Spin

24. The inertia moments generated in a spin are described below by assessing the effect of the
concentrated masses involved. Another explanation using a gyroscopic analogy, is given at the and
of the chapter.

(a) Roll. It is difficult to represent the rolling moments using concentrated masses, as is
done for the other axes. For an aircraft in the spinning attitude under consideration (inner
wing down pitching nose up), the inertia moment is anti-spin, i.e. tending to roll the aircraft out
of the spin. The equation for the inertia rolling moment is:

L = - (C-B) q r (9.3)

(b) Pitch. The imaginary


concentrated masses of the fuselage,
as shown in Fig 9-5, tend to flatten the
spin. The equation for the inertia
pitching moment is:

M = (C-A) r p (9.4)
Fig 9-5: Inertia pitching Moment
(c) Yaw. The inertia couple is
complicated by the fact that it comprises
two opposing couples caused by the
wings and the fuselage (Fig 9-6).
Depending on the dominant component,
the couple can be of either sign and
varying magnitude. The inertia yawing
moment can be expressed as:

N = - (B-A) p q (9.5) Fig 9-6: Inertia Yawing Moments

25. This is negative and thus anti-spin when B > A and is positive and pro-spin when A > B.
FIS Book 1: Aerodynamics 136

26. The B/A ratio has a profound effect on the spinning characteristics of an aircraft.

Aerodynamic Moments

27. It is now necessary to examine the contributions made by the aerodynamic factors in the
balance of moments in roll, pitch and yaw. These are discussed separately below.

28. Aerodynamic Rolling Moments. The aerodynamic contributions to the balance of


moments about the longitudinal axis to produce a steady rate of roll are as follows:

(a) Rolling Moment due to Sideslip. The design features of the aircraft, which
contribute towards positive lateral stability, produce an aerodynamic rolling moment as a
result of sideslip. It can be shown that, even at angles of attack above the stall, this still
remains true and the dihedral effect induces a rolling moment in the opposite sense to the
sideslip. In the spin the relative airflow is from the direction of the outer wing (outward
sideslip) and the result is a rolling moment in the direction in which the aircraft is spinning.
This contribution is therefore pro-spin.

(b) Autorotative Rolling Moment. In the chapter on flight controls, it is shown that the
normal damping in roll effect is reversed at angles of attack above the stall. This contribution
is therefore pro-spin.

(c) Rolling Moment due to Yaw. The yawing velocity in the spin induces a rolling
moment for two reasons:

(i) Difference in Speed of the Wings. Lift of the outside wing is increased
and that of the inner wing decreased, inducing a pro-spin rolling moment.

(ii) Difference in Angle of Attack of the Wings. In a spin the direction of the
free stream is practically vertical whereas the direction of the wing motion due to yaw
is parallel to the longitudinal
axis. The yawing velocity not
only changes the speed but
also the angle of attack of the
wings. Fig 9-7 illustrates the
vector addition of the yawing
velocity to the vertical velocity
of the outer wing. The effect
is to reduce the angle of Fig 9-7: Change in Angle of Attack
attack of the outer wing and Due to Yaw
increase that of the inner
137 Spinning

wing. Because the wings are stalled (slope of CL curve is negative), the CL of the
outer wing is increased and the CL of the inner wing decreased thus producing
another pro-spin rolling moment.

(d) Aileron Response. Experience has shown that the ailerons produce a rolling
moment in the conventional sense even though the wing is stalled.

29. Aerodynamic Pitching Moments. The aerodynamic contributions to the balance of


moments about the lateral axis to produce a steady rate of pitch are as follows:

(a) Positive Longitudinal Static Stability. In a spin the aircraft is at a high angle of
attack and therefore disturbed in a nose-up sense from the trimmed condition. The positive
longitudinal stability responds to this disturbance to produce a nose-down aerodynamic
moment. This effect may be considerably reduced if the tailplane lies in the wing wake.

(b) Damping in Pitch Effect. When the aircraft is pitching nose-up the tailplane is
moving down and its angle of attack is increased (the principle is same as the damping in roll
effect). The pitching velocity therefore produces a pitching moment in a nose-down sense.
The rate of pitch in a spin is usually very low and consequently the damping in pitch
contribution is small.

(c) Elevator Response. The elevators act in the conventional sense. Down-elevator
increases the nose-down aerodynamic moment whereas up-elevator produces a nose-up
aerodynamic moment. It should be noted, however, that down-elevator usually increases the
shielded area of the fin and rudder.

30. Aerodynamic Yawing Moments. The overall aerodynamic yawing moment is made up of a
large number of separate parts, some arising out of the yawing motion of the aircraft and some arising
out of the side slipping motion. The main contributions to the balance of moments about the normal
axis to produce a steady rate of yaw are as follows:

(a) Positive Directional Static Stability. When sideslip is present keel surfaces aft of
the CG produce an aerodynamic yawing moment tending to turn the aircraft into line with the
sideslip vector (i.e. directional static stability or ‘weathercock effect’). This is an anti-spin
effect, the greatest contribution to which is from the vertical fin. Vertical surfaces forward of
the CG will tend to yaw the aircraft further into the spin, i.e. they have a pro-spin effect. In a
spin outward sideslip is present which, usually produces a net yawing moment towards the
outer wing, i.e. in an anti-spin sense. Because of possible shielding effects from the tailplane
and elevator and also because the fin may be stalled, the directional stability is considerably
reduced and this anti-spin contribution is usually small.
FIS Book 1: Aerodynamics 138

(b) Damping in Yaw Effect. Applying the principle of the damping in roll effect to the
yawing velocity, it has been seen that the keel surfaces produce an aerodynamic yawing
moment to oppose the yaw. The greatest contribution to this damping moment is from the
rear fuselage and fin. In this respect the cross-sectional shape of the fuselage is critical and
has a profound effect on the damping moment. The following figures give some indication of
the importance of cross-section:

Cross-Section Damping Effect (Anti-Spin)


Circular 1.0
Rectangular 2.5
Elliptical 3.5
Round top / flat bottom 1.8
Round bottom / flat top 4.2
Round bottom / flat top with strakes 5.8

Table 9-1: Effect of Fuselage on Damping in Roll

Damping effect = Damping from body of given cross-section


Damping from circular cylinder

Fuselage strakes are useful devices for improving the spinning characteristics of prototype
aircraft. The anti-spin damping moment is very dependent on the design of the tailplane/fin
combination. Shielding of the fin by the tailplane can considerably reduce the effectiveness of
the fin. In extreme cases a low-set tailplane may even change the anti-spin effect into pro-
spin.

(c) Rudder Response. The rudder acts in the conventional sense, i.e. the in-spin
rudder produces pro-spin yawing moment and out-spin rudder produces anti-spin yawing
moment. Because of the shielding effect of the elevator (para 29c), it is usual during recovery
to pause after applying out-spin rudder so that the anti-spin yawing moment may take effect
before down-elevator is applied.

Balance of Moments

31. In para 16 it was seen that the balance of forces in the spin has a strong influence on the rate
of descent. It does not, however, determine the rate of rotation, wing tilt or incidence at which the spin
occurs. The balance of moments is much more critical in this respect. The actual attitude, rate of
descent, sideslip, rate of rotation and radius of a spinning aircraft can only be determined by applying
specific numerical values of the aircraft’s aerodynamic and inertia data to the general relationships
discussed below.
139 Spinning

32. Rolling Moments. The balance of rolling moments in an erect spin is:

(a) Pro-spin. The aerodynamic rolling moments in an erect spin are:


(i) Autorotative rolling moment.
(ii) Rolling moment due to sideslip.
(iii) Rolling moment due to yaw.

(b) Anti-spin. The inertia rolling moment, - (C - B) r q, is anti-spin.

These factors show that autorotation is usually necessary to achieve a stable spin. A small
autorotative rolling moment would necessitate larger sideslip to increase the effect of rolling moment
due to sideslip. This, in turn, would reduce the amount of wing tilt and make the balance of moments
in yaw more difficult to achieve, however the balance of moments in this axis is not as important as in
the other two.

33. Pitching Moments. In para 25 it was seen that the inertia pitching moment, (C – A) r p, of
the aircraft is always nose-up in an erect spin. This is balanced by the nose-down aerodynamic
pitching moment. The balance between these two moments is the main factor relating angle of attack
to rate of rotation in any given case and equilibrium can usually be achieved over a wide range. It can
be shown that an increase in pitch will cause an increase in the rate of rotation (spin rate). This, in
turn, will decrease the spin radius (para 16).

34. Yawing Moments. The balance of yawing moments in an erect spin is:

(a) Pro-spin.
(i) Yawing moment due to applied rudder.
(ii) A small contribution from the wing, due to yaw, is possible at large angles of
attack.
(iii) Yawing moment due to sideslip (vertical surfaces forward of CG).
(iv) Inertia yawing moment, (A - B) pq, if A > B.

(b) Anti-spin.
(i) Inertia yawing moment, (A-B) pq, if B>A.
(ii) Yawing moment due to sideslip (vertical surfaces aft of the CG).
(iii) Damping in yaw effect.

It can be seen that in-spin rudder is usually necessary to achieve balance of the yawing moments and
hold the aircraft in a spin.

35. Normal Axis. For conventional aircraft (A and B nearly equal), it is relatively easy to
achieve balance about the normal axis and the spin tends to be limited to a single set of conditions
FIS Book 1: Aerodynamics 140

(angle of attack, spin rate, attitude). For aircraft in which B is much larger than A, the inertia yawing
moment can be large and thus difficult to balance. This is probably the cause of the oscillatory spin
exhibited by these types of aircraft.

36. Yaw and Roll Axis. The requirements of balance about the yaw and roll axes greatly limit
the range of angle of attacks in which spinning can occur and determines the amount of sideslip and
wing tilt involved. The final balance of the yawing moments is achieved by the aircraft taking up the
appropriate angle of attack at which the inertia moments just balance the aerodynamic moments.
This particular angle of attack also has to be associated with the appropriate rate of spin required to
balance the pitching moments and the appropriate angle of sideslip required to balance the rolling
moments.

SPIN RECOVERY

Effect of Controls in Recovery from a Spin

37. The relative effectiveness of the three controls in recovering from a spin will now be
considered. Recovery is aimed at stopping the rotation by reducing the pro-spin rolling moment
and/or increasing the anti-spin yawing moment. The yawing moment is more important but, because
of the strong cross-coupling between motions about the three axes through the inertia moments, the
rudder is not the only means by which yawing may be induced by the pilot. Once the rotation has
stopped the angle of attack is reduced and the aircraft recovered.

38. The control movements which as experience has shown, are generally most favourable to the
recovery from the spin have been known and in use for a long time, i.e. apply full opposite rudder and
then move the stick forward until the spin stops, maintaining the ailerons neutral. The rudder is
normally the primary control but, because the inertia moments are generally large in modern aircraft,
aileron deflection is also important. Where the response of the aircraft to rudder is reduced in spin,
the aileron may even be the primary control although in the final analysis it is its effect on the yawing
moment, which makes it work.

39. The initial effect of applying a control deflection will be to change the aerodynamic moment
about one or more axes. This will cause a change in aircraft attitude and a change in the rates of
rotation about all the axes. These changes will, in turn, change the inertia moments.

Effect of Ailerons

40. Even at the high angle of attack in spin the ailerons act in the normal sense. Application of
aileron in the same direction as the aircraft is rolling will therefore increase the aerodynamic rolling
141 Spinning

moment. This will increase the roll rate (p) and affect the inertia yawing moment, (A - B) p q. The
effect of an increase in p on the inertia yawing moment depends on the mass distribution or B/A ratio:

(a) B/A > 1. In an aircraft where B/A > 1, the inertia yawing moment is anti-spin
(negative) and an increase in p will decrease it still further, i.e. make it more anti-spin. The
increase in anti-spin inertia yawing moment will tend to raise the outer wing (increase wing
tilt), which will decrease the outward sideslip. This will restore the balance of rolling moments
by decreasing the pro-spin aerodynamic moment due to lateral stability. The increase in wing
tilt will also cause the rate of pitch, q, to increase, which, in turn:

(i) Causes a small increase in the anti-spin inertia rolling moment, - (C - B) rq,
(C > B) and thus helps to restore balance about the roll axis (para 32).
(ii) Further increases the anti-spin inertia yawing moment.

(b) B/A < 1. A low B/A ratio will reverse the effects described above. The inertia
yawing moment will be pro-spin (positive) and will increase with an increase in p.

41. Due to secondary effects associated with directional stability, the reversal point actually
occurs at a B/A ratio of 1.3 (Fig 9-8). Thus:

(a) B/A > 1.3. Aileron with roll (in-spin) has an anti-spin effect.

(b) B/A < 1.3. Aileron with roll (in-spin) has a pro-spin effect.

42. Some aircraft change their B/A ratio in


flight as stores and fuel are consumed. The pilot
has no accurate indication of the value of B/A ratio
and, where this value may vary either side of 1.3, it
is desirable to maintain ailerons neutral to avoid an
unfavourable response, which may delay or even
prohibit recovery.

43. An additional effect of aileron applied with Fig 9-8: Yawing Moment (N)
per Degree of Aileron
roll is to increase the anti-spin yawing moments
due to aileron drag.

Effect of Elevators

44. In para 29 it was seen that down-elevator produces a nose-down aerodynamic pitching
moment. This will initially reduce the nose-up pitching velocity (q). Although this will tend to reduce
alpha, the effect on the inertia yawing and rolling moments is as follows:
FIS Book 1: Aerodynamics 142

(a) Inertia Yawing Moment (A - B) p q. If B > A, the inertia yawing moment is anti-
spin. A reduction in q will make the inertia yawing moment less anti-spin, i.e. a pro-spin
change. When A > B, however, down-elevator will cause a change in inertia yawing moment
in the anti-spin sense.

(b) Inertia Rolling Moment -(C - B) r q. The inertia rolling moment is always anti-spin
because C > B. A reduction in q will therefore make it less anti-spin, which is again a change
in the pro-spin sense.

The result of these pro-spin changes in the inertia yawing and rolling moments is to decrease the wing
tilt thus increasing the sideslip angle (Fig 9-2) and rate of roll. It can also be shown that the rate of
rotation about the spin axis will increase.

45. Although the change in the inertia yawing moment is unfavourable, the increased sideslip
may produce an anti-spin aerodynamic yawing moment if the directional stability is positive. This
contribution will be reduced if the down elevator seriously increases the shielding of the fin and
rudder.

46. The overall effect of down-elevator on the yawing moments therefore depends on:

(a) The pro-spin inertia moment when


B > A.
(b) The anti-spin moment due to
directional stability.
(c) The loss of rudder effectiveness
due to shielding.

In general, the net result of moving the elevators


down is beneficial when A > B and rather less so
when B > A, assuming that the elevator movement Fig 9-9: Yawing Moment (N) per Degree of
Down Elevator
does not significantly increase the shielding of the
fin and rudder.

Effect of Rudder

47. The rudder is nearly always effective in


producing an anti-spin aerodynamic yawing
moment though the effectiveness may be greatly
reduced when the rudder lies in the wake of the
wing or tailplane. The resulting increase in the Fig 9-10: Yaw Moment (N) per Degree of
Anti-spin Rudder
wing tilt angle will increase the anti-spin inertia
143 Spinning

yawing moment (when B > A) through an increase in pitching velocity. The overall effect of applying
anti-spin rudder is always beneficial and is enhanced when the B/A ratio is increased.

48. The effect of the three controls on the yawing moment is illustrated in Figs 9-8, 9-9 and 9-10.

Inverted Spin

49. Fig 9-11 shows an aircraft in an


inverted spin but following the same flight
path as in Fig 9-2. Relative to the pilot the
motion is now compounded of a pitching
velocity in the nose-down sense, a rolling
velocity to the right and a yawing velocity
to the left. Thus roll and yaw are in
opposite directions, a fact that affects the
recovery actions, particularly if the aircraft
has a high B/A ratio.

50. The inverted spin is


fundamentally similar to the erect spin and
the principles of moment balance
discussed in previous paragraphs are
equally valid for the inverted spin. The
values of aerodynamic moments however
are unlikely to be the same, since in the
inverted attitude, the shielding effect of
the wing and tail may change markedly.

51. The main difference will be


caused by the change in relative positions Fig 9-11: The Inverted Spin
of the fin and rudder and the tailplane. An
aircraft with a low-mounted tailplane will tend to have a flatter erect spin and recovery will be made
more difficult due to shielding of the rudder. The same aircraft inverted will respond much better to
rudder during recovery since it is unshielded and the effectiveness of the rudder is increased by the
position of the tailplane. The converse is true for an aircraft with a high tailplane.

52. The control deflections required for recovery are dictated by the direction of roll, pitch and
yaw, and the aircraft’s B/A ratio. These are:

(a) Rudder to oppose yaw as indicated by the turn needle.


FIS Book 1: Aerodynamics 144

(b) Aileron in the same direction as the observed roll, if the B/A ratio is high.

(c) Elevator up is generally the case for conventional aircraft but, if the aircraft has a high
B/A ratio and suffers from the shielding problems previously discussed, this control may be
less favourable and may even become pro-spin.

Oscillatory Spin

53. A combination of high wing loading and high B/A ratio makes it difficult for a spinning aircraft
to achieve equilibrium about the yaw axis. This is thought to be the most probable reason for the
oscillatory spin. In this type of spin the rates of roll and pitch are changing during each oscillation. In
a mild form it appears to the pilot as a continuously changing angle of wing tilt, from outer wing well
above the horizon back to the horizontal once each turn and the aircraft seems to wallow in the spin.

54. In a fully developed oscillatory spin the oscillations in the rates of roll and pitch can be quite
violent. The rate of roll during each turn can vary from zero to about 200 degrees per second. At the
maximum rate of roll the rising wing is unstalled which probably accounts for the violence of this type
of spin. Large changes in attitude usually take place from fully nose-down at the peak rate of roll, to
nose-up at the minimum rate of roll.

55. The use of the controls to effect a change in attitude can change the characteristics of an
oscillatory spin quite markedly. In particular:

(a) Anything, which increases the wing tilt, will increase the violence of the oscillations,
e.g. in-spin aileron or anti-spin rudder.

(b) A decrease in the wing tilt angle will reduce the violence of the oscillations, e.g. out-
spin aileron or down-elevator.

The recovery from this type of spin has been found to be relatively easy, although the shortest
recovery times are obtained if recovery is initiated when the nose of the aircraft is falling relative to the
horizon.

56. Conclusions. The foregoing paragraphs make it quite clear that the characteristics of the
spin and the effect of controls in recovery are specific to type. In general the aerodynamic factors are
determined by the geometry of the aircraft and the inertial factors by the distribution of the mass. In
the final analysis the only correct recovery procedure is laid down in the Aircrew Manual for the
specific aircraft.
145 Spinning

GYROSCOPIC CROSS-COUPLING BETWEEN AXES

Introduction

57. In the preceding paragraphs, the effects of the inertia moments have been explained by
considering the masses of fuselage and wings acting either side of a centerline. The effect of these
concentrated masses when rotating, can be visualized as acting rather in the manner of the bob-
weights of a governor.

58. Another, and more versatile, explanation of the cross-coupling effects can be made using a
gyroscopic analogy regarding the aircraft as a rotor.

Inertia Moments in a Spin

59. The inertia moments generated in a spin are essentially the same as the torque exerted by a
precessing gyroscope. Figs 9-12, 9-13 and 9-14 illustrate the inertia or gyroscopic moments about
the body axes. These effects are described as follows:

(a) Inertia Rolling Moments (Fig 9-12). The angular momentum in the yawing plane is
Cr, and by imposing a pitching velocity of q on it, an inertia rolling moment is generated equal
to (-Crq), i.e. in the opposite sense to the direction of roll in an erect spin. The inertia rolling
moment due to imposing of yawing velocity on the angular momentum in the pitching plane is
in a pro-spin sense equal to (+Brq). The total inertia rolling moment is therefore equal to (B -
C)rq, or since C > B: - (C - B) rq.

Fig 9-12: Total Inertia Rolling Moment

(b) Inertia Pitching Moments (Fig 9-13). The angular momentum in the rolling plane
is Ap and imposing a yawing velocity of r on the rolling plane rotor causes it to precess in
pitch in a nose-down sense due to inertia pitching moment (-Apr). Similarly, the angular
momentum in the yawing plane is Cr, and imposing a roll velocity of p on the yawing plane
FIS Book 1: Aerodynamics 146

rotor generates an inertia pitching moment (+Crp) in the nose-up sense. The total inertia
moment is therefore (C - A)rp. In an erect spin, roll and yaw are always in the same direction
and C is always greater than A. The inertia pitching moment is therefore positive (Nose-up) in
an erect spin.

(c) Inertia Yawing Moments (Fig 9-14). Replacing the aircraft by a rotor having the
same moment of inertia in the rolling plane, its angular momentum is the product of the
moment of inertia and angular velocity (Ap). Imposing a pitching velocity (q) on the rotor will
generate a torque tending to precess the rotor about the normal axis in the same direction as
the spin. It can be shown that this inertia yawing moment is equal in value to +Apq where the
positive sign indicates a pro-sign torque. Similarly, the angular momentum in the pitching
plane is equal to Bq and imposing a roll velocity of p on the pitching plane rotor will generate
an inertia yawing moment in an anti-spin sense equal to -Bpq. The total inertia yawing
moment is therefore equal to (A - B)pq, or if B > A: - (B - A)pq.

Fig 9-13: Total Inertia Pitching Moment

Fig 9-14: Total Inertia Yawing Moment


147

CHAPTER 10

WING PLANFORMS

Introduction

1. The preceding chapters discussed the basic considerations of lift, drag, stalling and spinning
and explained the causes of these phenomena. However, it is now necessary to examine another
important aspect of the design of wings, i.e. the planform. The planform is the geometrical shape of
the wing when viewed from above. It largely determines the amount of lift and drag obtainable from a
stated wing area and has a pronounced effect on the value of the stalling angle of attack. The
previous discussion of aerodynamic forces concerned the properties of aerofoil sections in two-
dimensional flow with no consideration given to the influence of the planform. When the effects of
wing planform are introduced, attention must be directed to the existence of flow components in the
spanwise direction. In other words, aerofoil section properties deal with flow in two dimensions while
planform properties consider flow in three dimensions.

2. This chapter is concerned mainly with the low-speed effects of variations in wing planforms.
The high-speed effects are dealt with in the chapters on High Speed Flight. In order to fully describe
the planform of a wing, several terms are required. The terms having the greatest influence on the
aerodynamic characteristics are illustrated in Fig 10-1 are:

(a) Wing Area, ‘S’. It is simply the plan surface area of the wing. Although a portion of
the area may be covered by fuselage or nacelles, the pressure carryover on these surfaces
allows legitimate consideration of the entire plan area e.g. 16.7 m2 for the hawk, while the
wing area of IL – 76 is about 283 m2.

(b) Wing Span, ‘b’. It is measured tip to tip e.g. 26.27m for the Avro, 9.5m of HPT – 32
and 8.2 m for the LCA.

(c) Average Chord, ‘c’. It is the geometric average. The product of the span and the
average chord is the wing area (b X c = S) e.g. mean chord of 1.78 m for the hawk and
4.56m for mirage 2000 and 3.3 for the MiG 29.

(d) Aspect Ratio, ‘AR’. It is the ratio of the span and the average chord, i.e. AR=b/c. If
the planform has curvature and the average chord is not easily determined, an alternate
expression is: AR= b² / S. The aspect ratio is a fineness ratio of the wing and this quantity is
very powerful in determining the aerodynamic characteristics and structural weight. Typical
AR vary from 35 for a high performance glider to 3.5 for a jet fighter to 1.28 for a flying saucer
e.g the AR of IL-76 is 8.5, Hawk is 5.29, MiG 29 is 3.4 while that of MiG 21 is 2.22.
FIS Book 1: Aerodynamics 148

(e) Root Chord, ‘CR’. It is the chord at the wing centerline and the tip chord, CT, is
measured at the tip.

S = WING AREA

b = SPAN

c = AVERAGE CHORD

AR = ASPECT RATIO

= b/c

= b2 / s

CR = ROOT CHORD

CT = TIP CHORD

λ = TAPER RATIO

λ = CR / CT

Λ = SWEEP ANGLE

MAC = MEAN
AERODYNAMIC CHORD

Fig 10-1: Description of Wing Planform

(f) Taper Ratio λ. Considering the wing planform to have straight lines for the leading
and trailing edges, the taper ratio, λ (lambda), is the ratio of the tip chord to the root chord.
λ=CT/CR. The taper ratio affects the lift distribution and the structural weight of the wing. A
rectangular wing has a taper ratio of 1.0 while the pointed tip delta wing has a taper ratio of 0.
The taper ratio of hawk for example is 0.34.

(g) Sweep Angle, ‘Λ’ (Cap Lambda). It is usually measured as the angle between the
line of 25 percent chords and a perpendicular to the root chord. The sweep of a wing causes
definite changes in compressibility, maximum lift and stall characteristics. It is for example
210 for hawk, 570 for MiG 21 and the variable sweep from 160 to 720 for the MiG 27.
149 Wing Planforms

(h) Mean Aerodynamic Chord, ‘MAC’. It is the chord drawn through the centroid
(geographical center) of plan area. A rectangular wing of this chord and the same span would
have identical pitching moment characteristics. The MAC is located on the reference axis of
the aeroplane and is a primary reference for longitudinal stability considerations. Note that the
MAC is not the average chord but is the chord through the centroid of area. As an example,
the pointed-tip delta wing with a taper ratio of zero would have an average chord equal to
one-half the root chord but an MAC equal to two-thirds of the root chord.

The aspect ratio, taper ratio, and sweepback of a planform are the principal factors which determine
the aerodynamic characteristics of a wing. These same quantities also have a definite influence on
the structural weight and stiffness of a wing.

Development of Lift by a Wing

3. In order to appreciate
the effect of the planform on the
aerodynamic characteristics, it is
necessary to study the manner
in which a wing produces lift. Fig
10-2 illustrates the three
dimensional flow pattern which
results when a rectangular wing
creates lift. If a wing is producing
lift, a pressure differential will
exist between the upper and
lower surfaces, i.e. for positive
lift, the static pressure on the
upper surface will be less than
that on the lower surface. At the
tips of the wing, the existence of
this pressure differential creates
the spanwise flow components
shown in Fig 10-2. For the
rectangular wing, the lateral flow
Fig 10-2: Wing Three Dimensional Flow
developed at the tip is quite
strong and a strong vortex is created at the tip. The lateral flow, and consequent vortex strength,
reduces inboard from the tip towards the root until it is zero at the centerline. The existence of the tip
vortex is described by the drawings of Fig 10-2. The rotational pressure flow combines with the local
airstream flow to produce the resultant flow of the trailing vortex. An increase in angle of attack
increases lift and increases the flow deflection and strength of the tip vortices.
FIS Book 1: Aerodynamics 150

ASPECT RATIO

4. The aspect ratio (AR) of a wing is found by dividing the square of the wing span by the area of
the wing. Thus if a wing has an area of 25 square meter and a span of 10 meter, the aspect ratio is 4.
Another wing with the same span but with an area of 15 square meters would have an aspect ratio of
6.66. Another method of determining the aspect ratio is by dividing the span by the mean chord of the
wing. For example, a span of 12 meter with a mean chord of 1.2 meter gives an aspect ratio of 10.
From the preceding examples it can be seen that the smaller the area or mean chord in relation to the
span, the higher is the aspect ratio. A rough idea of the performance of a wing can be obtained from
knowledge of the aspect ratio.

Effect of Aspect Ratio on Induced Drag

5. The effect of aspect ratio on the induced drag is the principal effect of the wing planform. The
value of induced drag is determined by the formula:

Di = CDI ½ ρ v2 S Where CDI is the coefficient of induced drag

The relationship between CDi and AR is given by the formula:

CDi = K CL²
π AR

The relationship for induced drag coefficient emphasizes the need of a high aspect ratio for the
aeroplane, which is continually operated at high lift coefficients. In other words, aeroplane
configurations designed to operate at high lift coefficients during the major portion of their flight
(Gliders, cargo, transport, patrol, and antisubmarine types) demand a high aspect ratio wing to
minimize the induced drag. While the high aspect ratio wing will minimize induced drag, long, thin
wings increase structural weight and have relatively poor stiffness characteristics. This fact will
hamper the preference of a very high aspect ratio aircraft. Aeroplane configurations which are
developed for very high speed flight (especially supersonic flight) operate at relatively low lift
coefficients and demand great aerodynamic cleanness. These configurations of aeroplanes do not
have the same preference for high aspect ratio as the aeroplanes which operate continually at high lift
coefficients. This usually results in the development of low aspect ratio planforms for these aeroplane
configurations.

6. The effect of aspect ratio on the lift and drag characteristics is shown in Fig 10-3 for wings of
a symmetrical aerofoil section. The basic aerofoil section properties are shown on these curves and
these properties would be typical only of a wing planform of extremely high (infinite) aspect ratio.
When a wing of some finite aspect ratio is constructed of this basic section, the principal differences
will be in the lift and drag characteristics, the moment characteristics remain essentially the same. The
effect of decreasing aspect ratio on the lift curve is to increase the wing angle of attack necessary to
151 Wing Planforms

produce a given lift coefficient. The


difference between the wing angle
of attack and the section angle of
attack is the induced angle of
attack, αi = K CL / π AR which
increases with decreasing aspect
ratio. The wing with the lower
aspect ratio is less sensitive to
changes in angle of attack and
requires higher angles of attack for
maximum lift. When the aspect
ratio is very low (below 5 or 6) the
induced angles of attack are not
accurately predicted by the
elementary equation for αi and the
graph of CL versus α develops
distinct curvature. This effect is
especially true at high lift
coefficients where the lift curve for
Figure 10-3: Effect of Aspect Ratio on Wing
the very low aspect ratio wing is Characteristics
very shallow and CL max and stall
angle of attack are less sharply defined. The effect of aspect ratio on wing drag characteristics may be
appreciated from inspection of Fig 10-3. The basic section properties are shown as the drag
characteristics of an infinite aspect ratio wing. When a planform of some finite aspect ratio is
constructed, the wing drag coefficient is the sum of the induced drag coefficient, CDi = K CL2 / π AR,
and the section drag coefficient. Decreasing aspect ratio increases wing drag coefficient at any lift
coefficient since the induced drag coefficient varies inversely with aspect ratio. When the aspect ratio
is very low, the induced drag varies greatly with lift and at high lift coefficients, the induced drag is
very high and increases very rapidly with lift
coefficient.

7. The origin and formation of trailing edge and


wing tip vortices was explained in the Chapter on Drag
where it was shown that induced downwash was the
cause of induced drag. The downwash imparted to
the air is a measure of the lift provided by a wing.
Induced drag is inversely proportional to aspect ratio.
A graph showing the curves of three different aspect
ratio wings plotted against CD and angle of attack can
be seen in Fig 10-4. Fig 10-4: Effect of Aspect Ratio on CD
FIS Book 1: Aerodynamics 152

Aspect Ratio and Stalling Angle

8. A stall occurs when the effective angle of attack reaches the critical angle. As has been
shown in the chapter on Drag, induced downwash reduces the effective angle of attack of a wing.
Since induced drag is inversely proportional to aspect ratio it follows that a low aspect ratio wing will
have high induced drag, high induced downwash and a reduced effective angle of attack. The low
aspect ratio wing therefore has a higher stalling angle of attack than a wing of high aspect ratio.

9. The reduced effective angle of attack of very low aspect ratio wings can delay the stall
considerably. Some delta wings have no measurable stalling angle up to 40° or more inclination to
the flight path. At this sort of angle the drag is so high that the flight path is usually inclined
downwards at a steep angle to the horizontal. Apart from a rapid rate of descent, and possible loss of
stability and control, such aircraft may have a fairly high nose up attitude with respect to the horizon
and this can be deceptive. This condition is called the super stall or deep stall, although the wing may
be far from a true stall and still be generating appreciable lift.

Use of High Aspect Ratio

10. Aircraft types such as gliders, transport, patrol and anti-submarine demand a high aspect ratio
to minimize the induced drag (High performance gliders often have aspect ratios between 25 and 30).
While the high aspect ratio wing will minimize induced drag, long thin wings increase weight and have
relatively poor stiffness characteristics. Also the effects of vertical gusts on the airframe are
aggravated by increasing the aspect ratio. Broadly it can be said that the lower the cruising speed of
the aircraft, the higher the aspect ratios that can be usefully employed. Aircraft configurations which
are developed for very high speed flight (especially supersonic flight) operate at relatively low lift
coefficients and demand great aerodynamic cleanness. This usually results in the development of low
aspect ratio planforms.

Use of Low AR Wings & Caution

11. While the effect of aspect ratio on lift curve slope and drag due to lift (lift dependent drag) is
an important relationship, it must be realized that design for very high speed flight does not favor the
use of high aspect ratio planforms. Low aspect ratio planforms have structural advantages and allow
the use of thin, low drag sections for high speed flight. The aerodynamics of transonic and supersonic
flight also favour short span, low aspect ratio surfaces. Thus, the modern configuration of aeroplane
designed for high speed flight will have a low aspect ratio planform with characteristic aspect ratios of
two to four. The most important impression that should result is that the typical modern configuration
will have high angles of attack for maximum lift and very high drag due to lift at low flight speeds. This
fact is of importance to the Military Aviator because the majority of pilot-error accidents occur during
this regime of flight-during takeoff, approach, and landing. Induced drag predominates in these
regimes of flight.
153 Wing Planforms

12. The modern configuration of high speed aeroplane usually has a low aspect ratio planform
with high wing loading. When wing sweepback is coupled with low aspect ratio, the wing lift curve has
distinct curvature and is very flat at high angles of attack, i.e., at higher values of CL, the lift coefficient
increases very slowly with increase in angle of attack. In addition, the drag curve shows extremely
rapid rise at high lift coefficients since the drag due to lift is so very large (Fig 10-3). These effects
produce flying qualities which are distinctly different from a more "conventional" high aspect ratio
aeroplane configuration. Some of the most important ramifications of the modern high speed
configuration are:

(a) During Take Off. During Take off the aeroplane must not be over-rotated to an
excessive angle of attack (very high nose up attitude). Any given aeroplane will have some
fixed angle of attack (and CL) which produces the best takeoff performance and this angle of
attack will not vary with weight, density altitude, or temperature. An excessive angle of attack
produces additional induced drag and may have an undesirable effect on takeoff
performance. Takeoff acceleration may be seriously reduced and a large increase in takeoff
distance may occur. Also, the initial climb performance may be marginal at an excessively low
airspeed. There are modern configurations of aeroplanes of very low aspect ratio (plus sweep
back) which, if over-rotated during a high altitude, high gross weight takeoff cannot fly out of
what is called the ground effect. With the more conventional aeroplane configuration, an
excess angle of attack produces a well defined stall. However, the modern aeroplane
configuration at an excessive angle of attack has no sharply defined stall but develops an
excessive amount of induced drag. To be sure, it will not go unsaid that, an excessively low
angle of attack on takeoff creates its own problems i.e. excess takeoff speed and distance
and critical tyre loads.

(b) During Approach. During Approach the pilot must exercise proper technique to
control the flight path. "Attitude plus power equals performance.” The modern high speed
configuration at low speeds will have low lift-drag ratios due to the high induced drag and can
require relatively high power settings during the approach. If the pilot interprets that his
aeroplane is below the desired glide path, his first reaction must not be to just ease the nose
up. An increase in angle of attack without an increase in power will reduce the airspeed and
greatly increase the induced drag. Such a reaction could create a high rate of descent and
lead to very undesirable consequences. The angle of attack indicator, if available, provides
reference to the pilot and emphasizes that during the steady approach "angle of attack is the
primary control of airspeed and power is the primary control of rate of climb or descent".
Steep turns during approach at low airspeed are always undesirable in any type of aeroplane
because of the increased stall speed and induced drag. Steep turns at low airspeeds in a low
aspect ratio aeroplane can create extremely high induced drag and can incur dangerous sink
rates. The aircrew manual therefore prescribes limitations of a maximum permissible bank
angle, while turning onto base leg or finals, during a circuit to land.
FIS Book 1: Aerodynamics 154

(c) Landing Phase. During the landing phase an excessive angle of attack (or
excessively low airspeed) would create high induced drag and necessitate a high power
setting to control rate of descent. A common error in the technique of landing modern
configurations is a steep, low power approach to landing. The steep flight path requires
considerable manoeuvre (rotation) to flare out the aeroplane for touchdown and necessitates
a definite increase in angle of attack. Since the manoeuvre of the flare out is a transient
condition, the variation of both lift and drag with angle of attack must be considered. The lift
and drag curves for a high aspect ratio wing (Fig10-3) show a continued strong increase in CL
with α up to stall and large changes in CD only at the point of stall. These characteristics imply
that the high aspect ratio aeroplane is usually capable of flare out without unusual results.
The increase in angle of attack at flare out provides the increase in lift to change the flight
path direction without large changes in drag to decelerate the aeroplane. The lift and drag
curves for a low aspect ratio wing (Fig10-3) show that at high angles of attack the lift curve is
shallow, i.e., small changes in CL with increased α. This implies a large rotation needed to
provide the lift to flare out the aeroplane from a steep approach. The drag curve for the low
aspect ratio wing shows large, powerful increases in CD with CL well below the stall. These lift
and drag characteristics of the low aspect ratio wing create a distinct change in the flare out
characteristics. If a flare out is attempted from a steep approach at low airspeed, the
increased angle of attack may provide such increased induced drag and rapid loss of
airspeed that the aeroplane does not actually flare out. A possible result is that an even
higher sink rate may be incurred. This is one factor favoring the use of the "no flare out" or
"minimum flare out" type landing technique for certain modern configurations. These same
aerodynamic properties set the best glide speeds of low aspect ratio aeroplanes above the
speed for (L/D)max. The additional speed provides a more favorable margin of flare out
capability for flameout landing from a steep glide path (low aspect ratio, low (L/D)max, low
glide ratio).

13. Landing Technique. This must emphasize proper control of angle of attack and rate of
descent to prevent high sink rates and hard landings. As before, to be sure that it will not go unsaid,
excessive airspeed at landing creates its own problems i.e. of excessive wear and tear on tyres and
brakes, excessive landing distance, etc. The effect of the low aspect ratio planform of modern
aeroplanes emphasizes the need for proper flying techniques at low airspeeds. Excessive angles of
attack create enormous induced drag which can hinder takeoff performance and incur high sink rates
at landing. Since such aircraft have intrinsic high minimum flying speeds, an excessively low angle of
attack at takeoff or landing creates its own problems.

THE EFFECTS OF TAPER

14. The aspect ratio of a wing is the primary factor in determining the three-dimensional
characteristics of the ordinary wing and its drag due to lift. However, certain local effects take place
155 Wing Planforms

throughout the span of the wing and these effects are due to the distribution of area throughout the
span. The typical lift distribution is arranged in some elliptical fashion.

15. The natural distribution of lift along the span of the wing provides a basis for appreciating the
effect of area distribution and taper along the span. If the elliptical lift distribution is matched with a
planform whose chord is distributed in an elliptical fashion (the elliptical wing), each square foot of
area along the span produces exactly the same lift pressure. The elliptical wing planform then has
each section of the wing working at exactly the same local lift coefficient and the induced downflow at
the wing is uniform throughout the span. In the aerodynamic sense, the elliptical wing is the most
efficient planform because the uniformity of lift coefficient and downwash incurs the least induced drag
for a given aspect ratio. The merit of any wing planform is then measured by the closeness with
which the distribution of lift coefficient and downwash approach that of the elliptical planform.

16. The effect of the elliptical planform is illustrated in Fig 10-5 by the plot of local lift coefficient
cl / CL to wing lift coefficient, against semi-span distance. The elliptical wing produces a constant value
of cl / CL = 1.0 throughout the span from root to tip. Thus, the local section angle of attack, αo, and
local induced angle of attack, α1 are constant throughout the span. If the planform area distribution is
anything other than elliptical it may be expected that the local section and induced angles of attack will
not be constant along the span.

17. A planform previously considered is the simple rectangular wing which has a taper ratio of
1.0. A characteristic of the rectangular wing is a strong vortex at the tip with local downwash behind
the wing which is high at the tip and low at the root. This large non-uniformity in downwash causes
similar variation in the local induced angles of attack along the span. At the tip, where high downwash
exists, the local induced angle of attack is greater than the average for the wing. Since the wing angle
of attack is composed of the sum of α1 and αo, a large local α1 reduces the local αo creating low local
lift coefficients at the tip. The reverse is true at the root of the rectangular wing where low local
downwash exists. This situation creates an induced angle of attack at the root, which is less than the
average for the wing, and a local section angle of attack higher than the average for the wing. The
result is shown by the line B in the graph of Fig 10-5 which depicts a local coefficient at the root
almost 20% greater than the wing lift coefficient.

18. The effect of the rectangular planform may be appreciated by matching a near elliptical lift
distribution with a planform with a constant chord. The chords near the tip develop less lift pressure
than the root and consequently have lower section lift coefficients. The great non-uniformity of local
lift coefficient along the span implies that some sections carry more than their share of the load while
others carry less. Hence, for a given aspect ratio, the rectangular planform will be less efficient than
the elliptical wing. For example, a rectangular wing of A = 6 would have 16% higher induced angle of
attack for the wing and 5% higher induced drag than an elliptical wing of the same aspect ratio.
FIS Book 1: Aerodynamics 156

19. At the other extreme of taper is


the pointed wing which has a taper ratio of
zero. The extremely small area at the
pointed tip is not capable of holding the
main tip vortex at the tip and a drastic
change in downwash distribution results.
The pointed wing has greatest downwash
at the root and this downwash decreases
toward the tip. In the immediate vicinity of
the pointed tip an upwash is encountered
which indicates that negative induced
angles of attack exist in this area. The
resulting variation of local lift coefficient
shows low cl at the root and very high cl at
the tip. The effect may be appreciated by
realizing that the wide chords at the root
produce low lift pressures while the very
narrow chords towards the tip are subject
to very high lift pressures. The variation
of cl / CL throughout the span of the wing
of taper ratio 0 is shown as curve E in the
graph of Fig 10-5. As with the rectangular
wing, the non-uniformity of downwash and
lift distribution result in inefficiency of this Fig 10-5: Lift Distribution and Stall Patterns
planform. For example, a pointed wing of
A = 6 would have 17% higher induced angle of attack for the wing and 13% higher induced drag than
an elliptical wing of the same aspect ratio.

20. Between the two extremes of taper will exist planforms of more tolerable efficiency. The
variations of cl / CL for a wing of taper ratio 0.5 (curve C in the graph of Fig 10-5) are similar to the lift
distribution of the elliptical wing and the drag due to lift characteristics are nearly identical. A wing of
A = 6 and taper ratio = 0.5 has only 3% higher induced angle of attack and 1% greater CDi than an
elliptical wing of the same aspect ratio.

21. The elliptical wing is the ideal of the subsonic aerodynamic planform since it provides
minimum induced drag for a given aspect ratio. However, the major objection to the elliptical planform
is the extreme difficulty of mechanical layout and construction. A highly tapered planform is desirable
from the standpoint of structural weight and stiffness and the usual wing planform may have a taper
ratio from 0.45 to 0.20. Since structural considerations are important in the development of an
aeroplane, the tapered planform is a necessity for an efficient configuration. In order to preserve the
157 Wing Planforms

aerodynamic efficiency, the resulting planform is tailored by wing twist and section variation to obtain
as near as possible the elliptic lift distribution.

Stall Patterns

22. An additional effect of the planform area distribution is on the stall pattern of the wing. The
desirable stall pattern of any wing is a stall which begins at the root sections first. The advantages of
the root stall first are that ailerons remain effective at high angles of attack, favourable stall warning
results from the buffet on the tailplane and aft portion of the fuselage, and the loss of downwash
behind the root usually provides a stable nose-down moment to the aircraft. Such a stall pattern is
favoured but may be difficult to obtain with certain wing configurations. The types of stall patterns
inherent with various planforms are illustrated in Fig 10-5. The various planform effects are separated
as follows:

(a) Elliptical Planform. The elliptical planform has constant lift coefficients throughout
the span from root to tip. Such a lift distribution means that all sections will reach stall at
essentially the same wing angle of attack and the stall will begin and progress uniformly
throughout the span. While the elliptical wing would reach high lift coefficients before an
incipient stall, there would be little advance warning of a complete stall. Also, the ailerons
may lack effectiveness when the wing operates near the stall and lateral control may be
difficult.

(b) Rectangular Wing. The lift distribution of the rectangular wing exhibits low local lift
coefficients at the tip, and high local lift coefficients at the root. Since the wing will initiate the
stall in the area of highest local lift coefficients, the rectangular wing is characterized by a
strong root stall tendency. Of course, this stall pattern is favorable since there is adequate
stall warning buffet, adequate aileron effectiveness, and usually strong stable moment
changes on the aircraft. Because of the great aerodynamic and structural inefficiency of this
planform, the rectangular wing finds limited application only to low cost, low speed, light
planes.

(c) Moderate Taper Wing. The wing of moderate taper (taper ratio = 0.5) has a lift
distribution which is similar to that of the elliptical wing. Hence the stall pattern is much the
same as the elliptical wing.

(d) Highly Tapered Wing, The highly tapered wing (taper ratio = 0.25) shows the
stalling tendency inherent with high taper. The lift distribution of such a wing has distinct
peaks just inboard from the tip. Since the wing stall is started in the vicinity of the highest
local lift coefficient, this planform has a strong “tip stall” tendency. The initial stall is not
started at the exact tip but at the station inboard from the tip where the highest local lift
coefficients prevail.
FIS Book 1: Aerodynamics 158

(e) Pointed Tip Wing. The pointed tip wing (taper ratio = 0) develops extremely high
local lift coefficients at the tip. For all practical purposes the pointed tip will be stalled at any
condition of lift unless extensive tailoring is applied to the wing. Such a planform has no
practical application to a subsonic aircraft.

(f) Swept Back Wing. The effect of sweepback on the lift distribution of a wing is
similar to the effect of reducing the taper ratio. The full significance of sweepback is
discussed in the following paragraphs.

SWEEPBACK

Swept-Back Leading Edges

23. This type of planform is used on high speed aircraft and may take the form of a sweptback
wing, or a delta, with a tailplane (MiG 21) or without it (Mirage 2000) a tailplane. The reason for the
use of these planforms is their low drag at the higher speeds. The chapter on Design for High Speed
Aerodynamics deals fully with this aspect. However, the high speed / low drag advantages are gained
at the cost of a poorer performance at the lower end of the speed range.

Effect of Sweepback on Lift

24. If a straight wing is changed to a swept


planform, with similar parameters of area, aspect
ratio, taper, section and washout, the CL max is
reduced. This is due to premature flow separation
from the upper surface at the wing tips. For a sweep
angle of 45°, the approximate reduction in CL max is
around 30%. Fig 10-6 shows typical CL curves for a
straight wing, a simple swept-back wing, and a
tailless delta wing of the same low aspect ratio. The
critical angle of attack of the MiG 21 aircraft is about
300 to 330 while that of the MiG 27 with wing sweep
160 is 180 to 210 and with sweep 720 it is 340 to 370. Fig 10-6: Effect of Sweepback on CL

25. A swept wing presents less camber and a greater fineness ratio to the airflow. However, the
reasons for the lowering of CL slope are more readily apparent from an examination of Figs 10-7 and
10-8. From Fig 10-7 it can be seen that the velocity V of the RAF can be divided into two
components, V1 parallel to the leading edge which has no effect on the lift, and V2 normal to the
leading edge which does affect the lift and is equal to V cos Λ. Therefore, all other factors being
equal, the CL of a swept wing is reduced in the ratio of the cosine of the sweep angle.
159 Wing Planforms

26. Fig 10-8 shows that an increase in fuselage


angle of attack ∆α will only produce an increase in the
angle of attack ∆α cos Λ in the plane perpendicular to
the wing quarter chord line. Since we have already
said that it is the airflow in the latter plane which
effects CL, the full increment of lift expected from the
∆α change is reduced to that of a ∆α cos Λ change.

27. Considering Fig 10-6, the stall occurs on all


three wings at angles of attack considerably greater
than those of wings of medium and high aspect ratios.
On all aircraft it is desirable that the landing speed
should be close to the lowest possible speed at which Fig 10-7: Flow Velocities on a Swept
Wing
the aircraft can fly. To achieve this desirable minimum
the wing must be at the angle of attack corresponding
to the CL max.

28. On all wings of very low aspect ratio, and


particularly on those with a swept-back planform, the
angles of attack giving the highest lift coefficients
cannot be used for landing. This is because, as
explained later, swept-back planforms have some
undesirable characteristics near the stall and because
the exaggerated nose-up attitude of the aircraft
necessitates, among other things, excessively long
and heavy undercarriages. The maximum angle at
which an aircraft can touch down without recourse to
Fig 10-8: Effect of Change in
such measures is about 15°, and the angle of attack at Angel of Attack
touchdown will therefore have to be something of this
order. Fig 10-6 shows that the CL corresponding to this angle of attack is lower than the CL max for
each wing.

29. Compared with the maximum


usable lift coefficient available for
landing aircraft with unswept wings,
those of the swept and delta wings are
much lower, necessitating higher
landing speeds for a given wing
Fig 10-9: Planform Areas Giving a
loading. It is now apparent that, to Common Stalling Speed
obtain a common minimum landing
FIS Book 1: Aerodynamics 160

speed at a stated weight, an unswept wing needs a smaller area than either of the swept planforms.
The simple swept wing needs a greater area, and so a lower wing loading, in order that the reduced
CL can support the weight at the required speed. The tailless delta wing needs still more area, and so
a still lower wing loading, to land at the required speed. Fig 10-9 shows typical planforms for the three
types of wing under consideration, with the areas adjusted to give the same stalling speed. The much
larger area of the delta wing is evident.

Effect of Sweepback on Drag

30. The main reason for employing sweepback as a wing planform is to improve the high speed
characteristics of the wing. Unfortunately this has adverse effects on the amount of drag produced at
the higher range of angles of attack. The induced drag increases approximately in proportion to 1 /
Cos Λ. This is because, as already explained, by sweeping the wing CL is reduced, and therefore to
maintain the same lift the angle of attack has to be increased. This increases the induced downwash
and hence the induced drag.

31. The practical significance of this high increase in drag is the handling problems it imposes
during an approach to landing. Because of the greater induced drag, the minimum drag speed is
higher than for a comparable straight wing, and the approach speed is usually less than the minimum
drag speed. Therefore, if a pilot makes a small adjustment to the aircraft’s attitude, for example, by
raising the nose slightly, the lift will be increased slightly, but there will be a large increase in drag
which will result in a rapid reduction in speed, and a large increase in power to restore equilibrium. In
fact, the stage may be reached where the use of full power is insufficient to prevent the aircraft from
descending rapidly.

32. On some aircraft this problem is overcome


by employing high drag devices such as airbrakes
or drag-chutes to increase the zero lift drag. This
results in a flatter drag curve with the minimum
drag speed closer to the approach speed (Fig 10-
10). A further advantage is that more power is
required on the approach, which on turbojet
aircraft, means better engine response.

Fig 10-10: Improvement in Approach


Effect of Sweepback on Stalling Speed Stability

33. When a wing is swept back, the boundary layer tends to change direction and flow towards
the tips. This outward drift is caused by the boundary layer encountering an adverse pressure
gradient and flowing obliquely to it over the rear of the wing.
161 Wing Planforms

34. The pressure distribution on a swept wing is shown


by isobars in Fig 10-11. The velocity of the flow has been
shown by two components, one at right angles and one
parallel to the isobars. Initially, when the boundary layer
flows rearwards from the leading edge it moves towards a
favorable pressure gradient, i.e. towards an area of lower
pressure. Once past the lowest pressure however, the
component at right angles to the isobars encounters an
adverse pressure gradient and is reduced. The component
parallel to the isobars is unaffected, thus the result is that
the actual velocity is reduced (as it is over an unswept
wing) and also directed outwards towards the tips. Fig 10-11: Outflow of Boundary
Layer

35. The direction of the flow continues to be changed until the component at right angles to the
isobars is reduced to zero, whilst the parallel component, because of friction, is also slightly reduced.
This results in a “pool” of slow moving air collecting at the tips.

36. The spanwise drift sets up a tendency towards tip stalling, since it thickens the boundary layer
over the outer parts of the wing and makes it more susceptible to separation, bringing with it a sudden
reduction in CL max over the wing tips.

37. At the same time as the boundary layer is flowing towards the tips, at high angles of attack,
the airflow is separating along the leading edge. Over the inboard section it re-attaches behind a
short “separation bubble”, but on the outboard section it re-attaches only at the trailing edge or fails to
attach at all. The separated flow at the tips combines with the normal wing tip vortices to form a large
vortex (the ram’s horn vortex). The factors which combine to form this vortex are:

(a) Leading edge separation.


(b) The flow around the wing tips.
(c) The spanwise flow of the boundary
layer.

These factors are illustrated in Fig 10-12, and the


sequence of the vortex development and its effect on
the airflow over the wing is shown in Fig 10-13. From
Fig 10-13 it can be seen that the ram’s horn vortex
has its origin on the leading edge, possibly as far
inboard as the wing root. The effect of the vortex on
the air above it (the external flow) is to draw the latter
down and behind the wing, deflecting it towards the
Fig 10-12: Vortex Development
fuselage (Fig 10-14).
FIS Book 1: Aerodynamics 162

38. The spanwise flow of the boundary


layer increases as angle of attack is
increased. This causes the vortex to
become detached from the leading edge
closer inboard (Fig 10-15). As a result,
outboard ailerons suffer a marked decrease
in response with increasing angle of attack.
This, in turn, means that comparatively large
aileron movements are necessary to
manoeuvre the aircraft at low speeds and
the aircraft response may be Fig 10-13: Formation of Ram’s
correspondingly sluggish. This effect may Horn Vortex
be countered by limiting the inboard
encroachment of the vortex as described
below, or by moving the ailerons inboard.
Another possible solution is the use of an
all-moving wing tip.
Fig 10-14: Influence on External Flow

Alleviating the Tip Stall

39. Most of the methods used to


alleviate the tip stall aim either at
maintaining a thin and therefore strong
boundary layer, or re-energizing the
weakened boundary layer (See also the
chapter on Lift Augmentation).
Fig 10-15: Shift of Ram’s Horn Vortex

(a) Boundary Layer Fences. Used originally to restrict the boundary layer out-flow,
fences also check the spanwise growth of the separation bubble along the leading edge (see
para 36).

(b) Leading Edge Slots. These have the effect of re-energizing the boundary layer
(see also the chapter on lift augmentation).

(c) Boundary Layer Suction. Suitably placed suction points draw off the weakened
layer and a new high-energy layer is then drawn down to take its place (see also the chapter
on lift augmentation).

(d) Boundary Layer Blowing. High velocity air is injected into the boundary layer to
increase its energy (see also the chapter on lift augmentation).
163 Wing Planforms

(e) Vortex Generators. These re-energize the boundary layer by making it more
turbulent. The increased turbulence results in high-energy air in layers immediately above the
retarded layer being mixed in and so re-energizing the layer as a whole. Vortex generators
are most commonly fitted ahead of control surfaces to increase their effect by speeding up
and strengthening the boundary layer. Vortex generators also markedly reduce shock-
induced boundary layer separation, and reduce the effects of the upper surface shockwave
(explained in chapters of lift augmentation and high speed aerodynamics).

(f) Leading Edge Extension. Also known as a “sawtooth” leading edge, the extended
leading edge is a common method used to avoid the worst effects of tip stalling. The effect of
the extension is to cut down the growth of the main vortex. A further smaller vortex, starting
from the tip of the extension, affects a much smaller proportion of the tip area and in lying
across the wing, behind the tip of the extension, it has the effect of restricting the outward flow
of the boundary layer. In this way the severity of the tip stall is reduced and with it the pitch-
up tendency. Further effects of the leading edge extension are:

(i) The t/c ratio of the tip area is reduced, with consequent benefits to the critical
Mach number.
(ii) The Centre of Pressure (CP) of the extended portion of the wing lies ahead of
what would be the CP position if no extension were fitted. The mean CP position for
the whole wing is therefore further forward and, when the tip eventually stalls, the
forward shift in CP is less marked, thus reducing the magnitude of the nose-up
movement.

(g) Leading Edge Notch. The notched leading edge has the same effect as the
extended leading edge insofar as it causes a similar vortex formation thereby reducing the
magnitude of the vortex over the tip area and with it the tip stall. Pitch-up tendencies are
therefore reduced. The leading edge notch can be used in conjunction with extended leading
edge, the effect being to intensify the inboard vortex behind the devices to create a stronger
restraining effect on boundary layer out-flow. The choice whether to use either or both of
these devices lies with the designer and depends on the flight characteristics of the aircraft.

Pitch-Up

40. Longitudinal Instability. Longitudinal instability results when the angle of attack of a swept
wing increases to the point of tip stall. The instability takes the form of a nose-up pitching moment,
called pitch-up, and is a self-stalling tendency in that the angle of attack continues to increase once
the instability has set in. The aerodynamic causes of pitch-up are explained in detail in the following
paragraphs.
FIS Book 1: Aerodynamics 164

41. Centre of Pressure (CP) Movement. When the swept-back wing is unstalled, the CP lies
in a certain position relative to the CG, the exact
position being the mean of the centres of
pressure for every portion of the wing from the
root to the tip. When the tip stalls, lift is lost over
the outboard sections and the mean CP moves
rapidly forward. The wing moment (Fig 10-16) is
reduced and a nose-up pitching moment results Fig 10-16: Nose-up Pitching Moment
which aggravates the tendency. Resulting from Tip Stalling

42. Change of Downwash over the Tailplane. Fig 10-17 shows that the maximum downwash
from the swept-back wing in
unstalled flight comes from
the tip portions. This is to be
expected since the CL is
highest over these parts of
the wing. When the wing tips
stall, effective lift production
is concentrated inboard and
the maximum downwash now
operates over the tailplane
and increases the tendency
to pitch up. This effect can
be reduced by placing the
tailplane as low as possible in
line with, or below, the wing
chord line, so that it lies in a
Fig 10-17: Variation of Downwash
region in which downwash
changes with angle of attack are less marked.

43. Washout Due to Flexure. When a swept wing


flexes under load, all chordwise points at right angles to the
main spar are raised to the same degree, unless the wing is
specially designed so that this is not so. Thus in Fig 10-18,
the points A and B rise through the same distance and the
points C and D rise through the same distance but through
a greater distance than A and B. Thus C rises further than
A and there is a consequent reduction in incidence at this
section. This aeroelastic effect is termed, ‘washout due to
flexure’, and is obviously greatest at the wing tips. It is Fig 10-18: Washout Due to
Flexure
165 Wing Planforms

most noticeable during high g manoeuvres when the loss of lift at the tips and the consequent forward
movement of the centre of pressure causes the aircraft to tighten up in the manoeuvre. A certain
amount of washout due to flexure is acceptable provided the control in pitch is adequate to
compensate for it, but it can be avoided by appropriate wing design.

44. Pitch-Up on Aircraft with Straight Wings. On aircraft with low aspect ratio, short-span
wings, pitch-up can be caused by the effect of the wing tip vortices. As the angle of attack is
increased the vortices grow larger until at or near the stall they may be large enough to affect the
airflow over the tailplane. As each vortex rotates inwards towards the fuselage over its upper half, the
tailplane angle of attack is decreased giving rise to a pitch-up tendency (Fig 10-17).

45. Rate of Pitch-Up. From the pilot’s point of view, pitch-up is recognized when the pull force
on the control column which is being applied to the aircraft near the stall has to be changed to a push
force to prevent the nose from rising further. The more the speed decreases, the further forward must
the control column be moved to restrain the nose-up pitch. Pitch-up in level flight or in any 1g stall is
usually gentle, since the rate at which the stall is spreading is comparatively slow and is usually
accompanied by the normal pre-stall buffeting. When the stall occurs in a manoeuvre, under g, the
onset of pitch-up can be violent and sudden, corresponding to the rate of spread of the stall.

The Crescent Wing

46. The crescent wing planform combines variable sweep with a changing thickness/chord ratio.
At the root section where the wing is thickest, the angle of sweep is greatest. As the t/c ratio is
reduced spanwise, so is the angle of sweep, so that the outboard sections are practically unswept.
Hence there is little or no outflow of the boundary layer at the tips. The advantages of the crescent
wing are:

(a) The critical drag rise Mach number is raised.


(b) The peak drag rise is reduced.
(c) Because of the lack of outflow of the boundary layer at the tips, tip stalling is
prevented.

FORWARD SWEEP

General

47. The benefits of wing sweep can be achieved by sweeping the wing backwards or forwards,
yet only in recent years has the forward swept wing (FSW), become a serious alternative to
sweepback. The reason for this lies in the behavior of wing structures under load (Para 43).
FIS Book 1: Aerodynamics 166

48. The main advantages lie in the subsonic / transonic regime. Taking the 70% chordline as the
average position for a shock-wave to form as the critical Mach number is approached, the sweep
angle of this chordline influences wave drag.

49. The FSW can maintain the same chord-line sweep as the swept-back wing (SBW) but due to
a geometric advantage, achieves this with less leading edge sweep and enjoys the advantages
accruing from this subsonically.

50. The decision to employ FSW or SBW will depend inter alia, on the speed regime envisaged
for the design. Due to better lift/drag ratio in the subsonic and near transonic speed range, fuel
consumption is improved over the SBW. For a high speed supersonic interception the higher
supersonic drag is a disadvantage.

Wing Flexure

51. Under flexural load the airflow sees a steady increase in effective angle of attack from root to
tip, the opposite effect to aft-sweep. Under g loading, lift increases at the tips, leading to pitch-up as
the centre of pressure moves forwards. Additionally, the increased angle of attack at the tips now
leads to increased wing flexure, which leads to increased effective angle of attack at the tips. The
result of this aero-elastic divergence is likely to be structural failure of the wing, so it is not surprising
that sweepback was considered to be a better option until recently. What changed the situation was
the development of carbon fiber technology, which can produce controlled wing twist under load, such
that the effect described is eliminated.

Vortex Generation

52. Fig 10-19 shows the difference


in ram’s horn vortex behavior. In the
swept forward design the ram’s horn
vortex develops inwards towards the
root, not outwards towards the tips.

53. There will, of course, still be Fig 10-19: Comparison of Ram’s Horn
vortices from the wing tips, but these Vortex Behaviour

no longer reinforce and aggravate the ram’s horn vortex, which now lies along the fuselage, or slightly
more outboard if a small section of the wing root is swept back.

Stalling

54. A swept forward wing will tend to stall at the root first. This stall can be controlled in a number
167 Wing Planforms

of ways. Since a conventional tailplane would tend to lie in a vortex, a popular option is to combine
sweep forward with a canard foreplane. Downwash from a carefully placed canard can delay root
stall, (see Para 71), and even the vortices from the canard can be used to energize the airflow over
inboard sections of the wing, maintaining lift to higher angles of attack.

55. The root-stall characteristics give better lateral control at the stall as aileron control is
retained, but may incur a penalty in directional control as the fin and rudder are acting in the chaotic
turbulence from the root separation.

DELTA WINGS

Tailless Delta

56. On aircraft using


this type of wing e.g.
Mirage 2000 (Fig 10-20)
the angle of attack is
controlled by movement of
the trailing edge of the
wing. An upward
movement produces a
downward force on the
trailing edge and so
increases the angle of
attack. When compared Fig 10-20: Tailless Delta Planform
with an identical wing
which uses a separate tailplane to control the angle of attack, the tailless delta reveals two main
differences:

(a) The CL max is reduced.


(b) The stalling angle is increased.

Reduction of CL MAX

57. The chord line of the wing is defined as being a straight line joining the leading edge to the
trailing edge. If a given wing/aerofoil combination has a hinged trailing edge for use as an elevator,
then it can be said that when the trailing edge is in any given angular position the effective aerofoil
section of the wing has been changed.

(a) When such a wing reaches its stalling angle in level flight, the trailing edge elevator
must be raised to impose a downward force on the trailing edge to maintain the wing at the
FIS Book 1: Aerodynamics 168

required angle of attack. The raised trailing edge has two effects viz. it deflects upwards the
airflow passing over it and so reduces the downwash, the amount of which is proportional to
the lift, and it reduces the extent of
both the low-pressure area over the
upper surface of the wing and the
high-pressure area below, thereby
lowering the CL.

(b) The curves of Fig 10-21 show


that a section with a raised trailing
edge will suffer a decreased CL max

compared to the basic section. Fig 10-21: Effect of Hinged Trailing Edge on
CL max and Stalling Angle

Increase in Stalling Angle

58. The planform of the delta wing gives it an inherently low aspect ratio and therefore a high
stalling angle and a marked nose-up attitude at the stall in level flight. If a certain delta wing is used
without a tailplane, i.e. the trailing edge is used as an elevator, then the stalling angle is higher than
when the same wing is used in conjunction with a tailplane.

59. All else being equal


(planform, aspect ratio, area,
etc.) changes in the amount of
camber (by altering the
angular setting of the trailing
edge elevator) do not affect
the stalling angle appreciably. Fig 10-22: Comparison of Stalling Angle

That is, the angle between the chord line (drawn through the leading and trailing edges) and the
direction of the airflow remains constant when at maximum CL irrespective of the setting of the hinged
trailing edge. Fig 10-22 illustrates this point and it can be seen that for both the “tailed” and “tailless”
aircraft the stalling angle is the same when measured on the foregoing principles.

60. However, it is normal practice and convention to measure the stalling angle with reference to
the chord line obtained when the moveable trailing edge is in the neutral position, and not to assume
a new chord line with each change in trailing edge movement. When the stalling angle is measured
with reference to the conventional fixed chord line, it can be seen from Fig 10-22 that the angle is
greater. Fig 10-22 also shows that, because the wing proper is set at a greater angle at the stall when
a trailing edge elevator is used, the fuselage attitude is more nose-up, giving a more exaggerated
attitude at stall in level flight.
169 Wing Planforms

61. Since it is easier to refer to angle of attack against a fixed chord line, the basic chord line is
always used as the reference datum. This convention is the reason for the apparently greater stalling
angles of tailless delta wings. It is perhaps a more realistic method, as the pilot is invariably aware of
the increased attitude of his aircraft relative to the horizontal, but is not always aware of increases in
the angle of attack.

The CL Curve

62. Reference to Fig 10-6 shows that the peak of the curve for the lift coefficient is very flat and
shows little variation of CL over a comparatively wide range of angles. This very mild stalling
behaviour enables the delta wing to be flown at an angle of attack considerably higher than that of the
CL max, possibly with no ill effects other than the very marked increase in drag. The flat peak denotes
a gradual stall, with a consequent gradual loss of lift as the stalling angle is exceeded.

The Slender Delta

63. The slender delta provides low drag at supersonic speeds because of its low aspect ratio.
This, combined with a sharp leading edge, produces leading edge separation at low angles of attack.
Paradoxically, this is encouraged. Up to now the vortex so produced has been an embarrassment as
it is unstable, varies greatly with angle of attack, causes buffet, increases drag and decreases CL max.
By careful design, the vortex can be controlled and used to advantage.

64. Vortex Lift. The vortex on a slender delta is different in character from that on a wing of
higher aspect ratio (greater than 3). On the slender delta the vortex will cover the whole leading edge
from root to tip, rather than start at the tip and travel inwards at higher angles of attack. Its behaviour
is therefore more predictable, and, as it is present during all aspects of flight, the following
characteristics may be exploited:

(a) Leading edge flow separation causes CP to be situated nearer mid-chord. Hence
there will be less difference between CP subsonic and CP supersonic than before, and
longitudinal stability is improved. (Supersonically the CP moves backwards to about 50%
chord.)

(b) The vortex core is a region of low pressure, therefore an increase in CL may be
expected. On the conventional delta this cannot be utilized as the vortex seldom approaches
anywhere near the wing root and most of its energy appears in the wake behind the wing,
where it produces high induced drag. On the slender delta the low pressure in the vortex is
situated above the wing and can result in an increase in CL of as much as 30% under
favourable conditions.
FIS Book 1: Aerodynamics 170

VARIABLE GEOMETRY WINGS

General

65. An aircraft which is designed to fly at supersonic speeds most of the time usually has poor
low speed characteristics which have to be accepted, although various high lift devices are available
for reducing take-off and landing speeds and improving the low speed handling qualities. In order to
achieve the desired high speed
performance, the aircraft has thin
symmetric wing sections and highly
swept or delta wing planforms. These
wings are very inefficient at low speeds
where unswept wing planforms and
cambered wing sections are required.

66. In the case of an aircraft which is


required to be operated efficiently at both
high and low speeds, variable wing
sweep is a desirable feature to be
incorporated in the design. The wings
Fig 10-23: Variable Geometry Wings
can then be swept back when the aircraft
is being flown at high speeds and forward again when flying at low speeds e.g. Mig 23 & Mig 27
aircraft (Fig 10-23).

Stability and Control Problems

67. When the wing of an aircraft is moved backwards the aerodynamic centre moves rearwards.
The CG of the aircraft also moves back at the same time, but, since most of the weight of an aircraft is
concentrated in the fuselage, the movement is less than that of the aerodynamic centre. The
rearwards movement of the aerodynamic centre produces a nose-down change of trim and an
increase in the longitudinal static stability of the aircraft. Additional up-elevator is required to trim the
aircraft and this results in additional drag called “trim drag”. This extra drag can be a relatively large
part of the total drag of an aircraft at supersonic speeds and it is essential that it should be kept as
small as possible. Various design methods are available for reducing or eliminating the trim changes
produced by sweeping the wings.

68. Wing Translation. The aerodynamic centre can be moved forward again by translating the
wing forwards as it is swept back. This method involves extra weight and structural complications.
171 Wing Planforms

69. CG Movement. The aircraft can be designed so that the CG moves rearwards in step with
the aerodynamic centre by mounting some weight in the form of engines, etc at the wing tips. As
engines would have to swivel to remain aligned with the airflow, additional weight and other
complications result. Another possible method of moving the CG is by transferring fuel to suitable trim
tanks in the rear fuselage.

70. Leading Edge Fillet and Pivot Position. Another solution can be obtained by positioning
the pivot point outboard of the fuselage inside a fixed, leading edge fillet, called a ‘glove’. The
optimum pivot position for minimum movement of the CP depends on the wing planform, but it is
usually about 20% out
along the mid-span.
However, the fixed
glove-fairing presents
a highly swept portion
of the span at low-
speed, forward sweep
settings. This incurs
the undesirable
penalties that variable
geometry is designed
to overcome. A
compromise between
sweeping the whole Fig 10-23: Movement of CP

wing and a long glove giving the minimum CP shift, is usually adopted (Fig 10-23).

CANARD WINGS

Advantages and Disadvantages

71. A canard-type configuration is one which has a fore-plane located forward of the wing instead
of, in the more conventional, aft position like Su 30 MK I. On an aircraft with a long slender fuselage
with engines mounted in the tail and a CG position well aft, this layout has the obvious geometric
advantage of a longer moment arm. This enables the stability and trim requirements to be satisfied by
a fore-plane of smaller area. The trim drag problem will also be reduced because, at high speeds, an
up-load will be required on the fore-plane to trim the aircraft. There are, however, certain
disadvantages with this layout:

(a) Stalling Problems. On a “conventional” rear-tailplane configuration, the wing stalls


before the tailplane, and longitudinal control and stability are maintained at the stall. On a
canard layout, if the wing stalls first, stability is lost, but if the fore-plane stalls first then control
FIS Book 1: Aerodynamics 172

is lost and the maximum value of CL is reduced. One possible solution is to use a canard
surface and a wing trailing edge flap in combination, with one surface acting as a trimming
device, and the other as a control. Alternatively, an auxiliary horizontal tailplane at the rear
may be used for trim and control at low speed.

Fig 10-24: Canard Design of the SU-30 aircraft

(b) Interference Problems. In the same way as the airflow from the wing interferes
with the tail unit on the conventional rear-tail layout, so the airflow from the fore-plane
interferes with the flow around the main wing and vertical fin in a canard layout. This can
cause a reduction in lift on the main wing, and can also result in stability problems. The
interference with the vertical fin can cause a marked reduction in directional static stability at
high angles of attack. The stability may be improved by employing twin vertical fins in place
of the single control vertical fin.

ENDPLATES AND WINGLETS

72. While the high aspect ratio wing will minimize induced drag, long, thin wings increase
structural weight and have relatively poor stiffness characteristics. Since a limit will be reached where
it is no longer practical to increase the aspect ratio of the wing, alternative methods have been
researched that can reduce wing tip vortices. These methods are used to create an effective aspect
ratio, which is greater than the physical aspect ratio. Since the induced drag is created by the cross-
flow at the wing tips, anything that can be done to reduce this flow will help in reducing the drag. One
method employed on many aircraft is modifying the shape of the wing tips.
173 Wing Planforms

73. One of the most effective designs


is a drooped wing tip, Fig 10-25. Drooped
wing tips give the outward flow of air a
downward direction as it attempts to flow
around the tip. This downward movement
of air results in the vortex being formed
farther out board than it would otherwise
be, in effect increasing the aspect ratio of
the wing.

74. Another method that has been


used on several aircraft is the installation of
an endplate. The endplate has had only Fig 10-25: Drooped wing tips and Wing lets

moderate success, with the result in many installations being that the reduction in induced drag is
offset by a corresponding increase in parasitic drag. A variation of the endplate concept is the use of
fuel tanks on the wing tips.

75. One of the most successful attempts at increasing the effective aspect ratio has resulted in
the development of a device called a winglet, shown in Fig 10-25, which looks like a small jib sail
mounted at and above the wing tips. The inward flow of air from the wingtip vortex acts on the winglet,
much like wind acting on a jib sail, and produces a forward component of lift. A winglet also reduces
the cross flow on the wing, which reduces the trailing vortex in strength. The reduction of wingtip
vortices causes a corresponding reduction in drag, thereby increasing the wings’ lift/drag ratio. A
winglet, however, also has some disadvantages. It will add weight to the aircraft and produce parasitic
drag. Winglets are most effective at high lift co-efficient.

SUMMARY

Planform Considerations

76. Planform is the geometrical shape of the wing when viewed from above, and it largely
determines the amount of lift and drag obtainable from a given area, it also has a pronounced effect
on the stalling angle of attack.

77. Aspect ratio (AR) is found by dividing the square of the wing span by the area of the wing:

A = Span2 = ____Span___
Area Mean Chord

78. The following wing characteristics are affected by aspect ratio:

(a) Induced drag is inversely proportional to aspect ratio.


FIS Book 1: Aerodynamics 174

(b) The reduced effective angle of attack of very low aspect ratio wings can delay the
stall considerably. (Some delta wings have no measurable stalling angle up to 40°).

79. In the aerodynamic sense, the elliptical wing is the most efficient planform because the
uniformity of lift coefficient and downwash incurs the least induced drag for a given aspect ratio.

80. Any swept-back planform suffers a marked drop in CL max when compared with an unswept
wing with the same significant parameters, also the boundary layer tends to change direction and flow
towards the tips.

81. The spanwise drift of the boundary layer sets up a tendency towards tip stalling on swept
wing aircraft. This may be alleviated by the use of one or more of the following:

(a) Boundary layer fences.


(b) Leading edge slots.
(c) Boundary layer suction.
(d) Boundary layer blowing.
(e) Vortex generators.
(f) Leading edge extension.
(g) Leading edge notch.

82. The factors effecting pitch-up are:

(a) Centre of pressure movement.


(b) Change of downwash over the tail-plane.
(c) Washout due to flexure.

83. The advantages of a crescent wing are:

(a) The critical drag rise Mach number is raised.


(b) The peak drag rise is reduced.
(c) Because of the lack of outflow of the boundary layer at the tips, tip-stalling is
prevented.

84. A FSW stalls at the root first, prolonging aileron control. The configuration may offer an
advantage in L/D ratio over sweepback in the appropriate speed range.

85. When compared with a delta which uses a separate tailplane to control angle of attack, the
tailless delta reveals two main differences:

(a) The CL max is reduced.


(b) The stalling angle is increased.
175 Wing Planforms

86. Vortex lift has the following characteristics:

(a) Leading edge flow separation causes the CP to be situated nearer to midchord.
(b) The vortex core is a region of low pressure, therefore an increase in CL may be
expected.

87. The canard configuration has the following advantages and disadvantages:

(a) Advantages:

(i) The control surface is ahead of any shocks which may form on the
mainplane.
(ii) On an aircraft with a long slender fuselage with engines mounted in the tail
and the CG position well aft, the fore-plane has the advantage of a long moment arm.
(iii) The stability and trim requirements can be satisfied with a smaller fore-plane
area.
(iv) Because up-loads will be required, the trim drag problem is reduced.

(b) Disadvantages:

(i) If the wing stalls first stability is lost.


(ii) If the fore-plane stalls first control is lost.
(iii) In the same way as the airflow from the wing interferes with the tail unit on
the conventional rear-tail layout, so the airflow from the fore-plane interferes with the
flow around the main wing and vertical fin of the canard configuration.
FIS Book 1: Aerodynamics 176
177

CHAPTER 11

LIFT AUGMENTATION

Introduction

1. Technology advancements have made modern day aircraft smaller and sleeker, but the
requirement to carry more loads (weapons, passengers, etc.) has also increased over the years. Thus
the wing loading (W/S), which represents the lift required in level flight per unit area of wing, has
increased. The same can be seen with the help of the lift equation:

2
L = CL ½ ρV S
2
W = L = CL ½ ρ V S (in straight and level flight)
2
Therefore W = L = CL max ½ ρ VB Where VB = BSS
S S

2. From the above equation it is clear that to balance certain value of W/S or L/S for a given
aircraft, we can either increase the value of CL max or increase VB. Increase in the basic stalling speed
would entail an increase in the landing speed and that would bring about the requirement of longer
runways, effective brake systems, tail chutes, arrester barriers, etc. An alternate and better way
therefore was to increase the value of CL max. This concept led to the development of Lift
Augmentation devices, which are primarily used to increase the max lift coefficient, in order to reduce
the airspeed at unstick and touchdown.

3. The chief devices used to augment CL max are:

(a) Slats. These are auxiliary aerodynamic surfaces, which are fitted along the leading
edge of the wing. These are invariably used in conjunction with slots. The main purpose
being to delay stall and improve low speed handling characteristics of aircraft (generally high
speed design aircraft). Their operation can either be automatic or manually controlled by the
pilot.

(b) Flaps. These are auxiliary aerodynamic surfaces, which augment CL max primarily
by changing the camber of the wing and also in some cases increasing the wing area. These
are generally fitted along the wing trailing edge, but in some cases are fitted along the leading
edge also.

(c) Boundary Layer Control. These are high lift devises used to energise the
boundary layer so as to delay separation of boundary layer thus delaying stall. This is
achieved either by boundary layer suction or by blowing air or by fitting vortex generators.
FIS Book 1: Aerodynamics 178

SLATS AND SLOTS

Principle of Operation

4. When a small auxiliary aerofoil slat of


highly cambered section is fixed to the leading
edge of a wing along the complete span, and
adjusted so that a suitable slot is formed
between the two, the CL max may be increased by
as much as 70%. At the same time, the stalling
angle of attack is increased by some 10°. The
graph at Fig 11-1 shows the comparative figures
for a slatted and unslatted wing of the same
basic dimensions.

5. The effect of the slat is to prolong the lift


curve by delaying the stall until a higher angle of
attack. When operating at high angles of attack
the slat itself is generating a high lift coefficient
because of its marked camber. The action of Fig 11-1: Effects of Flaps and Slats on Lift
the slat is to flatten the marked peak of the low-
pressure envelope at high angles of attack and
to change it to one with a more gradual pressure
gradient. The flattening of the lift distribution
envelope means that the boundary layer does
not undergo the sudden thickening that occurred
through having to negotiate the very steep
gradient that existed immediately behind the
former suction peak it therefore retains much of
its energy, thus enabling it to penetrate almost
the full chord of the wing before separating. Fig Fig 11-2: Effect of Slats on Pressure
Distribution
11-2 shows the alleviating effect of the slat on
the low-pressure peak and that, although flatter, the area of the low pressure region, which is
proportional to its strength, is unchanged or even increased. The passage of the boundary layer over
the wing is assisted by the fact that the air flowing through the slot is accelerated by the venturi effect,
thus adding to the kinetic energy of the boundary layer and so helping it to penetrate further against
the adverse pressure gradient.

6. As shown in Fig 11-1, the slat delays separation until an angle of about 25° to 28° is reached,
during which time the lift coefficient has risen steadily, finally reaching a peak considerably greater
179 Lift Augmentation

than that of an unslatted wing. Assuming that the CL max of the wing is increased by, say 70%, it is
evident that the stalling speed at a stated wing loading can be much reduced, for example, if an
unslatted wing stalls at a speed of about 200 kmph, its fully slatted counterpart would stall at about
160 kmph. The exact amount of the reduction achieved depends on the length of the leading edge
covered by the slat and the chord of the slat. In cases where the slats cover only the wing tips, the
increase in CL is proportionately smaller.

Automatic Slats

7. Since the slat is of use only at high


angles of attack, at the normal angles its
presence serves only to increase drag. This
disadvantage can be overcome by making the
slat moveable so that when not in use it lies
flushed against the leading edge of the wing as
shown in Fig 11-3. In this case the slat is
hinged on its supporting arms so that it can
move to either the operating position or the
closed position at which it gives least drag. This
type of slat is fully automatic in that its action Fig 11-3: Automatic Slat

needs no separate control.

8. At high angles of attack the lift of the slat raises it clear of the wing leaving the required slot
between the two surfaces which accelerates and re-energizes the airflow over the upper surface.
Slats which are movable but do not make use of the above principle have some different
arrangement. These slats may be operated hydraulically and the control to select them ‘out’ or ‘in’ is
given automatically by sensing the angle of attack the aircraft is at, thereby achieving the same result.

Uses of Slat

9. On some high performance aircraft the purpose of slats is not entirely that of augmenting the
CL max since the high stalling angle of the wing with a full-span slat necessitates an exaggerated and
unacceptable landing attitude if the full benefit of the slat is to be obtained. When slats are used on
these aircraft, their purpose may be as much to improve control at low speeds by curing any
tendencies towards wing tip stalling, as it is to augment the lift coefficient.

10. If the slats are small and the drag negligible they may be fixed, i.e. non-automatic. Large
slats are invariably of the automatic type. Slats are often seen on the leading edges of sharply swept-
back wings. On these aircraft the slats usually extend along most of the leading edge and besides
FIS Book 1: Aerodynamics 180

relieving the tip stalling characteristics they do augment CL considerably even though the angle of
attack may be well below the stalling angle.

11. Automatic slats are designed so that they open fully some time before the speed reaches that
used for the approach and landing. During this period they still accomplish their purpose of making
the passage of the boundary layer easier by flattening the pressure gradient over the front of the wing.
Thus whenever the slat is open, at even moderate angles of attack, the boundary layer can penetrate
further aft along the chord thus reducing the thickening effect and delaying separation and resulting in
a stronger pressure distribution than that obtained from a wing without slats. As the angle of attack is
increased so the effect becomes more pronounced.

12. Built-in Slots. Fig 11-4 shows a variation of


the classic arrangement, in which suitably shaped slots
are built into the wing tips just behind the leading edge.
At higher angles of attack, air from below the wing is
guided through the slots and discharged over the upper
surfaces, tangential to the wing surface, thereby re-
energizing the boundary layer to the consequent benefit
of the lift coefficient.

13. Stalling with Slats. The effect of the slat at the


highest angles of attack is to boost the extent of the low-
Fig 11-4: Built-in Slot
pressure area over the wing. At angles of attack of about
25° the low-pressure envelope has been considerably enlarged and a proportionately larger amount
of lift is being developed. When the wing reaches a certain angle the slat can no longer postpone
events and the stall occurs. When the powerful low pressure envelope collapses the sudden loss of
lift may result in equally sudden changes in the attitude of the aircraft. This applies particularly if one
wing stalls before the other. In this case a strong rolling moment or wing dropping motion would be set
up.

FLAPS

14. Purpose. The operation of the flap is to vary the camber of the wing section. High lift
aerofoils possess a curved mean camber line (the line equidistant from the upper and lower surfaces),
the greater the mean camber the greater the lifting capability of the wing. High-speed aerofoils
however, may have a mean camber line, which is straight, and, if either or both the leading edge and
trailing edge can be hinged downwards, the effect will be to produce a more highly cambered wing
section, with a resultant increase in the lift coefficient.
181 Lift Augmentation

15. Action of the Flap. Increased camber can be obtained by bringing down the leading or
trailing edge of the wing (or both). The use of leading edge flaps is becoming more prevalent on large
swept-wing aircraft, and trailing edge flaps are used on practically all aircraft except for the tailless
delta. The effect of this increased camber increases the lift, but since the change in camber is abrupt
the total increase is not as much as would be obtained from a properly curved mean camber line. Fig
11-1 shows the effect of flaps on the CL max and stalling angle.

Types of Flap

16. The trailing edge flap has many variations, all of which serve to increase the CL max. Some,
however, are more efficient than others. Fig 11-5 illustrates some representative types in use. The
more efficient ones are usually more complicated mechanically.

Fig 11-5: Types of Flaps


FIS Book 1: Aerodynamics 182

17. The increase in CL max obtained by the use of flaps varies from about 50% for the plain flap, to
90% for the fowler type flap. The effectiveness of a flap may be considerably increased if the air is
constrained to follow the deflected surface and not to break away or stall. One method of achieving
this is by the use of slotted flaps, in which, when the flap is lowered, a gap is made which operates as
a slot to re-energize the air in a similar manner to leading edge slots. Some aircraft have double, or
even triple-slotted flaps which may give a CL max increase of up to 120%.

18. The angle of attack at which the maximum lift coefficient is obtained with the trailing edge flap
is slightly less than with the basic aerofoil. Thus, the flap gives an increased lift coefficient without the
exaggerated angles made necessary with slats.

19. Comparison of Different Types of Flaps. Fig 11-6 shows lift curves representing the
effect of different kinds of flaps, all at the same moderate flap angle. The curves in Fig 11-7, again
plotted for the same flap angle for each flap, compare the drag increments for given lift increments.
The Fowler flap is seen to produce the largest lift increment for the least drag increment. The curves
of Fig 11-8 show how the increment in maximum
lift coefficient varies as flap deflection is increased.
The Fowler flap is seen to give a lift increment
even at zero flap angle. This is simply because it
slides backwards, giving increase in effective area,
before being deflected downwards. Also the
Fowler flap gives the greatest lift increment at any
given flap angle. However, the optimum deflection
is least in the case of a Fowler flap, and deflection
beyond the optimum produces a more sudden
drop in effectiveness than in the case of other
Fig 11-6: Effect of Flaps on Lift Curve
types.

Fig 11-7: Effect of Flaps on Fig 11-8: Effect of Flaps on Increment


CL Vs CD Curve in CL max
183 Lift Augmentation

Effect of Flap on the Stalling Angle

20. The main effect produced by flap deflection is an increase in the effective camber of the wing.
As we have seen, this reduces the zero-lift angle of attack, without affecting the lift curve slope. Thus,
at angles of attack well bellow stall, there is a constant increment in lift coefficient. The effect may be
slightly enhanced because flap deflection additionally produces a slight increase in the effective angle
of attack. A further effect, however, is to hasten the onset of the stall, because separation tends to
occur early over the rear of the flap, where the aerofoil is, in effect, very highly cambered. The
following additional points need to be noted: -

(a) Referring to the curves of basic aerofoil and the one using plain flaps in Fig 11-6, it
can be seen that, the increment in CL max (A) is generally less than the increment of CL at
lower angle of attack because of the reduced stalling angle. It is, of course, the increment in
CL max which is most important, since it is which determines the reduction in stalling speed
with flap deflection.

(b) Further increase in flap deflection generally increases the effects already described,
although at high flap deflections the rate of increase of lift increment begins to tail off. With a
plain flap, the maximum increment is usually obtained at 40° (Fig 11-9).

(c) The reduction in stalling angle is


beneficial in that it decreases the
incidence at landing and take-off. The
aircraft operates in a less nose-up
attitude, and the pilot’s range of vision is
improved.

21. The trailing edge flap is directly


comparable to the trailing edge elevator in so far
as the effect on stalling angle is concerned. The
raised trailing edge elevator at the stall
increases the stalling angle of attack and the
aircraft attitude in level flight but the lowered
trailing edge flap reduces the stalling angle and
the aircraft attitude at the level flight stall. Fig
11-9 illustrates how the lowered flap affects the
angle of attack and the aircraft attitude. Pilots
should take care not to confuse attitude with
angle of attack, as the attitude of the aircraft has Fig 11-9: Effect of Flaps on Stalling Angle
and Level Flight Stalling Attitude
no fixed relationship with the angle of attack
while manoeuvring.
FIS Book 1: Aerodynamics 184

22. Changes in Pressure Distribution


with Flap. All types of flaps when deployed,
change the pressure distribution across the
wing. Typical pressure distributions with flap
deployed are shown in Fig 11-10 for plain and
slotted flaps. The slotted flap has a much
greater intensity of loading on the flap itself.

23. Change in Pitching Moment with


Flap. All trailing edge flaps produce an
increased nose-down pitching moment due to
the change in pressure distribution around the
wing flap. Flaps, when lowered, may have the
added effect of increasing the downwash at the
tail. The amount by which the pitching moment
is changed resulting from this increase in Fig 11-10: Change in Pressure Distribution
downwash depends on the size and position of with Flaps

the tail. These two aspects of change in pitching moment generally oppose each other and on
whichever aspect is dominant will depend whether the trim change on lowering flap is nose-up or
nose-down. Leading edge flaps tend to reduce the nose-down pitching moment and reduce the wing
stability near the stall.

Lift / Drag Ratio

24. Although lift is increased by lowering flaps there is also an increase in drag and
proportionately the drag increase is much greater when considered at angles of attack about those
giving the best lift / drag ratio. This means that lowering flap almost invariably worsens the best lift /
drag ratio.

25. For a typical split or trailing edge flap, as soon as the flap starts to lower, the lift and drag start
increasing. Assuming that the flap has an angular movement of 90°, for about the first 30° there is a
steady rise in the CL. During the next 30° the CL continues to increase at a reduced rate and during
the final part of the movement a further very small increase occurs.

26. In conjunction with the lift the drag also increases, but the rate of increase during the first 30°
is small compared with that which takes place during the remainder of the movement, the final 30°
producing a very rapid increase in the rate at which the drag has been rising.

27. When flap is used for take-off or manoeuvring it should be set to the position recommended in
the Aircrew Manual. At this setting the lift / drag ratio is such that the maximum advantage is obtained
185 Lift Augmentation

for the minimum drag penalty. For landings, however, the high drag of the fully lowered flap is useful
since it permits a steeper approach without the speed becoming excessive (i.e. it has the effect of an
airbrake). The increased lift enables a lower approach speed to be used and the decreased stalling
speed means that the touchdown is made at a lower speed. The high drag has another advantage in
that it causes a rapid deceleration during the period of float after rounding out and before touching
down.

Use of Flap for Take-Off

28. The increased lift coefficient when the flaps are lowered shortens the take-off run provided
that the recommended amount of flap is used. The flap angle for take-off is that for the best lift/drag
ratio that can be obtained with the flaps in any position other than fully up. If larger amounts of flap
are used, although the lift is increased, the higher drag slows the rate of acceleration so that the take-
off run, although perhaps shorter than with no flap, is not the shortest possible.

29. When the take-off is made at or near the maximum permissible weight, the flaps should
invariably be set to the recommended take-off angle so that the maximum lifting effort can be obtained
from the wing.

30. Raising of Flaps in Flight. Shortly after take-off, while the aircraft is accelerating and
climbing slightly, the action of raising the flaps causes an immediate reduction in the lift coefficient and
the aircraft loses height or sinks unless this is countered by an increase in the angle of attack. If the
angle of attack is not increased, i.e. if the pilot makes no correcting movement with the control
column, the reduced lift coefficient results in a loss of lift which causes the aircraft to lose height until it
has accelerated to a higher air speed that counterbalances the effect of the reduced CL. When the
flaps are raised and the sinking effect is countered by an increased angle of attack, the attitude of the
aircraft becomes noticeably more nose-up as the angle of attack is increased. The more efficient the
flaps the greater is the associated drop in lift coefficient and the larger the subsequent corrections that
are needed to prevent loss of height. On some aircraft it is recommended that the flaps should be
raised in stages so as to reduce CL gradually and so avoid any marked and possibly exaggerated
corrections. This applies sometimes when aircraft are heavily loaded, particularly in the larger types
of aircraft.

Leading Edge Flap

31. The effect of leading edge flaps is, as with other flaps, to increase the CL and lower the
stalling speed. However, the fact that it is the leading edge and not the trailing edge that is drooped
results in an increase in the stalling angle, and the level flight stalling attitude. The difference is
explained as before, by the fact that although the stalling angle measured with respect to the chord
line joining the leading and trailing edges of the changed section may not be affected, the stalling
FIS Book 1: Aerodynamics 186

angle is increased
when measured, as is
conventional, with
respect to the chord
line of the wing with
the flap fully raised.
Fig 11-11 illustrates
this point. Leading
edge flaps are Fig 11-11: Effect of Leading Edge Flap on Stalling Angle

invariably used in conjunction with trailing edge flaps. The operation of the leading edge flap can be
controlled directly from the cockpit or it can be linked, for example, with the air speed measuring
system so that the flaps droop when the speed falls below a certain minimum and vice versa.

32. The effect of leading edge flaps is


similar to that of slats except that the stalling
angle is not increased as much. The amount
of increase in the CL max is about the same in
both cases. The leading edge flap is
sometimes referred to as a nose flap. The
effect of leading edge flap on the lift curve is
shown in fig 11-12.

Jet Flap

Fig 11-12: Effect of Leading Edge Flap


33. Jet flap is a natural extension of a on Lift Curve
slot blowing over trailing edge flaps for
boundary layer control, using much higher
quantities of air with a view to increasing the
effective chord of the flap to produce so-
called "super circulations" about the wing.

34. The term jet flap implies that the gas Fig 11-13: Jet Flap
efflux is directed to leave the wing trailing edge as a plane jet at an angle to the main stream so that
an asymmetric flow pattern and circulation is generated about the aerofoil in a manner somewhat
analogous to a large trailing edge flap. By this means, the lift from the vertical component of the jet
momentum is magnified several times by "pressure lift" generated on the wing surface, while the
sectional thrust lies between the corresponding horizontal component and the full jet momentum. To
facilitate variation of jet angle to the main stream direction, the air is usually ejected from a slot
forward of the trailing edge, over a small flap whose angle can be simply varied (Fig 11-13) Such
basic jet flap schemes essentially require the gas to be ducted through the wing and so are often
187 Lift Augmentation

referred to as "internal flow" systems. Fig 11-14


illustrates these, and the basics of their ‘external
flow’ configurations.

35. Effect of Sweepback on Flap. The


use of sweepback will reduce the effectiveness
of trailing edge control surfaces and high lift
devices. A typical example of this effect is the
application of a single slotted flap over the
inboard 60% span to both a straight wing and a
wing with 35° sweepback. The flap applied to
the straight wing produces an increase in the CL
max of approximately 50%. The same type flap Fig 11-14: Some Jet Flap Systems
applied to the swept wing produces an increase
in CL max of approximately 20%. The reason for this is the decrease in frontal area of flap with
sweepback.

36. Effect of Flap on Wing Tip Stalling. Lowering flaps may either increase or alleviate any
tendency towards the tip stalling of swept wings. When flaps are lowered the increased downwash
over the flaps and behind them induces a balancing upwash over the outer portions of the wing at
high angles of attack and this upwash may be sufficient to increase the angle of attack at the tip to the
stalling angle. On the other hand, because of the lowered flaps, the higher suctions obtained over the
inboard sections of the wing have the effect of restricting the outward flow of the boundary layer and
thus a beneficial effect is obtained on wing tip stalling tendencies. The practical outcome of these
opposing tendencies is dependent on the pressure configuration.

Additional Effects of Flap

37. Although flap deflection produces a nose down change in pitching moment, this is associated
simply with a negative change in the zero-lift pitching moment coefficient, Cmo. There is generally no
significant effect on the derivative d Cm / d Cmo, and therefore no change in the position of the
aerodynamic center.

38. Landing and take off flaps are usually part-span flaps only, situated in the inboard section of
the wings, because the corresponding outboard regions are occupied by the ailerons. This reduces
the flap effectiveness. It also changes the span wise load distribution, and may thus reduce the
induced drag. It has the further effect of moving the center of lift inwards on each wing, and this may
affect the lateral stability of the aircraft.
FIS Book 1: Aerodynamics 188

BOUNDARY LAYER CONTROL

General

39. Any attempt to increase the lifting effectiveness of a given wing directly and fundamentally
concerns the boundary layer. If the boundary layer can be made to remain laminar and unseparated
as it moves over the wing, then not only is the lift coefficient increased but also both surface friction
drag and form drag is reduced.

40. There are various methods of controlling the boundary layer so that it remains attached to the
aerofoil surface as far as possible. They all depend on the principle of adding kinetic energy to the
lower layers of the boundary layer.

Boundary Layer Control by Suction

41. If enough suction could be applied through a series of slots or a porous area on the upper
surface of the wing, separation of the boundary layer at almost all angles of attack could be
prevented. However, it has been found that the power required to draw off the entire boundary layer
so that it is replaced by completely undisturbed air, is so large that the entire output of a powerful
engine would be required to accomplish this.

42. However, even moderate amounts of suction have a beneficial effect in that the tendency to
separate at high angles of attack can be reduced. The effect of moderate suction is to increase the
strength and stability of the boundary layer.

43. The effect of the suction is to draw off the lower layer (the sub-layer) of the boundary layer, so
that the upper part of the layer moves on to the surface of the wing. The thickness of the boundary
layer is thereby reduced and also its speed is increased, since the heavily retarded sub-layer has
been replaced by faster moving air.

44. The suction is effected either through a slot or series of slots in the wing surface or by having
a porous surface over the area in which suction is required. These devices are positioned at a point
where the thickening effect of the adverse pressure gradient is becoming marked and not at the
beginning of the adverse pressure gradient. Generally, suction distributed over a porous area has a
better effect than the concentrated effect through a slot.

Boundary Layer Control by Blowing

45. When air is ejected at high speed in the same direction as the boundary layer at a suitable
point close to the wing surface, the result is to speed up the retarded sub-layer and re-energize the
189 Lift Augmentation

complete boundary layer, this again enables it to penetrate further into the adverse pressure gradient
before separating.

46. Very high maximum lift coefficients can be obtained by combining boundary layer control with
the use of flaps like in some Mig 21 variants. In this case the suction, or blowing, of air takes place
near the hinge line of the flap. An average CL max, for a plain aerofoil is about 1.5, for the same
aerofoil with a flap it may be increased to about 2.5. When boundary layer control is applied in the
form of blowing or suction over the flap
(Fig 11-15), the CL max may rise to as high
as 5. When this figure is put into the lift
formula under a given set of conditions it
can be seen that the amount of lift
obtained is greatly increased when
compared with that from the plain aerofoil Fig 11-15: Blown Flap
under the same conditions.

47. In addition to the use of boundary layer control over the flaps themselves, it can be used
simultaneously at the leading edge. In this way even higher lift coefficients can be obtained. The
practical limit, in the conventional fixed or rotating wing aircraft, is set by the large amount of power
needed to obtain the suction or blowing which is necessary to achieve these high figures. By the use
of the maximum amount of boundary layer control, wind tunnel experiments on a full size swept-wing
fighter have realized an increase in CL max such that the normal 200 kmph landing speed was reduced
to 120 kmph. This is evidence of the importance of the part that the boundary layer plays in
aerodynamics.

48. Vortex Generators. Vortex generators can either take the form of metal projections from
the wing surface or of small jets of air issuing normal to the surface. Both types work on the same
principle of creating vortices which entrain the faster moving air near the top of the boundary layer
down into the more stagnant layer near the surface thus transferring momentum which keeps the
boundary layer attached further back on the wing. There are many shapes for the metal projections,
such as plain rectangular plates or aerofoil sections, the exact selection and positioning of which
depends on the detailed particular requirement of the designer. In general, careful design selection
can ensure that the increase in lift due to the generators can more than offset the extra drag they
cause. An advantage of the air jet type is that they can be switched off for those stages of flight when
they are not required and thus avoid the drag penalty.

49. Deflected Slip-Stream. A simple and practical method of using engines to assist the wing
in the creation of lift is to arrange the engine / wing layout so that the wing is in the fan or propeller
slipstream. This, combined with the use of leading edge slats and slotted extending rear flaps,
provides a reliable high lift coefficient solution for STOL aircraft. A more extreme example involves
FIS Book 1: Aerodynamics 190

the linking together of engines and synchronization of propellers. The wings and flaps are in the
slipstream of the engines and gain effectiveness by deflecting the slipstream downwards.

SUMMARY

Devices to Augment Lift

50. The chief devices used to augment CL max are:

(a) Slats / Slots.


(b) Flaps.
(c) Boundary layer control.

51. The effect of the slat is to prolong the lift curve by delaying the stall until a higher angle of
attack. The effect of flap is to increase the lift by increasing the camber and the principle of boundary
layer control is to add kinetic energy to the sub-layers of the boundary layer, so as to delay boundary
layer separation.

52. Since the slat is of use only at high angles of attack, at the normal angles its presence serves
only to increase drag. This disadvantage is overcome by making the slat movable so that when it is
not in use it lies flush against the leading edge of the wing.

53. Flaps may be on the leading or trailing edge and many aircraft have them on both. Besides
increasing the camber of the wing, many types of flap also increase the wing area. Trailing edge flaps
reduce the stalling angle while the leading edge flaps increase the stalling angle.

54. The main methods of boundary layer control are:

(a) Blowing.
(b) Suction.
(c) Vortex generators.
191

CHAPTER 12

FLIGHT CONTROLS

General Considerations

1. All aircraft have to be fitted with a control system that will enable the pilot to manoeuvre and
trim the aircraft in flight about each of its three axes.

2. The aerodynamic moments required to rotate the aircraft about each of these axes are
usually produced by means of flap-type control surfaces positioned at the extremities of the aircraft so
that they have the longest possible moment arm about the CG.

3. There are usually three separate control systems and three sets of control surfaces, namely:

(a) Rudder for control in yaw.


(b) Elevator for control in pitch.
(c) Ailerons for control in roll (the use of spoilers for control in roll is also discussed in this
chapter).

4. On some aircraft the effect of two of these controls is achieved by a single set of control
surfaces, for example:

(a) Elevons, which combine the effects of ailerons and elevators, e.g. Mirage-2000.
(b) Ruddervator, also called Vee or butterfly tail, combining the effects of rudder and
elevators, e.g. F-117.
(c) Tailerons are slab tail surfaces that move either together, as pitch control, or
independently for control in roll, e.g. Jaguar (differential tail plane).
(d) Flaperons are flaps acting as ailerons when stick is moved laterally, e.g. SU-30.

5. It is desirable that each set of control surface should produce a moment only about its
corresponding axis. In practice, however, moments are often produced about the other axes as well,
e.g. adverse yaw due to aileron deflection. Some of the design methods used to compensate for
these cross effects are discussed in later paragraphs.

Control Characteristics

6. Control Power and Effectiveness. The main function of a control is to allow the aircraft to
fulfill its particular role. This aspect is decided mainly by:

(a) Size and shape of the control.


FIS Book 1: Aerodynamics 192

(b) Deflection angle.


(c) EAS2.
(d) Moment arm (distance from CG).

In practice the size and shape of the control surface are fixed and, since the CG movement is small,
the moment arm is virtually constant. The only variables in control effectiveness are speed and
effective deflection angle.

7. Control Moment. Consider the effect of


deflecting an elevator downwards. The angle of
attack and the camber of the tailplane are both
increased, thereby increasing the CL value (Fig
12-1). The change in moment produced by the tail
is the product of the change in lift produced by the
tail X the moment arm to the CG.

8. Effect of Speed. The aerodynamic


forces produced on an aerofoil vary as the square
of the speed. It follows then that for any given
control deflection, the lift increment, and so the
moment, will vary as the square of the speed i.e.
EAS2. The deflection angle required to give an Fig 12-1: Effect of Elevator Deflection on
attitude change, or response, is inversely the Lift Coefficient Curve
2
proportional to the EAS . To the pilot this means that when the speed is reduced by half, the control
deflection is increased fourfold to achieve the same result. This is one of the symptoms of level flight
stall and is erroneously referred to as ‘sloppy controls’.

9. Control Forces. When a control surface is deflected, e.g. down elevator, the aerodynamic
force produced by the control itself opposes the downward motion. A moment is thus produced about
the control hinge line, and this must be overcome in order to maintain the position of the control. The
stick force experienced by the pilot depends upon the hinge moment and the mechanical linkage
between the stick and the control surface. The ratio of stick movement to control deflection is known
as the stick-gearing, and is usually arranged so as to reduce the hinge moment i.e. stick-gearing
should be such that for a lesser stick movement a larger control deflection is achieved.

Aerodynamic Balance

10. If control surfaces are hinged at their leading edge and allowed to trail from this position in
flight, the forces required to change the deflection angle on all except light and slow aircraft would be
prohibitive. To assist the pilot to move the controls in the absence of powered or power-assisted
controls, some degree of aerodynamic balance is required.
193 Flight Controls

11. In all its forms, aerodynamic balance is a means of reducing the hinge moment and thereby
reducing the physical effort experienced in controlling an aircraft. The most common forms of
aerodynamic balance are:

(a) Horn balance.


(b) Inset hinge / Set back hinge.
(c) Internal balance / Sealed nose balance.
(d) Various types of tab balance.

12. Horn Balance. On most control surfaces,


especially rudders and elevators, the area ahead of the
hinge is concentrated on one part of the surface in the
form of a horn (Fig 12-2). The horn thus produces a
balancing moment ahead of the hinge-line. In its effect
the horn balance is similar to the inset hinge. It also
helps in mass balancing (as CG is moving closer to
hinge line) e.g. shielded horn in HPT - 32 and Avro.
Fig 12-2: Horn Balance

13. Inset Hinge. The most obvious way to


reduce the hinge moment is to set the hinge-line inside
the control surface thus reducing the moment arm.
The amount of inset is usually limited to 20-25% of the
chord length. This ensures that the CP of the control
will not move in front of the hinge at large angles of
control deflection (Fig 12-3) e.g. HPT – 32, Kiran etc.

14. Over Balance. Should the control CP move


ahead of the hinge line, the hinge moment would
assist the movement of the control and the control
would then be over-balanced.

15. Internal Balance / Sealed Nose Balance. Fig 12-3: Hinge Moment and
Although fairly common in use, this form of Inset Hinge

aerodynamic balance is not very obvious because it is


contained within the contour of the control. With this
type of balance (Fig 12-4), a plate projects forward
from the nose of the control surface. This pate, or
tongue, is joined to the main part of the wing, tailplane
or fin by a loose fold of impermeable fabric, which Fig 12-4: Internal Balance
constitutes a seal between the regions above and below the control, or on the two sides of the control
surface, in the case of rudder. The space above the tongue, in the case of an elevator, say, is open
FIS Book 1: Aerodynamics 194

to the air above the tailplane, and similarly the space below the tongue is open to the air below the
tail. When the elevator is deflected downwards, lift is created on the elevator. This is a result of a
generally reduced pressure on the upper surface and an increased pressure on the lower surface. In
particular, the pressure just above the tongue is low, and just below is high, so that there is an upward
force on the tongue which provides a nose-up hinge moment, which is the required balancing
moment.

16. Tab Balance. This subject is fully covered in the chapter on Trimming and Balance Tabs.

MASS BALANCE & FLUTTER

17. Torsional Aileron Flutter. This is caused by


the wing twisting under loads imposed on it by the
movement of the aileron. Fig 12-5 shows the sequence
for a half cycle, which is described as follows:

(a) The aileron is displaced slightly


downwards, exerting an increased lifting force on
the aileron hinge.

(b) The wing twists about the torsional axis.


The trailing edge rising, taking the aileron up with
it. The CG of the aileron is behind the hinge line,
its inertia tends to make it lag behind, increasing
aileron lift, and so increasing the twisting
moment.
Fig 12-5: Torsional Aileron Flutter

(c) The torsional reaction of the wing has arrested the twisting motion but the air loads on
the aileron, the stretch of its control circuit, and its upward momentum, cause it to overshoot
the neutral position, placing a down load on the trailing edge of the wing.

(d) The energy stored in the twisted wing and the reversed aerodynamic load of the
aileron cause the wing to twist in the opposite direction. The cycle is then repeated.

18. Torsional aileron flutter can be prevented either by mass-balancing the ailerons so that their
CG is on, or slightly ahead of, the hinge line, or by making the controls irreversible. Both methods are
employed in modern aircraft. Those aircraft with fully powered controls and no manual reversion do
not require mass-balancing. All other aircraft have their control surfaces mass-balanced.

19. Flexural Aileron Flutter. Flexural aileron flutter is generally similar to torsional aileron
flutter but is caused by the movement of the aileron lagging behind the rise and fall of the outer
195 Flight Controls

portion of the wings as it flexes, thus tending to increase the oscillation. This type of flutter is
prevented by mass-balancing the aileron. The positioning of the mass-balance weight is important.
The nearer the wing tip the smaller the weight required. On many aircraft the weight is distributed
along the whole length of the aileron in the form of a leading edge spar, thus increasing the stiffness
of the aileron and preventing a concentrated weight starting torsional vibrations in the aileron itself.

20. Torsional Flexural Flutter. This type of flutter takes place on wings, which have poor
(flexural) stiffness and their tortional axis lying ahead of its CG axis, “the control surface not playing
any significant role in this behaviour”. If such a wing, flying at very high speed, encounters a sudden
increases in α due to gust any other reason, the wing tips tend to flex upwards. The CG axis of the
wing, lying behind the torsional axis, causes the wing to twist, thereby causing a further increase in α
and the flutter similar to the flexural aileron flutter is set up. This type of flutter is prevented by mass
balancing the entire wing so as to get the wing CG axis ahead of the torsional axis.

21. Other Control Surfaces. So far only wing flutter has been discussed, but a few moments
consideration will show that mass balancing must be applied to elevators and rudders to prevent their
inertia and the springiness of the fuselage starting similar troubles. Mass balancing is extremely
critical, hence to avoid upsetting it, the painting of aircraft markings etc. should not be done on any
control surface. The danger of all forms of flutter is that the extent of each successive vibration is
greater than its predecessor, so that in a second or two the structure may be bent beyond its elastic
limit and fail.

Control Requirements

22. The main considerations are outlined briefly below:

(a) Control Forces. If the stick forces are too light the pilot may overstress the aircraft,
whereas if they are too heavy he will be unable to manoeuvre it. The effort required must be
related to the role of the aircraft and the flight envelope.

(b) Control Movements. If the control movements are too small the controls will be too
sensitive, whereas if they are too large the designer will have difficulty in fitting them into the
restricted space of the cockpit.

(c) Control Harmony. An important factor in the pilot’s assessment of the overall
handling characteristics of an aircraft is the “harmony” of the controls with respect to each
other. Since this factor is very subjective it is not possible to lay down precise quantitative
requirements. One method often used is to arrange for the aileron, elevator and rudder forces
in the ratio 1:2:4.
FIS Book 1: Aerodynamics 196

Control Response

23. Ailerons. Consider the effect of applying aileron. Deflection of the controls produces a
rolling moment about the longitudinal axis and this moment is opposed by the aerodynamic damping
in roll (the angle of attack of the down-going wing is increased while that of the up-going wing is
decreased, explained in the chapter on Stability). The greater the rate of roll, the greater the
damping. Eventually the rolling moment produced by the ailerons will be exactly balanced by the
damping moment and the aircraft will attain a steady rate of roll. Usually the time taken to achieve the
‘steady state’ is very short, probably less than one second. Thus, over most of the time that the
ailerons are being used, they are giving a steady rate of roll response and this is known as steady
state response. The transient response is that which is experienced during the initiation period
leading to a steady manoeuvre.

24. Rate versus Acceleration Control (Aileron). A conventionally operated aileron is


therefore described as a rate control that is the aircraft responds at a steady rate of movement for
almost all of the time. The stick force required to initiate a manoeuvre may be less than, or greater
than, the stick force required to sustain the manoeuvre, e.g. rolling. A favourable response is usually
assumed to be when the initiation force is slightly greater than the steady force required, the
difference being up to 10%. On some modern aircraft the transient response time to aileron deflection
is increased. By the time the steady rate of roll is obtained the aileron is being taken off again. In
these circumstances the aileron control is producing a rate of roll response, which is always
increasing, and it is then described as an acceleration control.

25. Elevators. When the elevators are used they produce a pitching moment about the lateral
axis. The resulting pitching movement is opposed by the aerodynamic damping in pitch and by the
longitudinal stability of the aircraft (see the chapter on Stability). The response to the elevator is a
steady state change in angle of attack with no transient time, that is a steady change of attitude.
Elevators are therefore described as a displacement control.

26. Rudders. The yawing moment produced by rudder deflection is opposed by aerodynamic
damping in yaw and by the directional stability of the aircraft. The response to the rudder is a steady
state of change of angle of attack on the keel surfaces of the aircraft, with no transient time. The
rudder control has a similar response to the elevator and is therefore also described as a
displacement control.

27. Steady State Response. In the chapter on Stability, it is shown that stability opposes
manoeuvre, or more precisely, the steady state response to control deflection is greatly affected by
the static stability. This is easily seen when considering the response to rudder. The heavy rudder
force required to sustain a steady yawing movement is a result of the strong directional static stability
of most aircraft.
197 Flight Controls

28. The steady state response to controls is of greater importance to the pilot and this subject is
discussed in the following paragraphs.

PRIMARY CONTROL SURFACES

Elevators

29. The elevators are hinged to the rear spar of the tailplane and are connected to the control
column so that forward movement of the control column moves the elevator (trailing edge) downwards
and backward movement moves the elevator upwards. When the control column is moved back and
the elevator rises, the effect is to change the overall tailplane/elevator section to an inverted aerofoil
which supplies a downward force on the tail of the aircraft and, as seen by the pilot, raises the nose.
The opposite occurs when a forward movement is made.

30. Elevators are normally free from undesirable characteristics, but large stick forces may be
experienced on some aircraft if the aerodynamic balance of the elevators or the stability
characteristics of the aircraft are at fault. The tail moment arm is determined by the position of CG,
and to retain satisfactory handling characteristics throughout the speed range, the CG position must
be kept within a certain limited range. If the CG moves too far forward, the aircraft becomes
excessively stable and the pilot will run out of up-elevator before reaching the lowest speeds required.

31. Consider an aircraft in the flare-out and landing where longitudinal static stability opposes the
nose-up pitch. The downwash angle at the tail is much reduced by the ground effect thereby
increasing the effective angle of attack at the tail. This results in a reduced tail-down moment in pitch,
therefore a greater elevator deflection angle is required to achieve the landing attitude than would be
required to achieve the same attitude at height. A forward movement of the CG would add to this
problem by increasing the longitudinal stability.

32. Some tailless delta aircraft require an aft movement of the CG (fuel transfer) to ensure that
the elevator movement is sufficient to achieve the landing attitude, e.g. Concorde.

33. Some light aircraft, notably civilian flying club types with a tricycle undercarriage have
undersized elevators in order to make the aircraft ‘unstallable’.

Ailerons

34. It has been seen that the aileron is a rate control and the rate of roll builds up rapidly to a
steady value dictated by the damping in roll effect. The value of the steady rate of roll produced by a
given aileron deflection will depend upon the speed and altitude at which the aircraft is flying.
FIS Book 1: Aerodynamics 198

Additional effects due to aero-elasticity and compressibility may be present and will modify the roll
response.

35. Effect of Altitude. In a steady


state of roll the rolling force and the
damping in roll force are balanced. The
rolling force is caused by the change in lift
due to aileron deflection and is
proportional to the amount of aileron
deflection and to EAS. The damping force Fig 12-6: Damping in Roll Effect
On Down going Wing
is due to the change in lift caused by the
increase in angle of attack of the downgoing wing and the decrease in angle of attack of the upgoing
wing. The value of the damping angle of attack can be found by the vector addition of the TAS and
the rolling velocity as illustrated in Fig 12-6. It can be seen that for a constant damping angle of
attack, the rolling force, and therefore the rate of roll, will increase in direct proportion to TAS. When
an aircraft is climbed at a constant EAS the rate of roll for a given aileron deflection therefore
increases because TAS increases.

36. Effect of Forward Speed. It has been shown that the rolling force is proportional to the
amount of aileron deflection and to EAS. It follows that rate of roll increases in proportion to EAS.

Aero-Elastic Distortion

37. Aileron Reversal. Ailerons are located towards the wing tips by the necessity in most
aircraft to utilize trailing edge flaps for landing and take-off. This may cause the wing to twist when
the ailerons are deflected and lead to an effect known as “aileron reversal”. It must be recognized
that aero-elastic distortion of the airframe may affect stability and control in pitch and yaw as well as in
roll. As the wings are usually the least rigid part of the airframe, aileron reversal assumes importance
as it reduces the ultimate rate of roll available at high forward speeds.

38. In Fig 12-7 it can be seen that downward


deflection of the aileron produces a twisting moment
about the torsional axis of the wing. The torsional
rigidity of the wing depends on the wing structure but
will normally be strong enough to prevent any
distortion at low speeds. Aileron power, however,
increases as the square of the forward speed,
whereas the torsional stiffness in the wing structure
is constant with speed. At high speeds therefore,
Fig 12-7: Aileron Reversal
the twisting moment due to aileron deflection
199 Flight Controls

overcomes the torsional rigidity of the wing and produces a change in incidence which reduces the
rate of roll. On the rising wing illustrated in Fig 12-7 the incidence is reduced whereas on the
downgoing wing the effect is to increase the incidence. A flight speed may be reached where the
increment of CL produced by deflecting aileron is completely nullified by the wing twisting in the
opposite sense. At this speed, called reversal speed, the lift from each wing is the same in spite of
aileron deflection, and the rate of roll will be zero. At still higher speeds the direction of roll will be
opposite to that applied by the pilot. Reversal speed is normally outside the flight envelope of the
aircraft but the effects of aero-elastic distortion may be apparent as a reduction in roll rate at the
higher forward speeds.

39. Wing Divergence. This type of aero


elastic distortion takes place at high dynamic
pressure (high speeds) and like aileron reversal,
occurs due to the interaction of aerodynamic
forces and elastic deflection of the wing
structure. Unlike the aileron reversal, this has
no contribution from the control surfaces. This
differs from aileron reversal in that it is a violent
instability which produces immediate failure. Fig
12-8 illustrates the process of instability. If the
wing surface is above divergence speed, any
disturbance in angle of attack can precipitate
this sequence. Any increase in α causes lift to
increase as well as the CP to shift forward. If
the tortional axis of the wing lies behind the
aerodynamic axis, an increase in lift force and
the increase in the moment arm between CP
and the tortional axis causes the wing to twist
(Wing leading edge up). This causes a further
increase in the α, greater lift, further forward
location of CP, greater twisting of the wing, more
lift and so on until the wing fails. At low speeds,
where dynamic pressure is low, the wing
tortional stiffness is able to overcome this Fig 12-8: Wing Divergence

twisting. However, the change in lift per degree α is proportional to V2 but structural stiffness of the
wing remains constant. This relationship implies that at some high speed the aerodynamic force build
up will overcome the resisting tortional stiffness and divergence will occur. This can be avoided by
restricting the maximum speed of the aircraft, designing the wing such that its tortional axis lies ahead
of its aerodynamic axis or by increasing the wing tortional stiffness e.g. by having multi spar
construction, sweep back, short span and high taper ratio and / or aero elastic tailoring.
FIS Book 1: Aerodynamics 200

40. Aileron Response at Low Speeds. Deflecting an aileron down produces an effective
increase in camber and a small reduction in the stalling angle of attack of that part of the wing to
which the aileron is attached (usually the wing tip). It is therefore important that the wings should stall
progressively from root to tip in order to retain aileron effectiveness at the stall. Rectangular straight
wings are not much of a problem because washout is usually incorporated to reduce vortex drag.
Swept wings, however, are particularly prone to tip stall and design features may have to be
incorporated to retain lateral control at low speeds. Some of these aspects were discussed in the
chapter on Wing Planforms. Loss of effectiveness of the down-going aileron due to tip stall will not
necessarily result in a reversal in the
direction of roll (i.e. wing-drop). The up-
going aileron will usually retain its
effectiveness and produce a lesser but
conventional response. The factor which
has the most significant effect on aircraft
response at high α is the damping in roll
effect. It has been seen that when the
aircraft is rolling, the angle of attack of the
down-going wing is increased while that of
the rising wing is decreased. If the aircraft is
close to stall the damping effect is reversed
and the change in rolling moments assists
the rotation (Fig 12-9). This leads to the
Fig 12-9: Damping in Roll
phenomenon known as autorotation.

41. Adverse Aileron Yaw. Unless they are carefully designed aerodynamically, the ailerons will
materially alter the drag force on the wing in addition to the desired change in lift force. When an
aileron is deflected downwards both the vortex drag and the boundary layer drag are increased. On
the aileron-up wing, however, the vortex drag is decreased, though the boundary layer drag may be
still increased. The changes in the drag forces are such as to produce a yawing moment causing the
aircraft to yaw in the opposite sense to the applied roll. Adverse yaw is produced whenever the
ailerons are deflected but the effect is usually reduced by incorporating one or more of the following
design features:

(a) Differential Ailerons. For a given stick deflection the up-going aileron is deflected
through a larger angle than the down-going aileron thus reducing the difference in drag and
the adverse yaw. e.g. Jaguar, Hawk, MiG 27 etc.

(b) Frise-Type Ailerons. The nose of the upgoing aileron protrudes into the airstream
below the wing to increase the drag on the down-going wing. This arrangement has the
additional advantage that it assists the aerodynamic balancing of the ailerons e.g. Kiran.
201 Flight Controls

(c) Coupling of Controls. A method used in some modern aircraft to overcome the
adverse yaw is to gear the rudder to the ailerons so that when the ailerons are deflected the
rudder moves to produce an appropriate yawing moment.

(d) Spoilers. On some aircraft, spoilers in the form of flat plates at right angles to the
airflow are incorporated. These extend into the airflow causing a breakdown of the flow and
the consequent loss of lift on the required wing, thus a roll response is produced. The
extended spoiler also increases the drag of the down-going wing thus inducing an appropriate
yaw. Spoilers are sometimes the only form of lateral control used at high speeds. The
ailerons are used at low speeds, where spoilers are least effective, but spoilers alone may be
used at high speeds e.g. Jaguar aircraft. Other uses of spoilers are as airbrakes, when the
spoilers of both wings operate together, or as lift dumpers, when the aircraft has landed.
They are usually hydraulically operated.

42. Cross-Coupling Response.


When an aircraft is rolling the increased
angle of attack of the down-going wing
increases both the lift and the drag,
whereas on the up-going wing the lift
and drag are reduced. The lift and drag
forces are by definition, however,
perpendicular and parallel respectively
to the local relative airflow on each wing.
The projections of the lift forces onto the
yawing plane produce a yawing moment
towards the rising wing (i.e. an adverse
yaw) as shown in Fig 12-10. This is
partially offset by the projections of the
drag forces onto the yawing plane which
usually produce a yaw in the same
direction as roll Fig 12-10. For a given
rate of roll the change in angle of attack
will be greatest at low forward speeds.
The adverse yaw due to roll is therefore
greatest at low speed and may Fig 12-10: Cross-Coupling Response

eventually become favorable at some high forward speed.

43. Response to Sideslip. The lateral response of the aircraft to sideslip is usually called the
‘dihedral effect’ and produces a rolling moment opposite to the direction of sideslip. In most
conventional aircraft this contribution is dominated by the yaw due to sideslip.
FIS Book 1: Aerodynamics 202

44. Overbalance. This may occur at any airspeed on some aircraft not fitted with power-
operated controls, but usually only at larger control angles. It is shown by a progressive decrease,
instead of an increase, of the aileron stick force as the control column is moved, i.e. a tendency for the
ailerons to move to their full travel on their own accord. In some cases this may happen fairly
suddenly.

45. Snatch. On manually operated controls, snatching may usually occur at or near the stall, or
at high Mach numbers. It is caused by a continuous and rapid shifting of the centre of pressure of the
aileron due to the disruption of the airflow over the surface, resulting in a snatching or jerking of the
control, which may be violent.

Rudder

46. The rudder, which is hinged to the rear of the fin, is connected to the rudder bar. Pushing the
right pedal will cause the rudder (its trailing edge) to move to the right, and in doing so, alter the
aerofoil section of the fin/rudder combination. This provides an aerodynamic force on the rear of the
aircraft which will move it to the left or, as the pilot sees it, the nose will yaw to right. Rudder
effectiveness increases with speed. Whereas a large deflection may be required at low speed to yaw
a given amount, a much smaller deflection is needed at high speeds. The steady response of an
aircraft to rudder deflection is complicated by the fact that yaw results in roll, and rudder deflection
may also cause the aircraft to roll.

47. Damping in Yaw. Just as


there is damping in pitch and roll, so
also is the aircraft damped in yaw.
The effect is similar in principle to
damping in roll, in that the yawing
velocity produces a change in the
angle of attack of the keel surfaces.
Keel surface moments fore and aft
Fig 12-11: Damping in Yaw
of the CG oppose the yawing
movement, and these moments exist only while the aircraft is yawing. For simplicity, Fig 12-11 shows
the damping effect on the fin and rudder while the aircraft is yawing to the right. The wings also
produce a small damping in yaw effect because the outer wing moves faster than the inner wing and
therefore produces more drag. The damping in yaw effect decreases with altitude for the reasons
stated in para 34.

48. Roll Due to Rudder Deflection. The rudder control inevitably produces a rolling moment
because the resultant control force acts above the longitudinal axis of the aircraft. Usually this effect
203 Flight Controls

is small but on aircraft with a tall fin and rudder it can be important and is sometimes eliminated by
linking up the rudder and aileron circuits.

49. Cross-Coupling Response. The response of the aircraft in roll to rudder deflection arises
from the resulting yawing motion and/or sideslip. The roll due to sideslip, i.e. the dihedral effect, has a
powerful cross-coupling effect, particularly on swept-wing aircraft. Rudder control is often used to pick
up a dropped wing at low speeds in preference to using large aileron deflection. If the starboard wing
drops and port rudder is applied, the aircraft sideslips to starboard and the dihedral effect produces a
rolling moment tending to raise the lower wing. The roll due to yaw (i.e. rate of yaw) arises because
the outer wing moves faster than the inner wing and therefore produces more lift. The roll with yaw is
greatest at high angles of attack/low speeds.

50. Rudder Overbalance. It is one of the problems that can be encountered in flight. This is
indicated by a progressive lessening of the foot loads with increasing rudder displacement. If, owing to
a weakness in design, the aerodynamic balance is too great, it will become increasingly effective as
the rudder is moved and may eventually cause it to lock hard over when the centre of pressure moves
in front of the hinge-line. At large angles of yaw (sideslip) the fin may stall causing a sudden
deterioration in rudder control and directional stability and, at the same time, rudder overbalance. If
this is encountered the yaw must be reduced by banking in the direction in which the aircraft is yawing
and not by stabilizing the yaw by instinctively applying opposite bank. The correct action reduces the
sideslip by converting the motion into a turn from which recovery is possible once the fin has
unstalled. Sometimes slight apparent rudder overbalance may be noticed under asymmetric power
when large amounts of rudder trim are used to decrease the foot load on the rudder bar. If this
happens the amount of rudder trim should be reduced.

51. Rudder Tramping. On some aircraft the onset of rudder overbalance may be shown by
‘tramping’, or a fluctuation in the rudder foot loads. If the yaw is further increased overbalance may
occur.

52. Elevons and Tailerons. Some aircraft with swept-back wings combine the function of the
elevators and ailerons in control surfaces at the wing tips called elevons, or if tail mounted, tailerons.
These are designed so that a backward movement of the control column will raise both surfaces and
so, acting as elevators, they will raise the nose. If the control column is held back and also moved to
one side, the surfaces will remain in a raised position and so continue acting as elevators, but the
angular position of each surface changes so that the lift at each wing tip is adjusted to cause a rolling
moment in the direction that the control column has been moved. If the control column is held central
and then moved to one side the surfaces act as normal ailerons and a subsequent forward movement
of the control column will lower both surfaces, maintaining the angular difference caused by the
sideways movement of the control column, and the nose will drop while the aircraft is rolling, e.g.
Mirage-2000 has elevons, while the Jaguar has tailerons.
FIS Book 1: Aerodynamics 204

53. The All-Moving (Slab) or Flying Tail. At high Mach numbers the elevator loses much of its
effectiveness for reasons given in the chapters on High Speed Flight. This loss of effectiveness is the
cause of a serious decrease in the accuracy with which the flight path can be controlled and in the
manoeuvrability. To overcome this deficiency the entire tailplane can be made to serve as the
primary control surface for control in the looping plane. When this is done some form of power
assistance is usually employed to overcome the higher forces needed to move the tailplane during
flight. With the flying tail, full and accurate control is retained at all Mach numbers and speeds.
Forward movement of the control column increases the incidence of the tailplane to obtain the upward
force necessary to lower the nose. On some aircraft the elevator is retained and is linked to the
tailplane in such a way that movement of the tailplane causes the elevator, by virtue of its linkage, to
move in the usual direction to assist the action of the tailplane. When no elevator is used the whole is
known as a slab tailplane, e.g. the Mig-21, Mig-27, Jaguar etc.

54. The Variable Incidence (VI) Tail. This is used on some aircraft as an alternative, and
sometimes in addition, to trimming tabs. By suitably varying the incidence of the tailplane any out-of-
trim forces can be balanced as necessary. The VI tailplane is generally more effective than tabs at
high Mach numbers. Its method of operation is usually electrical with the control being a switch which
is spring-loaded to a central off position, e.g. Kiran aircraft.

The Vee Tail

55. The Vee, or butterfly tail is an arrangement whereby two surfaces forming a high dihedral
angle perform the functions of the conventional horizontal and vertical tail surfaces. The effective
horizontal tail area is the area of both surfaces projected on the horizontal plane. The effective vertical
tail area is the area of both surfaces projected on vertical plane. As the lift force from each surface
acts normal to its span line, the vertical component acts to provide a pitching moment, and the
horizontal component acts to produce a yawing moment. The two moveable portions therefore are
capable of performing the functions of both the elevator and rudder. They are sometimes called a
ruddervator. It can be seen that if both surfaces are moved up or down by an equal amount, the net
result is a change in the vertical force component only, and a pitching moment only results. If the two
surfaces are moved equal amounts in opposite directions, the result is a change in the net horizontal
force component only, and a yawing moment only results. Any combination of the two movements
results in combined pitching and yawing moments.

56. The elevator control in the cockpit is connected to give an equal deflection in the same
direction. The rudder control is connected to give equal deflections in opposite directions to the two
surfaces. The two cockpit controls are connected to the two surfaces through a differential linkage or
gearing arrangement. Thus the Vee tail performs the horizontal and vertical tail control functions with
normal cockpit controls. Some of the advantages claimed are:
205 Flight Controls

(a) Weight saving as less total tail surface required.


(b) Performance gain as less total tail surface and lower interference drag, as only 2
surfaces intersect the fuselage.
(c) Removal of the tail from the wing wake and downwash.
(d) Better spin recovery, as the unblanketed portion of the tail acts both to pitch the
aircraft down and to stop the rotation.

AIRBRAKES

General

57. Jet engined aircraft, having no propeller drag when the engine is throttled back, have
comparatively low drag and lose speed only slowly. Further, having eventually reached the desired
lower speed, any slight downward flight path causes an immediate and appreciable increase in speed.

58. An air brake is an integral part of the airframe and can be extended to increase the drag of an
aircraft at will, enabling the speed to be decreased more rapidly, or regulated during a descent. On
some aircraft the undercarriage may be lowered partially or completely to obtain the same effect.

59. Although the area of the air brakes on a typical fighter is small, considerable drag is produced
at high speeds. For example, an air brake with an assumed CD of 1.2 and a total area of about 0.4 m2
when operated at 1000 kmph at sea level produces a drag of about 22,500 N. This figure is indicative
of the large loads imposed on an aircraft when flying at high indicated speeds. The effectiveness of
an air brake varies as the square of the speed and therefore at about 250 kmph the same air brake
gives a drag of about 1400 N only. The decelerating effect of air brakes can be seen from figures
obtained from an aircraft flying at 800 kmph at low altitude. With the air brakes in and power off the
aircraft takes 2 min 58 sec to slow down to 300 kmph, but with air brakes out the time is reduced to 1
min 27 sec.

60. Ideally, air brakes should not produce any effect other than drag, although on some aircraft
the air brakes are designed to produce an automatic nose-up change of trim when opened. In
practice, however, the opening of most air brakes is accompanied by some degree of buffet, with or
without a change of trim. The strength of these adverse effects is usually greatest at high speeds,
becoming less as the speed decreases.

Effect of Altitude on Airbrake Effectiveness

61. Air brakes derive their usefulness from the fact that they are subjected to dynamic pressures,
the ½ρV2 effect, and so provide drag in proportion to their EAS. At high altitudes therefore, the
FIS Book 1: Aerodynamics 206

effectiveness of all air brakes is much reduced since the drag, which is low in proportion to the TAS,
takes longer to achieve a required loss in speed, i.e. the rate of deceleration is reduced.

62. An air brake which develops, say, 10,000 N drag at a stated IAS/TAS at sea level will develop
the same drag at the same IAS at high altitude, but whereas the TAS at sea level was equal to the
IAS, the TAS at altitude may be as much as 2 or more times the IAS. Since the drag is required to
decrease the kinetic energy, which is proportional to the TAS, it is apparent that the decelerating
effect of the air brake (proportional to the IAS) is decreased, e.g. the time taken to decelerate over a
given range of IAS will be doubled at 40,000 ft compared to that at sea level since the IAS at 40,000 ft
is about half the TAS.

BRAKE PARACHUTE

General

63. Brake parachutes are used to supplement the aircraft’s wheel brakes and so reduce the
length of the landing run. In general they produce enough drag to cause a steady rate of deceleration
varying from about 0.25 g to 0.35 g depending upon the particular installation. Below 125 kmph
(about 60 to 70 kt) the drag, varying as the square of the speed, falls to a much lower figure and the
wheel brakes become the primary means of deceleration.

Parachute Diameter

64. The diameter of the parachute depends on the weight and


size of the aircraft. For aircraft with a landing weight of around 5,000
kg, the ‘flying’ diameter of the parachute is from 2 - 3 m. At a
touchdown speed of 240 kmph this gives a drag of about 1,250 kg
and a deceleration rate of about 0.25g.

65. For large aircraft with landing weights around 50,000 kg the
flying diameter of the brake parachute is about 11 m. This produces,
at a touchdown speed of about 280 kmph, a drag of some 25,000 kg
and an initial rate of deceleration of about 0.35 g.

Cross-wind Landings

66. When used in a cross-wind landing the parachute aligns itself


along the resultant between the vectors representing the forward
speed of the aircraft and the 90° component of the cross-wind vector Fig 12-12: Effect of
(Fig 12-12). This causes a yawing moment, which increases the Cross-wind.
207 Flight Controls

weather-cock characteristics of the aircraft. For 90° cross-wind components of about 35 kmph the
effect is small and can easily be countered by the pilot. At higher cross-wind speeds it becomes
progressively more difficult to keep the aircraft straight. As the aircraft decelerates, the flying angle
between the parachute and the centre line of the aircraft increases, but the retarding force of the
parachute will decrease with the reduction in V2. Thus the overall effect of decreasing speed is a
reduction in the weather cocking tendency due to the brake parachute.

67. Any difficulty in directional control tends to occur fairly early in the landing run, although not
necessarily at the highest speed. This is mainly due to the very small flying angle and the relatively
high rudder effectiveness. Tyre sideslip due to the side load imposed by the parachute may be
another significant factor.

68. Jettisoning of Brake Parachute. At any time after the parachute has been streamed it can
be disconnected or jettisoned by the pilot in an emergency. It is also usual to jettison the parachute at
the end of the landing run.

69. Inadvertent Streaming in Flight. If the parachute is opened inadvertently at high speed
the opening load causes failure of a weak link and the parachute breaks away from the aircraft. If
opened too early on the approach the pilot can disconnect it.

SUMMARY

Flight Controls

70. The main control surfaces are:

(a) Rudder.
(b) Elevator.
(c) Aileron.

On some aircraft the effect of two of these controls is combined in a single set of control surfaces, e.g.
elevons, ruddervator and elevons (tail mounted ‘tailerons’).

71. Control power and effectiveness is decided by:

(a) Size and shape of the control.


(b) Deflection angle.
(c) EAS2.
(d) Moment arm.
FIS Book 1: Aerodynamics 208

72. Aerodynamic balance is a means of reducing the hinge moment and thereby reducing the
physical effort experienced in controlling an aircraft. The most common forms of aerodynamic
balance are:

(a) Horn balance.


(b) Inset hinge.
(c) Internal balance.
(d) Various types of tab balance.

73. Flutter can be prevented by:

(a) Mass balancing.


(b) Increasing structural rigidity.
(c) Irreversible controls.

74. The main considerations for control requirements are:

(a) Control force.


(b) Control movement.
(c) Control harmony.

75. Different control responses are given different terms. The ailerons are normally described as
a rate control, and the rudder and elevator as displacement controls.

76. Methods of overcoming adverse aileron yaw are:

(a) Differential ailerons.


(b) Frise ailerons.
(c) Control coupling.
(d) Spoilers.

77. Some advantages claimed for the Vee tail are:

(a) Weight saving.


(b) Performance gain.
(c) Removal of the tail from the wing wake and downwash.
(d) Better spin recovery.
209

CHAPTER 13

TRIMMING AND BALANCE TABS

Introduction

1. The chapter on flight controls covered as to how the controls can be balanced and also how
the control forces can be reduced with the help of aerodynamic balance. This chapter is really an
extension of the same, but deals mainly with tabs and their operation. Various control surface tab
devices can be utilized to modify the control forces.

2. A tab is a small hinged surface


forming part of the trailing edge of a
primary control surface. The principle of
operation is described below and also
shown in Fig 13-1.

3. Assume that the aircraft is


slightly tail heavy and therefore requires
a constant push force to maintain level
flight. If the position of the elevator
trimming tab is then adjusted so that it is
moved upwards the result is a download Fig 13-1: Principle of Operation of Balance Tab
on the elevator trailing edge, which
moves the elevator down. The downward movement continues until the downward moment of the tab
is balanced by the upward moment of the elevator.

Fixed Tabs

4. Fixed tabs can only be adjusted on the


ground and their setting determined by one or more
test flights. Thus under conditions of no stick force,
the trailing position of the control surfaces is
governed by the position of the tab. This type of tab
is used on the ailerons of some aircraft (Fig 13-2
and 3) and also on the rudders of single engine jet
aircraft.

5 An early form of fixed tab, still used on light


Fig 13-2: Fixed Tab on the Aileron of
aircraft, consists of small strips of cord doped above Deepak Aircraft
FIS Book 1: Aerodynamics 210

or below the trailing edge of the control surface. A strip of


cord above the trailing edge deflects the surface downwards,
the amount of deflection depending on the length and
diameter of the cord used.

Trim Tabs

6. Trim tabs are used to trim out any holding forces


encountered in flight, such as those occurring after a change
of power or speed, or when the CG position changes owing
to fuel consumption, dropping bombs or expending
ammunition. Whenever the speed, power or CG position is
altered, one at a time or in combination, the changed trim of
the aircraft necessitates resetting of the tabs (Fig 13-3). Fig 13-3: Fixed and Trim Tab

7. Operation of Trim Tabs. The tabs are pilot operated, either by hand wheels in the cockpit
or electrically. The hand wheels always operate in the natural sense, i.e. the top of the elevator trim
hand wheel is moved forward to produce a nose-down trim change and vice versa. Tabs may also be
electrically operated by small switches, spring-loaded to the central (off) position. To eliminate a
holding force the pilot moves the appropriate switch in the same direction as the holding force until the
stick load is zero, then releases the switch.

8. Variation of Tab Effectiveness with Speed. The sensitivity and power of a trim tab varies
with speed in the same way as the control surfaces. At low speeds large tab deflections may be
required to trim an aircraft, but at high speeds small movements have a marked and immediate effect
on the trim.

Other Methods of Trimming

9. Spring-Bias Trimmers. Trimming is also done on some light aircraft by adjusting the
tension in springs, which are connected directly to the pilot’s controls. Thus an adjustment in the
spring tension will bias the control column or rudder bar in any desired position.

10. Variable-Incidence (VI) Tail planes and All Flying (Slab) Tail planes. These are used to
counter large changes in trim in the pitching plane. They were discussed in the chapter on Flight
Controls.

11. Variable-Incidence Wings. Strictly speaking the VI wing is not a trimming device at all but
a means of altering the attitude of the aircraft in flight. On swept-wing aircraft which utilize high lift
devices on the leading edge, the net effect is to increase the stalling angle. A disadvantage of this is
211 Trimming and Balance Tabs

the nose high attitude of the aircraft in the take-off and landing configuration, with attendant poor
cockpit vision and the requirement of long undercarriages. The variable incidence wing overcomes
this disadvantage by the use of a variable rigging angle of incidence, controllable by the pilot, which
allows a large angle of attack on the wing without a nose high attitude to the fuselage.

BALANCE TABS

General

12. Balance tabs are used to balance or


partially balance the aerodynamic load on the
control surfaces, thus reducing stick loads. In its
basic form this type of tab is not under the
control of the pilot, but the tab angle is changed
automatically whenever the main control surface
is moved. The cross-sectional shape of the
control surface may have an important effect on
aerodynamic balance. A convex shape can tend
towards overbalance whilst a concave shape
("hollow-ground") can have the opposite effect.
The latter shape is sometimes achieved by
fitting narrow flat plates to the trailing edge of the Fig 13-4 a & b: Balance and
Anti-Balance Tabs
tab to give the "hollow-ground" effect.

Geared Tabs

13. This tab is linked with the fixed surface ahead of the control surface and is geared to move at
a set ratio, in the opposite direction to the movement of the control (Fig 13-4 a). Since the tab moves
in the opposite sense to the control
surface, it produces a small loss of lift,
but the reduction in control effectiveness
is usually negligible.

14. It is sometimes necessary to


incorporate a large horn balance to
improve the stick-free stability, and it
may then be necessary to fit an anti- Fig 13-5: Balance Tabs on the Aileron of HPT 32 Ac
balance tab to increase the stick force
(Fig 13-4 b)
FIS Book 1: Aerodynamics 212

15. Servo-Tabs. These tabs are connected directly to the cockpit controls and the tab can be
made to apply the hinge moment required to move the control surface. The pilot’s control input
deflects the tab and the moment produced about the hinge line of the control surface causes this
surface to "float" to its position of equilibrium. The floating control will then produce the required
moment about the CG of the aircraft. The stick forces involved are only those arising from the hinge
moments acting on the tab, which are much less than those on the main control surface. A schematic
diagram of a servo-tab is shown in Fig 13- 6. One of the chief disadvantages of the simple servo-tab
is that it lacks effectiveness at low speeds, since any tab loses most of its effect when deflected
through an angle of more than about 20°. The large
control surface deflections required at low speeds
need correspondingly large tab deflections and
therefore this system is not always satisfactory. The
spring tab, described in Para 16, is used to
overcome this fault.

Spring Tabs
Fig 13-6: Servo-Tab System

16. The spring tab is a modification of the


geared tab in which a spring is incorporated in the
linkage permitting the tab gearing to be varied
according to the applied stick forces. The actual
arrangement of the mechanism may vary with
different designs according to the space limitations
where it is to be applied. A schematic arrangement
of a spring tab is shown Fig 13-7.
Fig 13-7: Spring Tab System

17. Movement of the input rod deflects the tab against spring tension. The input force is
transmitted through the spring to the control surface, which moves as a result of the combined effect
of input force and aerodynamic assistance provided by the tab.

18. The amount of servo-action depends on the rate (strength) of the spring employed. It can be
seen that an infinitely strong spring produces no assistance from the tab whereas an infinitely weak
spring causes the tab to behave as a servo-tab. The effect of the spring tab can be regarded as
producing a change in the basic speed squared law.

19. Spring tabs may be pre-loaded to prevent them from coming into operation until the stick (or
rudder) force exceeds a predetermined value. This is done to keep the spring tab out of action at low
speeds, thus avoiding excessive lightening and lack of feel. A disadvantage of this arrangement is
the sudden change in stick force gradient that occurs when the tab comes into operation.
213 Trimming and Balance Tabs

20. The spring tab has been widely used in modern aircraft with much success. On very large
surfaces however the spring tab or servo-tab may produce a "spongy" feel of controls to which the
pilot may object. On ground, when the control column is held in one position, the control surfaces can
be moved by hand.

21. Dual Purpose Tabs. In some aircraft the features of two or more of the foregoing types of
tabs are combined. For example, servo, geared or spring tabs can be connected to the pilot-operated
trim wheel so that the basic position of the tab in relation to the control surface can be varied. In this
way the tab functions both as a trim tab and as a pure servo, geared or spring tab.

Control Locks on Servo and Spring Tab Control Surfaces

22. On aircraft using servo or spring tabs, locking the control column or rudder bar will not prevent
high winds from moving the control surfaces. In these cases external control surface clamps are
essential.

23. On certain spring tab installations, partial or full movement of the control column or rudder bar
is possible when the external clamps are still in position. For this reason it is vital that the clamps are
removed before flight and a visual check of the correct movement of the control surfaces is made
when the control column is moved. If the surface cannot be seen from the cockpit, movement should
be checked with the assistance of the ground crew. Any restriction in the movement of the control
column, other than the normal friction, must be investigated before flight.

24. Following images show various tabs used on HPT-32 and HJT-16 aircraft.

Fig 13-8: Various Tabs and Balances on the Rudder Fig 13-9: The Fixed and Tab Balances
and Elevator of Deepak Aircraft on the Aileron of Kiran Aircraft
FIS Book 1: Aerodynamics 214
215

CHAPTER 14

STABILITY

Introduction

1. The study of aircraft stability can be extremely complex and so for the purpose of this chapter
the subject is greatly simplified. Stability is first defined in general terms and it will thereafter be seen
how the aircraft designer incorporates stability into an aircraft. The subject is dealt under two main
headings Static stability and Dynamic stability.

STATIC STABILITY

Definitions

2. To quote Newton’s first law again, "a body will tend to remain in a state of rest or uniform
motion unless disturbed by an external force". When such a body is so disturbed, stability is
concerned with the motion of the body after the external force has been removed. Static stability
describes the immediate reaction of the body after disturbance, while dynamic stability describes the
subsequent reaction. The response is related to the original equilibrium state by use of the terms
positive, neutral and negative stability. Positive stability indicates a return towards the position prior to
disturbance, neutral stability indicates taking up of a new position of a constant relationship to the
original, whereas negative stability indicates a continuous divergence from the original state. Note
that in colloquial usage positively stable and negatively stable are usually "stable" and "unstable"
respectively.

3. In order to relate the response of a body to its initial equilibrium state it is useful at this stage
to use an analogy of the "bowl and ball" which can be used to illustrate this idea (Fig 14-1). If the ball
is displaced from its initial position to a new position, the reaction of the ball will describe its static
stability. If it tends to roll back to its original position, it is said to have positive stability, if it tends to
roll further away from its original position, it has negative stability and if the ball tends to remain in its
new position, it has neutral stability.

Fig 14-1: Static Stability Analogies


FIS Book 1: Aerodynamics 216

4. The concept of stability "degree"


can be expressed more usefully in
graphical form (Fig 14-2). Displacement,
plotted on the vertical axis may refer to
any system, e.g. distance, moments,
volts, etc. No scale is given to the
horizontal axis, which may vary from
microseconds to hours, or even years.

5. Plotting the response in this form


Fig 14-2: Graphs of the Degrees of Stability
makes it possible to measure the actual
degree of stability using the following two
parameters:

(a) The sign of the slope


indicates whether the response is
favorable or unfavorable.

(b) The slope of the curve is


a measure of the static stability.

6. Before considering the response


of the aircraft to disturbance it is
necessary to resolve the motion of the
aircraft into components about the three Fig 14-3: Aircraft Reference Axes

body axes passing through the CG (Fig 14-3).

Axis Longitudinal (x) Lateral (y) Normal (z)

Motion (About the Axis) Roll (p) Pitch (q) Yaw (r)

Stability Lateral Longitudinal Directional (Weathercock)

7. It is important to realize that the motion involved is angular velocity and the disturbance
assumed is an angular displacement. In the first instance it is helpful to consider these components
separately although, in other than straight and level flight, motion of the aircraft is more complex, e.g.
in a level turn the aircraft is pitching and yawing.

8. A reasonable degree of stability is essential for an aircraft around its three axes if it is to be
handled with comfort & ease in flight. However this does not mean that an unstable aircraft cannot be
217 Stability

flown at all. If the instability is slight with a low rate of divergence, the aircraft can be maintained under
control with quick reactions of the pilot and effective control input on the aircraft. Such an instability
would require constant monitoring and frequent control inputs to maintain the aircraft along its desired
flight path. However, if some design features or computer control is incorporated, any aircraft can be
made reasonably stable. This chapter deals only with the aerodynamic design features incorporated
in aircraft to achieve stability in subsonic flight.

9. It is quite possible that the stability characteristics of a given aircraft vary with speed, altitude
and configuration. There have been practical examples of aircraft which were fairly unstable in straight
and level flight but extremely stable in inverted flight. In such aircraft, a roll out from the inverted flight
has been found to be difficult, if not impossible with the available aileron power. However, stability
should not be confused with balance. For example, there may be an aircraft which flies steady with a
wing low. If you “disturb" the aircraft from this condition to bring it to wings level flight, it may even
return to its original wing low condition after the “disturbing force" is removed. Such an aircraft may be
out of balance, but it is certainly not unstable. Similarly an aircraft that has a rolling tendency during
level flight, with the wing continuing to drop, is unstable in level flight.

DIRECTIONAL STABILITY

10. In our study, we will discuss the contributions of various components (aerodynamic) of an
aircraft towards static stability. However, numerical values can not be given and hence, the words
"stable", "less stable" and "unstable" would be used to correlate and compare the relative merits and
demerits.

11. A simple approach to both directional


and longitudinal stability is to consider a
simple dart. The flights or vanes of a dart
ensure that the dart is aligned with the flight
path. Consider first the pair of vanes which
impart positive directional stability to the dart,
these may be referred to as the vertical
stabilizers. Fig 14-4 shows how a
displacement in yaw through an angle β,
resulting in sideslip, produces a restoring
moment and therefore positive directional
Fig 14-4: The Positive Stability of a Dart
(static) stability. Two points are worth noting:

(a) The dart rotates about the centre of gravity (CG).


(b) The momentum of the dart momentarily carries it along the original flight path, i.e. the
Relative Air Flow (RAF) is equal and opposite to the velocity of the dart.
FIS Book 1: Aerodynamics 218

12. An aerodynamic shape like a


fuselage or drop-tank may be unstable.
Reference to Fig 14-5 shows that this occurs
when the centre of pressure (CP) is in front
of the CG. It is necessary therefore to add a
vertical stabilizer or "fin" to produce positive
Fig 14-5: The Negative Static Stability of a
directional stability and this has the effect of Streamline Body when CP is Ahead of CG
moving the CP behind the CG, as depicted
on the MiG 21 (Fig 14-6). In general, it may
be said that the keel surface of the fuselage
(ahead of the CG) has an unstable influence,
while the keel surface behind the CG has a
stable influence. For simplicity, the rudder is
considered to be "locked".

13. The directional stability of an aircraft


is essentially weathercock stability and
involves moments about the normal (vertical)
axis and their relationship with yaw and side
slip angle. Consider an aircraft disturbed
from a state of equilibrium (Fig 14-6) i.e.
temporarily deflected from its course. If the
aircraft develops yawing moment as result of
this deflection in such a manner so that the
aircraft tends to return to its original state of
equilibrium than it possesses positive static Fig 14-6: Contribution of Aeroplane
Components Towards Directional Stability
directional stability.

14. Before moving on let us see the


definition of yaw and sideslip which will help
us in understanding the directional stability
and further the lateral stability better.

(a) Sideslip Angle. It is the


angle between the aircraft centerline
and the RAF and is considered
positive when the RAF is from the
right. It is denoted by β (Fig 14-7).
Fig 14-7: Yawing & Sideslip
219 Stability

(b) Yaw angle. It is the angle formed due to displacement of the aircraft centerline
angle from some reference azimuth and is positive when the nose of the aircraft is displace to
the right of the reference azimuth. It is denoted by psi (Ψ) (Fig 14-7).

15. From the definitions of yaw and sideslip angle it is evident that there is no direct relationship
between the two for an aircraft in free flight. For example, an aircraft turned through 360 deg has
yawed through 360 deg but the side slip may have been zero through the entire turn. Thus static
directional stability of an aircraft can be best appreciated by response to a given sideslip. The static
directional stability of an aircraft can be
studied by a graph of yawing moment
co-efficient Cn versus sideslip angle β
as shown in Fig 14-8. From the graph it
is evident that when an aircraft is
subjected to a positive sideslip angle,
static directional stability will be evident
if positive yawing moment coefficient
results. From the graph it can be seen
that directional stability will exist if the
curve of Cn versus β has a positive Fig 14-8: Directional Stability

slope and the degree of the stability will be function of this slope. Thus when the RAF comes from the
right, positive yawing moment, i.e. to the right should be established, which would tend to return the
nose of the aircraft back into wind and thus exhibit positive directional stability. From the graph it can
also be seen that at small to moderate angles of sideslip strong positive slope indicates strong
positive directional stability. However, at large angle of sideslip there is a sudden decay in directional
stability producing neutral stability to directional instability.

16. Contribution of the Aeroplane Components. Static directional stability must be in


evidence for all the critical conditions of flight. Generally, good directional stability is a fundamental
quality directly affecting the pilots' handling of an aeroplane. The static directional stability of the
aeroplane is a result of contribution of each of the various aeroplane components. While the
contribution of each component is somewhat dependent upon and related to other components, it is
necessary to study each component separately.

(a) The Vertical Tail. This is the primary source of directional stability for the
aeroplane. As shown in Fig 14-6, when the aeroplane is in a sideslip the vertical tail will
experience a change in angle of attack. The change in lift or side force on the vertical tail
creates a yawing moment about the center of gravity which tends to yaw the aeroplane into
the RAF. The magnitude of the vertical tail contribution to static directional stability then
depends on the change in tail lift and the tail moment arm. Obviously, the tail moment arm is
a powerful factor but is essentially dictated by the major configuration properties of the
aeroplane.
FIS Book 1: Aerodynamics 220

(b) Wing. The contribution of the wing to static directional stability is usually small. The
swept wing provides a stable contribution depending upon the amount of sweepback but the
contribution is relatively weak when compared with other components.

(c) Fuselage and Nacelles. The contribution of the fuselage and nacelles is of primary
importance since these components furnish maximum destabilizing influence (Fig 14-6). The
subsonic Centre of pressure of the fuselage will be located at or ahead of the quarter length
point. Since the CG of the aircraft is usually well aft of this point, the fuselage contribution will
be destabilising. However at large angles of sideslip the large destabilizing contribution of the
fuselage reduces.

17. The contribution of the vertical tail alone is highly stabilizing upto the point where the surface
begins to stall. The contribution of the vertical tail must be large enough so that the complete aircraft
displays the required degree of stability. Thus for a given displacement and therefore a given side slip
angle the degree of positive stability would depend on the size of restoring moment which is mainly
determined by:

(a) Design of the vertical stabilizer


(b) The moment arm

18. Design of the Vertical Sabilizer. The vertical stabilizer is a symmetrical aerofoil and will
produce an aerodynamic force at positive angles of attack. In sideslip therefore the total side force on
the fin and the rudder will be proportional to lift
coefficient and the surface area. Increasing the
surface area though would increase the lift but would
also result in increased drag. The lift coefficient will
vary like any aerofoil with aspect ratio and
sweepback. Though it is desirable to have high lift
curve slope but because it would have lower stalling
angle, practically it is not desirable. Thus there is a
need for low aspect ratio and sweep back on the fin
(Fig 14-9). In supersonic flight since the there is
decay in the lift curve slope there is a need for large
vertical surfaces. Typical examples being SU-30 and
MiG-29 aircraft which have twin and large vertical
fins in order to have sufficient directional stability at
higher mach numbers. Ventral fins (Jaguar, Mig 21,
23 & 27) also increase the surface area behind the
aircraft CG without altering the aspect ratio or the Fig 14-9: Vertical Stabilizers

angle of sweep of the fin.


221 Stability

19. Moment Arm. The position of CG and therefore the distance between the CG and the CP
of the vertical stabilizer would also play a vital role in determining the directional stability of an aircraft
(Fig 14-6). The greater the distance between the two, greater is the directional stability and thus
smaller the size of the vertical surface needed. Forward movement of CG would therefore increase
directional stability and its aft movement would decrease it. The requirement of large vertical surfaces
on the SU-30 and MiG-29 ac is also due to the CG being situated in an aft position.

LONGITUDINAL STABILITY

20. The analogy used in para 11 can usefully be used to introduce the concept of static
longitudinal stability. In this case the dart is viewed from the side and the horizontal stabilizers
produce a pitching moment (M) tending to reduce the displacement in pitch.

21. On an aircraft, the tailplane and elevators perform the functions of a horizontal stabilizer and
the conclusions reached in para 16 will be equally valid. For simplicity, the explanation is limited to
stick fixed static stability, i.e. elevators locked.

22. Fig 14-10 (a) shows a wing with the CP forward of the CG by the distance y. A nose-up
displacement will increase the angle of attack, increase the lift (L) by the amount dL and increase the
wing pitching moment by the amount dLy. The result is to worsen the nose-up displacement, an
unstable effect. In Fig 14-10(b), the CP is aft of the CG and the wing moment resulting from a
displacement in pitch will be stabilizing in effect. The pitching moment is also affected by the
movement of the CP with angle of attack and it follows therefore that the relative positions of the CP
and CG determine whether the wings have a stable or unstable influence. Taking the worst case,
therefore, the wing may have an unstable influence and the horizontal stabilizer must be designed to
overcome this.

Fig 14-10: Variations in the Position of CP and CG

23. It can therefore be said that an aircraft will exhibit positive static longitudinal stability if it tends
to return to the trim angle of attack when displaced by a gust or inadvertent control inputs. The
aircraft, which is unstable, will continue to pitch in the disturbed direction until the displacement is
FIS Book 1: Aerodynamics 222

resisted by opposing control forces where as a stable aircraft will tend to pitch in the opposite direction
i.e. pitch down if there is an increase in angle of attack.

24. Since static longitudinal stability depends upon the relationship of angle of attack and pitching
moment, it is necessary to study the pitching moment contribution of each component of the aircraft.
In a manner similar to all other aerodynamic forces, the pitching moment is studied in the co-efficient
form:

M = CM q S C Where: M is pitching moment about CG


q is dynamic pressure
S is wing surface area
C is Mean Aerodynamic chord and
CM is pitching moment co-efficient

25. If we sum up the pitching moment coefficients contributed by all the components of the
aircraft and plot it against lift coefficient, a study of this CM Vs CL graph will relate the static
longitudinal stability of the aircraft (Fig 14-11). The graph shows approximate values attainable for an
aircraft as a whole for positive values of lift coefficient. In the curve marked as "stable", it can be seen
that any increase in CL (say by increasing α ), results in a nose down pitching moment, thus reducing
the α & the CL to the original value. Thus positive static longitudinal stability is indicated by a negative
slope of CM versus CL. In the unstable case, the
reverse occurs and once an increase in CL
occurs due to some reason, it continues to
increase. In the third case of neutral stability, it
achieves an equilibrium at the new value
because there is no change in the CM with CL.
The Trim Point indicates a value of CL for which
CM is zero, that is, the aircraft will continue to Fig 14-11: CM Vs. CL Slopes Depicting
maintain the CL. Aircraft Static Longitudinal Stability

26. Ordinarily, the static longitudinal stability


of a conventional aircraft does not vary with CL..
However, if the aircraft has sweepback or large
contribution of power effects to stability or
significant changes in downwash at the
horizontal tail, noticeable changes in static
stability can occur at high lift coefficients. A
typical graph of a swept back wing displaying Fig 14-12: Static Longitudinal Stability of
Swept Back Wing Aircraft
the "pitch up" instability phenomenon at high α
is as shown in Fig 14-12.
223 Stability

27. The factors affecting aircraft static longitudinal stability are as follows:

(a) Contribution from the wing.


(b) Effect of fuselage.
(c) Design of the tailplane.
(i) Distance from CP tail to CG aircraft.
(ii) Tail area.
(iii) Tail volume.
(iv) Plan form.
(v) Wing downwash.
(d) Longitudinal dihedral.
(e) Position of CG.

Contribution from the Wing

28. This depends upon the relative position of AC of the wing to CG of the aircraft. All changes in
CL effectively take place at the wing AC. Thus, if the wing experiences some change in CL, the
pitching moment about the CG will vary unless the AC and the CG are co-located. A little imagination
will make it clear that CG should be ahead of the AC for the wing to contribute towards positive
stability. If CG happens to be behind AC, unstable contribution, and if they are co-located a neutral
contribution from the wing can be expected. Since the wing is the most important aerodynamic
surface in an aircraft, any change in its contribution will produce a significant change in aircraft
stability. This fact would be most apparent in case of a tailless aircraft. For such an aircraft to achieve
positive stability, CG must be ahead of AC. The
wing must also have a positive pitching moment
about AC to achieve trim at positive values of CL.
We know from the basic theory that that CM for
most aerofoil about the AC is negative at zero lift.
That means, for such an aerofoil to be stable with
the CG at AC, there will be no positive value of
CL where CM is zero. Thus, in order to achieve
trim at a positive value of CL, the aircraft tail has
to produce a down load at all values of CL for the
aircraft to fly in straight & level trimmed flight.
The only means available to achieve trim at
positive CL with a wing having negative CM about
AC is an unstable CG position aft of the AC
(assuming that tail plane is not available).
Therefore, in a tailless aircraft, high lift devices at
the trailing edge, which incur any significant Fig 14-13: Wing Contribution
changes in CM, cannot be utilized (Fig 14-13).
FIS Book 1: Aerodynamics 224

29. Taking the worst case, therefore, the wing may have an unstable influence and the horizontal
stabilizer must be designed to overcome this. Note that the sentence says "Taking the worst case ".
Hence, it need not necessarily be the case always.

Effect of the Fuselage

30. In most cases the contribution of the fuselage and nacelles is destabilizing. A symmetrical
body of revolution in the flow field of an ideal fluid develops an unstable pitching moment when given
an angle of attack. Even though there may not be any appreciable lift being produced, an increase of
α produces an increase in the unstable
pitching moment. In the actual case of
subsonic flow, taking the fuselage to be
a symmetrical body, same effect is
produced essentially. Secondly, there is
an up wash ahead of the wing and
downwash behind, which increase with
increase in α. The upwash ahead of the
wing increases the destabilizing
influence from the portions of fuselage
ahead of the wing while the downwash
behind the wing, acting on the portions
of the fuselage behind it, reduces this
destabilizing influence. Hence the
location of the wing relative to the Fig 14-14: Pressure Distribution over the Fuselage
fuselage is important in determining the
contribution to stability (Fig 14-14).

31. In the study of supersonic aerodynamics, it will be seen that a body of revolution at a positive
value of α will also produce positive lift from the front portions (in addition to the unstable moment
described above). Since the fuselage CP is now well ahead, the body contributes a destabilising
influence. In supersonic design, the fuselage and nacelles may be quite large in comparison to the
wing and their contribution to stability will also be large. The formation of shockwaves and the
ineffectiveness of areas behind the shockwaves, also generally create problems of stability in
transonic flight.

Design of the Tailplane

32. Among all the components in the aircraft, the tail plane has the greatest influence in deciding
its stability characteristics. Usually, the tail plane is a symmetrical aerofoil (need not necessarily be)
and it creates lift like any other surface depending on its area, Aspect Ratio, angle of attack, camber,
225 Stability

etc. The simplified


diagram at Fig 14-15
illustrates a system of
forces due to
displacement in pitch, in
this case an increase in
angle of attack. The tail
contribution must
overcome the unstable
wing (and any other)
contribution for positive
static longitudinal Fig 14-15: Changes in Forces and Moments due to a Small Nose-
Up Displacement (dα)
stability. The degree of
positive stability for a given change in angle of attack depends upon the difference between the wing
moment and the tail moment; this difference is called the restoring moment, i.e.

(Total Lift Tail) X y - (Total Lift wing) X x = Net pitching moment.

33. The tailplane is an aerofoil and the lift force resulting from a change in angle of attack will be
proportional to the CL tail and the area. The increment in lift from the tail will depend upon the slope
of its CL curve and will also be affected by the downwash angle behind the wing (if the downwash
changes with angle of attack). The tail design features which may affect the restoring moment are
therefore:

(a) Distance from CPTAIL to CG (Moment Arm). From para 32, greater the value of y
(distance from aircraft CG to the CP of tail plane), better the stability.

(b) Tail Area. The total lift provided by the wing = CL WING qS and the total lift produced
by the tail = CL TAIL qS. For a given aerofoil of given planform, the CL varies with angle of
attack at a constant q (EAS). Therefore, in comparing tail moments with wing moments, it is
necessary only to compare the respective area(s) and moment arms (CG position).

(c) Tail Volume. The product of the tailplane area X moment arm is known as the tail
volume. The ratio of the tail volume to the wing volume is the main parameter used by the
designer in determining the longitudinal stability of the aircraft.

(d) Planform. As was seen in the chapter on wing planforms, aspect ratio, taper and
sweepback affect the slope of the CL curve for a lifting surface. The planform of the tailplane
therefore affects the change in CL with change in angle of attack caused by a disturbance.
For example, the CL increments will be lower on a swept-back tail than on one of rectangular
planform.
FIS Book 1: Aerodynamics 226

(e) Wing Downwash. Usually, the location of the tail is such that it is almost always
subjected to a slight downwash coming from the wing. In addition, due to the effect of any
slipstream or any turbulence created by the wing boundary layer, the lift force for the tail plane
may be far different from that experienced by the wing. All these factors generally tend to
make tail less effective. Where a disturbance in angle of attack results in a change in the
downwash angle from the wings, the effective angle of attack at the tail is also changed. For
example, if the aircraft is displaced nose-up and the α of the mainplane is changed by 100,
then this will also increase the downwash angle. This increased downwash angle will result in
a smaller (approx. 60) change in α on the tailplane. The total tail lift will not be as great as it
would otherwise have been and so the restoring moment is reduced. Moving the CG further
forward, thereby increasing the moment arm compensates for this decrease in stability.

34. Longitudinal Dihedral. Longitudinal dihedral means that the tailplane is set at a lower
angle of incidence than the mainplane, as illustrated in Fig 14-16. The effect of longitudinal dihedral is
shown in Fig 14-17 where it
can be seen that when the
mainplane is at zero lift angle
of attack, the tailplane is at a
negative angle of attack, at
Fig 14-16: Longitudinal Dihedral
which it produces a
downward force which tends
to pitch the aircraft nose-up.
When the mainplane is at a
small positive angle of attack,
such as is normally to be
expected in straight and level
flight, the tailplane produces
no pitching moment. When
the mainplane angle of attack
increases the tailplane
Fig 14-17: CM CG vs. CL – Tailplane Contribution
produces a nose-down
pitching moment. Thus if the
aircraft is disturbed in pitch, the tailplane will tend to restore it to level flight. The longitudinal dihedral
therefore provides a stabilizing influence.

35. Trim Point (Stick Fixed). The combined effects of wing and tail are shown in Fig 14-18.
The negative slope of the graph shows positive longitudinal stability. The amount of longitudinal
stability is measured by the slope of the graph. Trim point is the CL or angle of attack at which the
overall moment about the CG is zero. This is where the wing moment is equal to the tail moment and
227 Stability

the aircraft is then "in trim". The


angle of attack or level flight speed
at which this occurs is determined
by the degree of longitudinal
dihedral.

Position of the CG

36. The position of the CG may


be marginally under the control of
the pilot. From Fig 14-15 it can be Fig 14-18: Combined Effect of Wing & Tail
seen that its position affects the ratio of the tail moment to the wing moment and therefore the degree
of stability. In particular:

(a) Aft movement of the CG decreases the positive stability.

(b) Forward movement of the CG increases the positive stability. Because the position of
the CG affects the positive longitudinal stability, it also affects the handling characteristics in
pitch. The aerodynamic pitching moment produced by deflecting the elevators must override
the restoring moment arising from the aircraft’s positive stability, i.e. the stability opposes
manoeuvre. For a given elevator deflection there will be a small response in an aircraft with a
forward CG (stable condition) and a large response in an aircraft with an aft CG (less stable
condition).

37. Neutral Point. Aircrew Manuals for every aircraft give the permitted range of movement of
the CG. The forward position is determined mainly by the degree of manoeuvrability required in the
particular aircraft type. Of greater importance to the pilot is the aft limit for the CG. If the CG is
moved aft, outside the permitted limits, a position will eventually be reached where the wing moment
(increasing) is equal to the tail moment (decreasing). In this situation the restoring moment is zero
and the aircraft is therefore neutrally stable. This position of the CG is known as the neutral point.
The aft limit for the CG, as quoted in the Aircrew Manual, is safely forward of the neutral point. If the
loading limits for the aircraft are exceeded, it is possible to have the CG position on, or aft of, the
neutral point. This unsafe situation is aggravated when the controls are allowed to "trail", i.e. stick
free.

38. CG Margin (Stick Fixed). The larger the tail area, the larger the tail moment, and so the
further aft is the CG position at which the aircraft becomes neutrally stable. The distance through
which the CG can be moved aft from the quoted datum, to reach the neutral point, is called the static
or CG margin, and is an indication of the degree of longitudinal stability. The greater the CG margin,
the greater the stability, e.g. a training, or fighter aircraft, may have a margin of a few inches but a
large passenger aircraft may have a margin of a few feet. For stability in pitch it is necessary that
FIS Book 1: Aerodynamics 228

when the angle of attack is


temporarily increased a
correcting force will lower
the nose to reduce the
angle of attack. As
illustrated in Fig 14-19, if the
CG is behind the centre of
Fig 14-19: Variations in Position of CG and CP
pressure an increase in the
angle of attack results in a pitch-up or positive pitching moment coefficient. Thus the mainplane on its
own has an unstable influence. If the CG is forward of the centre of pressure the pitching moment
coefficient will be negative, a stable contribution.

39. Till now, we have studied the contribution of some components of an aircraft towards static
longitudinal stability. Some more factors that affect longitudinal stability are:

(a) Power.
(b) High lift devices.
(c) Elevator deflection.
(d) Control system friction.
(e) Manoeuvre.
(f) Altitude.

Power Effects

40. The Effect of power


may cause significant changes
in trim lift coefficient and static
longitudinal stability. Power
effects will be most significant
when the aeroplane operates
at high power and low speeds
such as the takeoff and go
around. These effects are to be
considered under two main
categories. Firstly, the direct
effect. This is caused by the
force created by the propulsion
unit and is straightforward. If
the thrust line is above the CG,
Fig 14-20: Direct Power Effects
it produces a stabilizing effect.
A propeller or intake duct located ahead of the CG contributes a destabilizing effect. A rotating
229 Stability

propeller, inclined to the flight path, causes a deflection to the airflow. The momentum change of the
slipstream creates a normal force at the plane of the propeller, similar to a wing, by deflecting the air
stream (Fig 14-20). The magnitude of the unstable contribution depends upon the distance from the
CG to propeller and is greatest at high power and high angle of attack. The normal force created at
the inlet of the jet engine contributes a similar destabilizing effect when the inlet is in front of the CG.
The effect is largest at high thrust and high angle of attack.

41. The indirect effects of power are usually


of a greater concern in a propeller powered
aircraft than in a pure jet. Due to the presence of
the propeller(s), the airflow over a part of the
wing, fuselage and the tail plane as applicable,
is modified (Fig 14-21). Quite often, there is a
change in the direction of airflow apart from the
increased magnitude of q. This effect thus
produces stability contributions quite different
from the pure aeroplane considered so far and
so, depending on the power output from the
propeller, the aircraft stability characteristics
vary. Even in a jet aircraft, at times, a
destabilising effect is produced by an induced
flow at the exhaust. This is usually more
prominent when the jet exhaust is below, the tail
plane. In cases where the jet is above the tail,
the contribution may even be destabilizing as
shown in the Fig 14-21. Summarizing the effects Fig 14-21: Indirect Effects of Power and
of power, both direct and indirect, a general Slipstream

reduction of static stability results at high power, high CL and low EAS combination. Because of the
magnitude of both direct and indirect effects of power, a propeller powered aircraft usually
experiences a greater effect than a jet aircraft.

42. High Lift Devices. In general, any high lift device in an aircraft when extended, tends to
increase the downwash angle at the tail and also reduces the dynamic pressure at the tail, both of
which are destabilising. However, they generally tend to reduce the unstable moment from the wing
itself. Overall effect of these depends on the aircraft design features (the exact position of these parts
on the aircraft). The usual effect is destabilising. Hence most critical forward neutral point is
experienced in a condition while the aircraft is attempting to 'go round' off a powered approach to
land. The power ON neutral point usually decides the most aft limit of CG position.
FIS Book 1: Aerodynamics 230

43. Elevator Deflection. The effect of


elevator deflection on the pitching moments is
shown in the Fig 14-22. As is clear, an upward
deflection of the elevator (negative camber for the
tailplane) increases the CL value where trim is
achieved and vice versa. However, there is hardly
any change in the slope of the curve, which means
that the elevator or stabilizer deflection does not
alter the tail contribution to stability. The graph
also implies that stick is to be brought back to
increase angle of attack and to trim at lower
airspeed and at higher speeds, down elevator Fig 14-22: Longitudinal Control

deflection is required for trim. If the aeroplane were to have stick position instability, it would call for an
aft movement of stick to trim at higher airspeed and vice versa.

44. Elevator Angle to Trim. If the angle of attack is increased from the trim point, the aircraft’s
longitudinal stability will produce a stable, nose-down pitching moment. To maintain the new angle of
attack, an equal and opposite moment, nose-up, will be required from the elevators. When this is
achieved, by raising the elevators, a new trim point is established, i.e. at a higher angle of attack on
the mainplane, the tail has been made to produce a greater nose-up moment by altering the effective
camber on the tail. The reverse applies when the angle of attack on the mainplane is reduced. This
does not usually affect the positive longitudinal stability.

45. Stick-Free Longitudinal Stability. When the elevators are allowed to 'float free', that is,
stick is not held rigidly, they may tend to streamline with the RAF at the tail when the horizontal tail is
given a change in α. If the tail is subject to an increased α and the elevator tends to 'float up', the
change in lift on the tail is less than that if the elevators had remained fixed, and hence tail
contribution to stability is reduced. Thus, the stick free stability of an aeroplane is usually less than the
stick fixed stability. However, in case the controls use very large horn balances, the opposite results
may be obtained. If the controls are fully powered and are actuated by an irreversible mechanism, the
surfaces are not allowed to float free and hence no difference exists between stick free and stick fixed
stability.

46. Control System Friction. Most aircraft control systems pose a certain amount of
resistance to movement due to the friction of the various components. This has the effect of having a
small band of trim settings for the same value of trim CL or a small band of trim speed (CL) for a
constant trim setting. A wide friction force band can completely mask the stick force stability when the
stick force stability is low. With this, we conclude longitudinal static stability and now we will proceed
to longitudinal Dynamic stability.
231 Stability

Manoeuvre Stability (Steady Manoeuvres Only)

47. In the preceding paragraphs the longitudinal static stability was discussed with respect to a
disturbance in angle of attack from the condition of trimmed level flight. A pilot must also be able to
hold an aircraft in a manoeuvre and the designer has to provide adequate elevator control appropriate
to the role of the aircraft. The following paragraphs consider the effects on an aircraft of a disturbance
in angle of attack and normal acceleration.

48. It should be carefully noted that the initial condition is, as before, steady level flight. The
difference between static and manoeuvre stability is that manoeuvre stability deals with a disturbance
in angle of attack (α) and load factor (n) occurring at constant speed, whereas static stability deals
with a disturbance in angle of attack at constant load factor (n = 1).

49. If an aircraft is trimmed to fly straight and


level (the initial condition of Fig 14-23), and is then
climbed, dived and pulled out of the dive so that at
the bottom of the pull-out it is at its original
trimmed values of speed and height (Fig 14-23),
then the aircraft can be considered as having been
"disturbed" from its initial condition in two ways,
both contributing to the overall manoeuvre
stability:

(a) It now has a greater angle of


attack to produce the extra lift required to
maintain a curved flight path (L = nW).
This is the same as the static stability
contribution discussed earlier.

(b) It has a nose-up rotation about its Fig 14-23: Forces Acting on an Aircraft in
a Level Flight and Steady Manoeuvre.
CG equal to the rate of rotation about its
centre of pull-out.

50. Because the aircraft is rotating about its


own CG, the tailplane can be considered to be
moving downwards relative to the air or,
alternatively, the air can be considered to be
moving upwards relative to the tailplane. In either
Fig 14-24: Increase in Tailplane AOA Due
case the effective angle of attack of the tailplane to its Vertical Velocity
will be increased (Fig 14-24), thus the manoeuvre
stability is greater than the static stability in level flight.
FIS Book 1: Aerodynamics 232

51. If the aircraft’s longitudinal stability is greater in manoeuvre, the position of the CG which
achieves neutral stability will be further aft than for the straight and level case. This position of the CG
is called the manoeuvre point (corresponding to the neutral point) and the distance between the CG
and the manoeuvre point is called the manoeuvre margin. It will be seen that for a given position of
the CG, the manoeuvre margin is greater than the CG margin.

52. Effect of Altitude. Consider an aircraft


flying at two different heights at the same EAS
and apply the same amount of up elevator in
each case. The elevator produces a downward
force on the tailplane to rotate the aircraft about
its centre of gravity and increase the angle of
attack of the wing. The rate of change of pitch
attitude is dependent upon the magnitude of the
Fig 14-25: Effect of Altitude on
force applied to the tailplane and also the TAS. Tailplane Contribution
As shown in Fig 14-25 the force is proportional to
EAS2 but its effect upon the change in angle of attack of the elevator is a function of the EAS2/ TAS
ratio. Fig 14-25 shows that the change in angle of attack at sea level is greater than at altitude and
therefore the elevator is less effective at altitude and stability is decreased.

LATERAL STABILITY

53. Having seen and studied in detail some of


the problems of longitudinal stability and also
directional stability, we shall now discuss the stability
of an aircraft in the lateral mode. Though there’s a
complex interaction between the lateral and
directional stability, for the sake of understanding
them better we have described them separately.

54. Damping in Roll. It is the rolling moment


associated not simply with the displacement in roll
but with the angular velocity in roll. When an aircraft
is disturbed in roll about its longitudinal axis say to
the port, the angle of attack of the down-going wing
is increased because of the downward velocity and
that on the up-going wing is decreased because of
the upward velocity (Fig 14-26). As long as the Fig 14-26: Damping in Roll
aircraft is not near the stall, the difference in angle of
233 Stability

attack produces an increase of lift on the down-going wing and a decrease on the up-going wing. The
rolling moment produced thus opposes the initial disturbance and results in a "damping in roll" effect.

55. Since the damping in roll effect is


proportional to the rate of roll of the aircraft, it
cannot bring the aircraft back to the wings-level
position. Therefore in the absence of any other
leveling force, an aircraft disturbed in roll would
remain with the wings banked. Therefore, by
virtue of the damping in roll effect, an aircraft
possesses neutral static stability with respect to an
angle of bank disturbance. However, when an
aircraft is disturbed laterally it experiences not only
a rolling motion but also a side slipping motion
caused by the inclination of the lift vector (Fig 14- Fig 14-27: Vector Addition of Forward &
Sideslip Velocities
27). Now if the various forces and moments
acting on the different
parts of the aircraft due
to side slip produce a
rolling moment tending to
restore the aircraft to its
initial wings-level position
the aircraft seems to
possess static lateral
stability.
Fig 14-28: Rolling Moment due to Sideslip
56. Thus the lateral
stability of an aircraft involves considerations of
rolling moment due to sideslip. The static lateral
stability of an aircraft can be studied with the help
of a graph of rolling moment co-efficient CL and
sideslip angle β (Fig 14-28 & 29). From the graph
it can be seen that lateral stability will exist if the
curve of CL versus β has a negative slope and the
degree of the stability will be function of this slope

57. In order to appreciate the development of


lateral stability in an aircraft contribution from
different parts of an aircraft will have to be studied Fig 14-29: Effect of Sideslip on Lateral
Stability
and these contributions will be of different
FIS Book 1: Aerodynamics 234

magnitude depending on the condition of flight and the particular configuration of the aircraft. The
more important of these contributions are from: -

(a) Dihedral.
(b) Sweepback, aspect ratio & taper ratio.
(c) Wing/fuselage interference.
(d) Position of keel surfaces.
(e) Undercarriage & flap.
(f) Slip stream effect.
(g) Power effects.

58. Dihedral Effect. The most common


method of obtaining lateral stability is by the
use of dihedral angle on the main planes. It is
the angle between each main plane and the
horizontal and not the angle formed between
the both wings. So if the wings are inclined
upward towards the tip, the dihedral is positive
and if downwards it is negative and called as
anhedral. From the Fig 14-30 it is clear that
due to the geometric dihedral, a point nearer
the wing tip (A or D) is higher than a point
inboard (B or C). Therefore a sideslip to
starboard will produce the following effects:

(a) Starboard Wing. The


relative airflow will cross the wing
(from A to B) at an angle equal to the
sideslip angle. Since point A is higher
than point B this will produce the
same effect as raising the leading Fig 14-30: Dihedral Effect
edge and lowering the trailing edge, i.e. increasing the angle of attack. As long as the aircraft
is not flying near the stalling speed the lift would increase.

(b) Port Wing. By a similar argument, the angle of attack on the port wing will reduce
and its lift would decrease.

A stable rolling moment is thus produced whenever sideslip is present. This is one of the most
important contributions to the overall stability and it is for this reason that the lateral static stability is
often referred to as the "dihedral effect" although there are a number of other important contributions.
235 Stability

59. Sweepback. Wing sweepback has the effect of producing an additional stabilizing
contribution thus increasing the "effective" dihedral of the wing (10° of sweep has about the same
effect as1° of dihedral). Fig 14-31 illustrates the principal effect of wing sweep back.

Fig 14-31: Effect of Sideslip on Swept Planform

(a) Angle of Sweep. The angle of sweep of the leading wing (wing into the side slip) is
decreased and that of the trailing wing is increased by the amount of sideslip angle. If the
wings are operating at a positive lift co-efficient, the leading wing generates more lift than the
trailing wing and thus a stable rolling moment is induced because of the sideslip (Fig 14-31).

(b) Aspect Ratio. On an aircraft with considerable sweep back placed in a sideslip, the
effective span of the leading (low) wing is increased and the chord decreased, which is an
effective increase in aspect ratio. On the trailing (high) wing, the span is decreased and the
chord is increased resulting in a reduction in aspect ratio. This again produces a stable rolling
moment because the more efficient (low) wing produces more lift.

(c) Taper Ratio. A small effect arises from a tapered wing. An increase in taper ratio,
defined as tip chord to root cord ratio, affects the lift coefficient and also produces a small
stable rolling moment in sideslip.

(d) Variation with Speed. The changes in the slope of the lift curve associated with
changes in aspect ratio and sweep result in variations in lift forces of the "leading" and
"trailing" wings. It follows, therefore, that the contribution of sweep to the lateral (static)
stability becomes more important at the higher values of CL, i.e. at the lower forward speeds,
FIS Book 1: Aerodynamics 236

because the CL curves are divergent. This is very important because it means that the
"dihedral effect" varies considerably over the
speed range of the aircraft. At high speeds a
lower angle of attack is needed than at low
speeds, therefore the stability at high speeds is
much less than at low speeds. To reduce the
stability to a more reasonable value at the higher
angles of attack, it may be necessary to
incorporate some negative dihedral (i.e.
anhedral) on a swept-wing aircraft, e.g. MiG 21 Fig 14-32: Anhedral of the MiG 21
Aircraft
(Fig 14-32).

60. Wing / Fuselage Interference

(a) Shielding Effect. Most aircraft will be


affected by the shielding effect of the fuselage.
In a sideslip, the section of the trailing (outer)
wing near the root lies in the "shadow" of the
fuselage (Fig 14-33). The dynamic pressure
over this part of the wing may be less than over
the rest of the wing and therefore produces less
lift. This effect will tend to increase the
"dihedral effect" and thus adds to lateral stability.

(b) Vertical Location. The vertical


location of the wings with respect to the fuselage Fig 14-33: Shielding Effect
contributes towards the lateral stability. To make
things simpler it would be helpful to start by
considering the fuselage to be cylindrical in
cross-section and placed in a sideslip. The
sideslip velocity will flow round the fuselage,
being deflected upwards across the top and
downwards underneath. Superimposing a wing
in this flow has the following effect, illustrated in
Fig 14-34.

(i) High Wing. A high-mounted Fig 14-34: Vertical Location of Wing


wing root will lie in a region of upwash
on the up-stream side of the fuselage tending to increase its overall angle of attack.
Conversely, on the down-stream side of the fuselage the wing root is influenced by
the downwash tending to reduce its angle of attack. The difference in lift produced by
237 Stability

each wing will cause a restoring moment bringing the aircraft back to wings level.
This effect has been demonstrated to be equivalent to 1° to 3° of dihedral.

(ii) Low Wing. The effect of locating the wing on the bottom of the fuselage is
to bring it into a region of downwash on the up-stream side and into upwash on the
down-stream side of the fuselage. The angle of attack of the leading (low) wing will
be decreased and that of the trailing wing increased. This gives rise to an unstable
moment equivalent to about 1° to 3° of anhedral.

From these facts it can be seen that there is zero effect on lateral stability when the wing is mounted
centrally on the fuselage. The effect is lessened as separation occurs at the wing/fuselage junction.

61. Position of vertical surfaces.


Since the aircraft is side slipping, there
will be a component of drag opposing
the sideslip velocity. If the drag line of
the aircraft is above the CG the result
will be a restoring moment tending to
raise the low wing. This configuration is
therefore a contribution towards positive
lateral stability. Conversely, a drag line
below the CG will be an unstable
contribution. The position of the drag
line is determined by the geometry of
the entire aircraft but the major
contributions, illustrated in Fig 14-35
are:

(a) High wing.


(b) High fin and rudder.
(c) Tee-tail configuration.
Fig 14-35: De-Stabilizing Effect of Slipstream &
Thus if the vertical surfaces / side Flaps

surfaces viz. vertical fin, T-Tail etc. are


above the CG then it would restore the aircraft back to wings level and if they are below the CG the
pressure on them would tend to make the ac roll even more and so cause lateral instability. Aircraft
such as IL-76 have high tail plane mounted on top of a high fin and such aircraft have anhedral on the
main plane to counterbalance this effect and prevent too great a degree of lateral stability (Fig 14-35).
FIS Book 1: Aerodynamics 238

62. Slipstream. The effect of slipstream in


case of a propeller aircraft is destabilizing as
illustrated in Fig 14-36. Due to sideslip the
slipstream behind the propeller or propellers is no
longer symmetrical about the longitudinal axis.
The dynamic pressure in the slipstream is higher
than the free stream and covers more of the
trailing (outer) wing in sideslip. The result is an
unstable moment tending to increase the
displacement (wing low). This unstable
contribution is worse with flaps down. Fig 14-36: De-Stabilizing Effect of
Slipstream

63. Flaps. Partial-span flaps alter the span wise distribution of pressure across a wing. The
local increase in lift coefficient near the root (Fig 14-37) has the effect of moving the "half-span" centre
of pressure towards the fuselage (in a span wise sense). If the local increase in lift co-efficient
reduces the moment arm of the wing, then a given change in CL due to the dihedral effect will produce
a smaller restoring moment and
thus become destabilizing. The
design geometry of the flap itself
can be used to control this
contribution. In particular, a
swept-back flap hinge-line will
increase the dihedral effect,
whereas a swept-forward hinge- Fig 14-37: De-Stabilizing Effect of Flaps
line will decrease it.

64. Handling Considerations (lateral stability). It has been shown that the "dihedral effect" of
sweepback in sideslip produces a strong rolling moment. This has been referred to somewhat
imprecisely as roll with yaw. Two applications of this effect at low speeds, where it is strongest, are
worth considering:

(a) Cross-Wind Landing. After an approach with the aircraft heading into a cross-wind
from the right, the pilot must yaw the aircraft to left to align it with the runway prior to
touchdown. This action will induce a sideslip to right and the pilot must anticipate the
subsequent roll to left, in order to keep the wings level.

(b) Wing Drop. The greater tendency of a swept-wing aircraft to drop a wing at a high
angle of attack may be further increased by a large deflection of corrective aileron. In such
cases the dihedral effect of sweepback may be utilized by applying rudder to yaw the nose
towards the high wing - sideslip to the left, roll to the right. It must be said, however, that
modern design has reduced the tip-stalling tendency and improved the effectiveness of
239 Stability

ailerons at high angles of attack and the problem is not as acute as it might have been in the
"transonic era".

DYNAMIC STABILITY

General

65. When an aircraft is disturbed from the equilibrium state, the resulting motion and
corresponding changes in the aerodynamic forces and moments acting on the aircraft may be quite
complicated. This is especially true for displacement in yaw which affects the aircraft in both yawing
and rolling planes.

66. Some of the factors affecting the long-term response of the aircraft are listed below:

(a) Linear velocity and mass (momentum).


(b) The static stabilities in roll, pitch and yaw.
(c) Angular velocities about the three axes. Angular
(d) Moments of inertia about three axes. Momentum
(e) Aerodynamic damping moments due to roll.

67. Consider a body which has been disturbed from its equilibrium state and the source of the
disturbance then removed. If the subsequent system of forces and moments tends initially to
decrease the displacement, then that body is said to have positive static stability. It may, however,
overshoot the equilibrium condition and then oscillate about it. The terms for possible forms of motion
which describe the dynamic stability of the body are listed below:

(a) Amplitude increased - negative stability.


(b) Amplitude constant - neutral stability.
(c) Amplitude "damped" - positive stability.
(d) Motion heavily damped - oscillations cease and the motion becomes "dead-beat"
positive stability.
(e) Motion diverges - negative
dynamic stability.

Fig 14-38 illustrates these various forms of


dynamic stability; in each case shown, the
body has positive static stability.

68. Dynamic stability is more readily


understood by use of the analogy of the "bowl
and ball" described and illustrated at para 3. Fig 14-38: Forms of Motion
FIS Book 1: Aerodynamics 240

For example, when the disturbance is removed from the ball, it returns to the bottom of the bowl and
is said to have static stability. However, the ball will oscillate about a neutral or equilibrium position
and this motion is equivalent to dynamic stability in an aircraft.

69. If the oscillations are constant in amplitude


and time, then a graph of the motion would be as
shown in Fig 14-39. The amplitude shows the extent
of the motion, and the periodic time is the time taken
for one complete oscillation. This type of motion is
known as simple harmonic motion.
Fig 14-39: Simple Harmonic Motion

70. Periodic Time. The time taken for one complete oscillation will depend upon the degree of
static stability, i.e. the stronger the static stability, the shorter the periodic time.

71. Damping. In this simple analogy it is assumed that there is no damping in the system and
the oscillations will continue indefinitely and at constant amplitude. In practice there will always be
some damping. The viscosity of the fluid (air) is a damping factor which is proportional to the speed of
the mass. Damping can be expressed as the time required (or number of cycles) for the amplitude to
decay to one half of its initial value (Fig 14-38 - Damped Phugoid). An increase in the damping of the
system (e.g. a more viscous fluid) will cause the oscillations to die away more rapidly and, eventually,
a value of damping will be reached for which no oscillations will occur. In this case, after the
disturbance has been removed, the mass returns slowly towards the equilibrium state but does not
overshoot it, i.e. the motion is "dead-beat" (Fig 14-38 - Positive Dynamic Stability).

72. Dynamic Stability of Aircraft. The dynamic stability of an aircraft depends on the
particular design of the aircraft and the speed and height at which it is flying. It is usually assumed
that for "conventional" aircraft the coupling between the longitudinal (pitching) and lateral (including
directional) motions of an aircraft can be neglected. This enables the longitudinal and lateral dynamic
stability to be considered separately.

73. Design Specification. Oscillatory motions which have a long periodic time are not usually
important. Even if the motion is not naturally well damped, the pilot can control the aircraft fairly easily.
To ensure satisfactory handling characteristics, however, it is essential that all oscillatory motions with
a periodic time of the same order as the pilot's response time are heavily damped. This is because
the pilot may get out of phase with the motion and pilot-induced oscillations (PIOs) may develop. The
minimum damping required is that oscillation should decay to one half of their original amplitude in
one complete cycle of the motion. Some modern aircraft, however, do not satisfy this requirement
and in many cases it has been necessary to incorporate autostabilization systems to improve the
basic stability of the aircraft, e.g. pitch dampers or yaw dampers.
241 Stability

Longitudinal Dynamic Stability

74. When an aircraft is disturbed in pitch from trimmed level flight it usually oscillates about the
original state with variations in the values of speed, height and indicated load factor. If the aircraft has
positive dynamic stability, these oscillations will gradually die away and the aircraft returns to its initial
trimmed flight condition. The oscillatory motion of the aircraft in pitch can be shown to consist of two
separate oscillations of widely differing characteristics, the phugoid and the short-period oscillation.

75. Phugoid. This is usually a long period, poorly damped motion involving large variations in
the speed and height of an aircraft but with negligible changes in load factor (n). It can be regarded
as a constant energy motion in which potential energy and kinetic energy are continuously
interchanged. The phugoid oscillation is usually damped and the degree of damping depends on the
drag characteristics of the aircraft. The modern trend towards low drag design has resulted in the
phugoid oscillation becoming more of a problem.

Fig 14-40: Basic Components of Longitudinal Fig 14-41: Short Period Oscillation
Dynamic stability

76. Short-Period Oscillation. This oscillatory motion is usually heavily damped and involves
large changes of load factor with only small changes in speed and height. It can be regarded simply
as a pitching oscillation with one degree of freedom. In para 70 it was stated that the time taken for
one complete oscillation will depend upon the static stability, in this case it is the periodic time of the
short-period oscillation (Fig 14-41).

77. Stability Factors. The longitudinal dynamic stability of an aircraft, i.e. the manner in which
it returns to a condition of equilibrium, will depend upon:

(a) Static longitudinal stability.


(b) Aerodynamic pitch damping.
(c) Moments of inertia in pitch.
(d) Angle of pitch.
(e) Rate of pitch.
FIS Book 1: Aerodynamics 242

78. Having studied some aspects of dynamic stability and the lateral and directional stability of an
aircraft in isolation, we are now in a position to connect these two forms of stability. Both lateral as
well as directional stability are so interlinked that in flight both lateral and directional responses are
coupled. When an aircraft in free flight is placed in a sideslip the aeroplane produces both rolling
moment as well as yawing moment due to sideslip. Thus, the lateral dynamic motion of an aircraft in
free flight must consider the coupling or interaction of the lateral and directional effects. The principal
effects which determine the lateral and dynamic characteristics of an aircraft are:-

(a) Rolling moment due to sideslip or dihedral effect (lateral stability).


(b) Yawing moments due to sideslip or static directional stability.
(c) Yawing moment due to rolling velocity or adverse (or proverse) yaw.
(d) Rolling moment due to yawing velocity.
(e) Aerodynamic side force due to sideslip.
(f) Rolling moment due to rolling velocity (damping in roll).
(g) Yawing moment due to yawing velocity (damping in yaw).
(h) The moment of inertia of the aircraft about the roll axis.

79. The complex interaction of these factors produces three types of motion of the aircraft:

(a) Directional divergence. Directional divergence is a condition which cannot be


tolerated. If the reaction to a small initial side slip is such as to create moments which tend to
increase the sideslip, directional divergence will exists. The side slip would increase until the
aircraft is broadside to the wind or structural failure occurs. Increasing the static directional
stability reduces the tendency for directional divergence.

(b) Spiral instability. Spiral instability will exist when the static directional stability of
an aircraft is large as compared to lateral stability. The lateral stability of an aircraft depends
on the forces that tend to right the aircraft when a wing drops. However, at the same time the
keel surface (including the fin) tends to yaw the aircraft into the airflow, in the direction of the
lower wing. Once the yaw is started, the higher wing, being on the outside of the turn and
travelling slightly faster than the lower, produces more lift. A rolling moment is then set up
which opposes, and may be greater than, the correcting moment of the dihedral, since the roll
due to yaw will tend to increase the angle of bank. If the total rolling moment is strong enough
to overcome the restoring force produced by dihedral and the damping in yaw effect, the
angle of bank will increase and the aircraft will enter a diving turn of steadily increasing
steepness. This is known as spiral instability. A reduction in fin area, reducing directional
stability and the tendency to yaw into the sideslip results in a smaller gain in lift from the
raised wing and therefore in greater spiral stability. The character of spiral divergence is by no
means violent and is easily controllable by the pilot.
243 Stability

(c) Dutch roll. Dutch Roll or the oscillatory instability is more serious than spiral
instability and is commonly found to a varying degree in combinations of high wing loading,
sweepback (particularly at low IAS) and high altitude. Dutch roll is characterized by a
combined rolling and yawing movement or "wallowing" motion. When an aircraft is disturbed
laterally, the subsequent motion may be either of the two extremes. The aerodynamic causes
of oscillatory instability are complicated and a simplified explanation of one form of Dutch roll
is given in the following paragraph. Consider a swept-wing aircraft seen in planform. If the
aircraft is yawed, say to starboard, the port wing generates more lift due to the larger expanse
of wing presented to the airflow and the aircraft accordingly rolls in the direction of yaw.
However, in this case the advancing port wing also has more drag because of the larger area
exposed to the airflow. The higher drag on the port wing causes a yaw to port which results
in the starboard wing obtaining more lift and reversing the direction of the roll. The final result
is an undulating motion in the directional and lateral planes and results in Dutch roll. Since
the motion is caused by an excessive restoring force one method of tempering its effect is to
reduce the lateral stability by setting the wings at a slight anhedral angle.

SUMMARY

Static and Dynamic Stability of Aircraft

80. Stability is concerned with the motion of a body after an external force has been removed.
Static stability describes its immediate reaction while dynamic stability describes the subsequent
reaction.

81. Stability may be of the following types:

(a) Positive - the body returns to the position it held prior to the disturbance.
(b) Neutral - the body takes up a new position of constant relationship to the original.
(c) Negative - the body continues to diverge from the original position.

82. The factors affecting static directional stability are:

(a) Design of the vertical stabilizer.


(b) The moment arm.

83. The factors affecting static longitudinal stability are:

(a) Design of the tailplane.


(i) Tail area.
(ii) Tail volume.
(iii) Planform.
FIS Book 1: Aerodynamics 244

(iv) Wing downwash.


(v) Distance from CP tail to CG.
(b) Position of CG.
(i) Aft movement of the CG decreases the positive stability.
(ii) Forward movement of the CG increases the positive stability.

84. Manoeuvre stability is greater than the static stability in level flight and a greater elevator
deflection is necessary to hold the aircraft in a steady pull-out.

85. The factors affecting static lateral stability are:

(a) Wing contributions due to:


(i) Dihedral.
(ii) Sweepback.
(b) Wing/fuselage interference.
(c) Fuselage and fin contribution.
(d) Undercarriage, flap and power effects.

86. Some of the factors affecting the long-term response of the aircraft are:

(a) Linear velocity and mass.


(b) The static stabilities in roll,pitch and yaw.
(c) Angular velocities about the three axes
(d) Moments of inertia about the three axes.
(e) Aerodynamic damping moments due to roll, pitch and yaw.

87. The longitudinal dynamic stability of an aircraft depends upon:

(a) Static longitudinal stability.


(b) Aerodynamic pitch damping.
(c) Moments of inertia in pitch.
(d) Angle of pitch.
(e) Rate of pitch.
245

CHAPTER 15

LEVEL FLIGHT

Introduction

1. For any aircraft, there are many occasions when it is in unaccelerated, steady straight and
level flight. Therefore the forces that act on the aircraft during this phase of straight and level flight
need to be understood to appreciate the balance of these forces as well as the factors that govern
certain design features to be used on aircraft.

2. During flight four main forces act upon an aircraft, viz. lift, weight, thrust and drag. When an
aircraft is in steady level flight a condition of equilibrium must prevail. The unaccelerated condition of
flight is achieved with the aircraft trimmed for lift to equal weight and the engine set for thrust to equal
the aircraft drag.

3. The four forces mentioned in para 1 each have their own point of action viz. lift through the
centre of pressure (CP), weight through the centre of gravity (CG), and for the purpose of this chapter
(although not strictly valid), thrust and drag act in opposite directions parallel to the direction of flight
through points varying with aircraft design and attitude. Although the opposing forces are equal there
is a considerable difference between each pair of forces. The lift or weight figure, for example, may be
25,000 kg while the drag or thrust maybe as low as 2,500 kg, depending on the speed and power
used.

Pitching Moments

4. The positions of the CP and CG


are variable and under most conditions
of level flight they are not coincident.
The CP changes its position with
changes in angle of attack and the CG
changes its position with reduction in
fuel or when stores are expended. The
outcome is that the opposing forces of
lift and weight set up a couple causing
either a nose-up or a nose-down
pitching moment depending on whether
the lift is in front of or behind the CG, as Fig 15-1: Pitching Moment
illustrated in Fig 15-1. The same
consideration applies to the position of the lines of action of the thrust and drag. Ideally the pitching
FIS Book 1: Aerodynamics 246

moments arising from these two couples should neutralize each other in level flight so that there is no
residual moment tending to rotate the aircraft.

5. This ideal is not easy to attain by juggling with the lines of action of the forces alone but, as
far as possible, the forces are arranged as shown in Fig 15-2. With this arrangement the thrust/drag
couple produces a nose up moment and the
lift/weight couple a nose-down moment, the
lines of action of each couple being positioned
so that the strength of each couple is equal. If
the engine is throttled back in level flight the
thrust/drag couple is weakened and the
lift/weight couple pitches the nose down so
that the aircraft assumes a gliding attitude. If
the power is then reapplied the growing
strength of the thrust/ drag couple raises the
nose towards the level flight attitude. Fig 15-2: Disposition of four forces

6. Tailplane and Elevator. The


function of a tailplane is to supply any force
necessary to counter residual pitching
moment arising from inequalities in the two
main couples, i.e., it has a stabilizing function.
This is achieved most commonly by fixing a
hinged flap behind the tailplane by which the
direction of force being generated, as well as
Fig 15-3: Action of Tailplane
its magnitude can be varied. The force on the
tailplane is positioned some distance from the CG means that it can apply a large moment to the
aircraft (Fig 15-3). For this reason the area and lift of the tailplane is small compared with the
mainplane.

7. Additional Information on the Tailplane

(a) Design. The tailplane may be required to produce either an upload or a download.
Therefore the tailplane is usually symmetrical in design.

(b) Tailplane Flight Longitudinal Dihedral. In most conventional aircraft, the tailplane
operates in the downwash of the mainplane. The downwash reduces the effective angle of
attack of the tailplane (usually the tailplane operates at half the angle of attack of the main
plane). The reduced angle of attack of the tailplane is termed as the Tailplane Flight
Longitudinal Dihedral.
247 Level Flight

(c) Tailplane Riggers Dihedral. In order to offset the reduction in angle of attack due
to mainplane downwash, the angle of incidence of the tailplane is usually rigged to some
positive incidence relative to the main plane. This is referred to as Tailplane Riggers Dihedral.

(d) Trim Drag. In order to maintain straight and level flight, the tailplane is required to
produce an up/down force to counter any imbalance between the thrust - drag and lift -
weight couples. If the tailplane has to provide a downward force, this additional force adds to
the effective weight of the aircraft. The increase in weight has to be countered by increasing
lift (by increasing speed or angle of attack). The additional drag generated in this process is
called trim drag. It is of great consequence for high-speed flight.

(e) Case of Reducing Speed and Stabilized Low speed. When speed is being
reduced, the stick is being brought back, i.e., the elevator is deflected upwards, implying that
the tailplane is generating downward force (lift). However, this is a transient phase. In a
stabilized low speed condition of level flight, the center of pressure has moved up (as the
angle of attack is higher) and the nose down lift-weight couple is reduced. This would cause
the nose down pitch to reduce. Now, to maintain straight and level flight, the force being
produced by the tailplane is less downward. The converse occurs in a case of increasing
speed and stabilized High Speed.

8. Variation of Speed in Level Flight. For level flight the lift must equal the weight. From the
lift formula (L= CL ½ ρ V2 S) it can be seen that, for a given aircraft flying at a stated weight, if the
speed factor is decreased, then the lift coefficient (angle of attack) must be increased to keep the lift
at the same value as the weight.

9. Aircraft Attitude in Level Flight. At low speed the angle of attack must be high, while at
high speed only a small angle of attack is needed to obtain the necessary amount of lift. Since level
flight is being considered, these angles become evident to the pilot as an attitude, which will be nose-
up at low speeds and nose lower at high speeds. The difference between low and high-speed
attitudes is most marked on aircraft having sweptback or unswept wings of low aspect ratio, for the
reasons given in earlier chapters.

10. Aircraft Performance in Level Flight. The variation of power required and thrust required
with velocity is illustrated in Fig 15-4. Each specific curve of power or thrust required is valid for a
particular aerodynamic configuration, at a given weight and altitude. These curves define the power or
thrust required to achieve equilibrium, i.e., lift equal to weight or constant altitude flight, at various air
speeds. As shown by the curves of Fig 15-4, if it is desired to operate the aircraft at the air speeds
corresponding to point A, the power or thrust required curves define a particular value of thrust or
power that must be made available from the engine to achieve equilibrium. Some different air speed,
such as that corresponding to point B, changes the value of thrust or power required to achieve
FIS Book 1: Aerodynamics 248

Fig 15-4: Level Flight Performance

equilibrium. Of course, the change of airspeed to point B would also require a change in angle of
attack to maintain a constant lift equal to the aircraft weight. Similarly, to establish airspeed and
achieve equilibrium at point C will require a particular angle of attack and engine thrust or power. In
this case, flight at point C would be in the vicinity of the minimum flight speed and a major portion of
the thrust or power required would be to counter induced drag. The maximum level flight speed for the
aircraft will be obtained when the power required equals the maximum power or thrust available from
the engine. The minimum flight airspeed is not usually defined by thrust or power requirements since
conditions of stall or stability and control problems generally predominate.

Effect of Weight on Level Flight

11. If an aircraft is flying level at a given angle of attack, for example that for best L/D ratio (about
0
4 ) and at a known weight and EAS, then if the weight is reduced by dropping stores the lift must also
be reduced to balance the new weight. To maintain optimum conditions with the angle attack at 40,
the speed must be decreased until the lift reduces to the same value as the new weight.

12. The lower the weight, the lower is the EAS corresponding to a given angle of attack, the EAS
at the required angle being proportional to the square root of the all-up weight.

13. Effect of Altitude on Level Flight. Since not only the speed but also the lift and drag vary
2
with the ½ ρ V S factor, the relationship between EAS and angle of attack is unchanged at altitude,
provided of course that the weight is constant. However compressibility effects tend to alter the
relationship by reducing the CL, appropriate to a given angle of attack.
249

CHAPTER 16

CLIMBING AND GLIDING

General

1. This chapter deals primarily with the principles of flight involved in both climbing and gliding,
but in considering the factors affecting the climb, some performance details are discussed.

CLIMBING

2. During a climb an aircraft gains potential energy by virtue of elevation. This is achieved by
one or both of the following two means:

(a) The expenditure of propulsive energy over and above that required to maintain level
flight.
(b) The expenditure of aircraft kinetic energy, i.e. loss of velocity by a zoom.

Zooming for altitude is a transient process of exchanging kinetic energy for potential energy and is of
considerable importance for aircraft which can operate at very high levels of kinetic energy. However,
the major portion of climb performance for most aircraft is a near steady process in which additional
propulsive energy is converted into potential energy.

Forces in the Climb

3. To maintain a climb at a given EAS, more power has to be provided than in level flight. This
is, firstly, to overcome the drag as in level flight (PREQ = D X V), secondly, to lift the weight at a finite
vertical speed, which is known as the rate of climb (W X Vc), and thirdly to accelerate the aircraft
slowly as the TAS steadily increases with increasing altitude for the same EAS ( W X a X v).
g

PREQ = D.V + W.Vc + W.V a


g
Where Vc is the rate of climb
a is the acceleration.

4. The acceleration term can be ignored in low performance aircraft but has to be taken into
account in jet aircraft with high rates of climb. It can be seen from Fig 16-1 that in the climb, L =
W Cos θ i.e., lift is less than weight, as a component of weight (W sin 2) is being balanced by Thrust.
If this is so then the lift dependent drag is less than that at the same speed in level flight. However, it
is still considered sufficiently accurate to assume that lift equals weight up to about 15° climb angle
(since Cos 15° is 0.9659, the error due to this assumption is less than 3%).
FIS Book 1: Aerodynamics 250

Maximum Rate of Climb

5. Fig 16-1 (a) & (b) can be used to show that rate of climb is determined by the amount of
excess power and the angle of climb is determined by the amount of excess thrust left after opposing
drag. The expression for maximum rate of climb can be found as follows:

16-1 a & b: Forces in Climb

From Fig 16-1(a): Sin θ = Rate of Climb where V is TAS (16.1)


V
From Fig 16-1(b): Thrust = D + W Sin θ
∴ Sin θ = Thrust – Drag (16.2)
Weight

Equating right hand sides of equation (16.1) and (16.2), we get:

Rate of Climb = Thrust – Drag


V Weight
∴ Rate of Climb = V (Thrust – Drag)
Weight
Rate of Climb = PAV - PREQ (16.3)
Weight
Rate of Climb = Excess power
Weight

6. In practice, aircraft do not, for varying reasons (e.g. engine cooling, avoidance of an
exaggerated attitude etc.) always use the exact speed for maximum rate of climb. In jet aircraft this
speed is quite high and at low altitude is not very critical due to the shape of the power available curve
(Fig 16-3). In piston aircraft the speed is much lower and is normally found to be in the vicinity of the
minimum drag speed.

Maximum Angle of Climb

7. When the maximum angle of climb is required it can be seen from Fig 16-1 that:

Sin θ = Thrust – Drag


Weight
251 Climbing and Gliding

The aircraft therefore should be flown at


the speed which gives the maximum
difference between thrust and drag. For
piston aircraft, where thrust is reducing
as speed is increased beyond unstick,
the best speed is usually as low as is
safe above unstick speed. For a jet
aircraft, since thrust varies little with
speed, the best speed is at minimum
drag speed (Fig 16-2).

Power Available and Power Required

8. The thrust horsepower


available curve is calculated by
multiplying the thrust by the
corresponding speed. In SI, the value of
power in Watts could be obtained by
simply multiplying thrust in Newtons
with speed in m/s. The power available
curve for a jet engine differs from that of
a piston engine as shown in the upper
curves of Fig 16-3. The main reason for Fig 16-2: Maximum Angle of Climb,
Piston and Jet Aircraft
this difference is that the thrust of a jet
engine remains virtually constant at a given altitude, irrespective of the speed, therefore when this
constant thrust is multiplied by the appropriate air speed to calculate thrust horsepower, the result is a
straight line. The piston engine, on the other hand, under the same set of circumstances and for a
given bhp, suffers a loss of thrust horsepower at both ends of its speed range because of reduced
propeller efficiency. The thrust horsepower available curves are representative of a more powerful,
Second World War class piston engine fighter and a typical jet fighter having a high subsonic
performance.

9. The power required to propel an aircraft in level flight can be found by multiplying the drag by
the corresponding TAS. The lower curve of Fig 16-3 is a typical example and can be assumed to
apply to both a piston or jet-propelled airframe, i.e. the airframe drag is the same irrespective of the
power unit used. Note that the speed for minimum drag, although low, is not the lowest possible nor
is it that for minimum power. The increase in power required at the lowest speed is caused by the
rapidly rising effects of induced drag.
FIS Book 1: Aerodynamics 252

Climbing Performance

10. The vertical distance between


the power available and the power
required curves represents the power
available for climbing at a particular
speed. The best climbing speed
(highest rate of climb) is that at which
the excess power is at a maximum, so
that, after expending some power in
overcoming the drag, the maximum
amount of power remains available for
climbing the aircraft. From Fig 16-3, for
the piston engine aircraft the best speed
is seen to be about 300 kmph and for
the jets about 750 kmph. Notice that in
the latter case a fairly wide band of
speeds would still give the same
Fig 16-3: Typical Power Available and
amount of excess power for the climb Power Required Curves
but in practice the highest speed is
used since better engine efficiency is obtained. At the intersection of the curves (points X and Y) all
the available power is being used to overcome drag and none is available for climbing. These points
therefore represent the minimum and maximum speeds possible for the particular power setting.

11. If power is reduced the power available curve is lowered. Consequently the maximum speed
and maximum rate of climb are reduced, while the minimum speed is increased. When the power is
reduced to the point when the power available curve is tangential to the power required curve, the
points X and Y coincide and the aircraft cannot climb.

Effect of Altitude on Climbing

12. The thrust horsepower of


both jet and piston engines decreases
with altitude. Even if it is possible to
prolong sea-level power of the piston
engined aircraft to some greater
altitude by supercharging or some
other method of power boosting, the
power will inevitably decline when the
Fig 16-4: Effect of Altitude on Typical Power Available
boosting method employed reaches a and Power Required Curve – Piston
253 Climbing and Gliding

height at which it can no longer


maintain the set power.

13. At the highest altitudes the


power available curves of both types
of engines are lowered as shown in
Fig 16-4 & 5 and the power required
curve is displaced upwards and to the
right. Notice that the power required
to fly at the minimum drag speed is
increased. This effect is caused by
the fact that although the minimum Fig 16-5: Effect of Altitude on Typical Power Available
and Power Required Curves - Jets
drag speed, in terms of EAS, remains
same at all heights, the speed used in the calculation of THP is TAS, which increases with altitude for
a given EAS. Therefore the power required to fly at any desired EAS increases with altitude. From
Fig 16-4 & 5 it may be seen that the speed for the best rate of climb reduces with altitude and the
range of speeds between maximum and minimum level flight speeds is also reduced.

Note: The height of 40,000 ft used in Fig 16-5 was chosen because at that height EAS = ½ TAS.
A piston engine would not normally operate at that height.

14. Ceilings. The altitude at which the maximum power available curve only just touches the
power required curve and a sustained rate of climb is no longer possible is known as the Absolute
Ceiling. It is possible to exceed this altitude by the zoom climb technique, which converts the
aircraft’s kinetic energy (speed) to potential energy (altitude). Another ceiling is the Service Ceiling,
which is defined as the altitude at which the maximum sustained rate of climb falls to 2.5 m/sec (500
fpm) for jets and 0.5 m / sec (100 fpm) for a piston aircraft.

15. Operating Data Manuals. Since climb performance forms an important part of flight and
fuel planning of modern aircraft, especially jets, take-off and climb data with respect to distance
covered, fuel consumed and time taken are covered in the operating data manuals in detailed tabular
form catering for various variables like temperature, atmospheric pressure, weight of the aircraft,
winds and even runway slope.

GLIDING

General

16. Gliding is the state of flight when the aircraft flies without any thrust being provided by the
engines of the aircraft. Such a flight also is in a state of equilibrium. The aircraft loses height steadily
to maintain the desired speed for flying. The only source of energy in such a state can be the
FIS Book 1: Aerodynamics 254

potential energy of the aircraft, which exists by virtue of its weight and height. Before we see the
balance of forces, let us consider the following aspects of glide. Had the case been of free fall, the
aircraft would have come down vertically, accelerating downwards. Whereas in the glide, the aircraft
is not falling vertically downwards and instead it has a steady forward and downward components of
velocities. Both these components have to be provided by the force of weight only since that is the
only force at the disposal of the gliding aircraft.

Angle / Rate of Descent in Glide

17. As the aircraft descends on a steady glide path, it forms a finite angle with respect to the
horizontal. This angle is called the Gliding Angle. While gliding at a steady speed, the aircraft would
also have a finite rate of descent, which will be proportional to its gliding angle and the speed of
gliding.

18. An aircraft in glide may be flown at any speed by choosing the angle of descent of the aircraft,
i.e. by steepening the descent. The aircraft may be flown at a higher speed with a correspondingly
higher rate of descent or it may be possible to fly the aircraft at speeds as low as its stalling speed.
However, it should be noted that the relationship between speed and rate of descent is not direct
throughout. From the pilot’s point of view, lowering the nose of the glider may not necessarily mean
that the rate of descent would increase and the range of glide reduce. Similarly, slowest speed during
glide neither means the shallowest angle of descent nor does it mean the slowest rate of descent. To
understand the mechanics of a gliding descent let us understand the forces that act in a glide as
indicated in Fig 16-6.

Forces in a Glide

19. The four forces in


flight are Lift, Weight, Drag
and Thrust. Since there is
obviously no thrust
available in a glide, the
balance of forces of the
remaining three forces is as
Fig 16-6 a & b: Forces in a Glide
follows:

(a) Weight is balancing the lift as well as drag, however the addition is not ‘algebric’ but
‘vectoral’. From the Fig 16-6, it can be seen that the total (aerodynamic) reaction obviously is
balanced by the only available force, that is ‘Weight’.
(b) The drag accordingly, is being balanced by a component of weight i.e. W sin θ.
(c) The lift is being balanced by another component of weight i.e. W cos θ.
255 Climbing and Gliding

The Lift continues to be at right angles to the flight path and Drag continues to be along the flight path
in the reverse direction of flight. The only source of energy to do the work, which is being done
against the drag, is the potential energy of the aircraft, which is being slowly traded off as the height of
the aircraft is reducing.

Gliding for Endurance

20. Gliding for ‘endurance’ means gliding for maximum time in air. Obviously, from a specified
commencement height, an aircraft with the least rate of descent would stay afloat in air longest. A
commercial glider pilot who charges people for the glider rides proportional to the time spent in the air
would like to find such a speed for his glider so that he can make maximum money out of every
launch. Let us see what he would need to know.

21. In Fig 16-6 we can see that for minimum rate of descent, V sin θ must be minimum. From
balance of forces in a glide:

D = W sin θ

Multiplying the equation with velocity V

DV = W sin θ.V OR
Vsin θ = D.V (16.4)
W

From Fig 16-6, we know that

ROD = V sin θ

Therefore, from equation (16.4), for the ROD to be minimum, the value of D.V/W should be minimum.
Assuming the weight to be constant, the minimum ROD and hence maximum endurance of the
aircraft would be achieved at the speed requiring minimum power.

22. The commercial glider pilot, when flying at the endurance speed for his glider would find his
Vertical Speed Indicator showing the least ROD and would find his ROD stabilizing at a higher value
whether he reduces speed by raising the nose of the glider or increases speed by lowering the nose.

Gliding for Range

23. Gliding for range means gliding in a fashion so as to cover maximum distance by the time the
aircraft touches down on ground. From Fig 16-6, for a given commencement height, the maximum
range would mean a point farthest on the ground from the point vertically below the commencement
point. This in turn would mean the smallest gliding angle or the ‘flattest’ glide. For example, Kiran
FIS Book 1: Aerodynamics 256

aircraft has approximately an angle of glide of 4.7 degrees and the HPT-32 aircraft has an angle of
glide of about 12.95 degrees when flying at respective best L/D angles of attack.

24. From Fig 16-6, the triangle formed by Lift, Drag and Total reaction is geometrically similar to
that formed by Distance, Height and Glide-path. Now, if distance is to be maximum, gliding angle
must be minimum. θ is minimum when Cot θ is maximum or Tan θ is minimum. From the balance of
forces, since L = W . Cos θ and D = W . Sin θ, ∴ Cot θ = L / W. Therefore θ is maximum when
L / D is maximum. Further, in most aircraft the gliding angle is quite shallow, being less than 15
degrees hence the lift can be safely assumed to be equal to weight. For the more mathematically
inclined reader, truly speaking the lift will be equal to Cos function of the gliding angle, which in case
of a glide angle of 15 deg works out to be L X Cos 15 = L X 0.9659 i.e. less than 2% error, which will
be even lesser at lesser glide angles.

25. Therefore best (smallest) angle of glide depends on maintaining an angle of attack that gives
the best lift/drag ratio. Since lift is assumed to be equal to the weight and hence constant, best L/D
ratio would correspond to the ‘minimum drag’ condition. For the mathematician, like mentioned earlier,
in case of steeper gliding angles, the lift would be lesser than the weight by a factor of cos θ. In case
the aircraft is equipped with an angle of attack indicator, the pilot needs to just maintain the same
angle of attack as earlier; however if the aircraft does not have the AOA indicator and the pilot has to
go by the EAS, he would have to reduce gliding speed by a factor of √cos θ to continue to maintain
the optimum angle of attack for best range.

26. Effect of Altitude on Range, Speed and ROD. From the explainations above it can be
seen that the best glide situation would continue to be achieved irrespective of altitude, at the optimal
AOA for best L/D ratio. Even if the pilot does not have an AOA indicator, since ignoring
compressibility, drag depends on EAS, the best gliding speed at a given weight would be at a
constant EAS regardless of altitude. Consequently, since the ratio of Lift / Drag would remain
unchanged, the angle of glide and hence the glide range for a given loss of height would remain same
throughout, irrespective of the altitude. The rate of descent will, however, reduce at lower altitudes
because the TAS for the same EAS reduces at lower altitudes and therefore the rate at which the
aircraft ‘physically’ loses height reduces as the TAS reduces. Mathematically, since V sin θ means the
rate of descent, where V is the TAS, as V reduces the ROD reduces at lower altitudes.

Effect of Wind on Range and Endurance (At Constant Weight)

27. Endurance. Head or tail wind does not have any effect on the rate of descent since the
ROD depends only on the product of V (TAS) and sin θ (glide angle). Here, one has to understand
that the glide angle in this case is not to be measured with respect to the ground but with respect to
the air mass in which the aircraft is gliding, because the factor ‘V’ is also being measured with respect
257 Climbing and Gliding

to the air mass and not the ground. Therefore, in a glide for best endurance, since the rate of descent
is not being affected by the wind there will be no change in the net endurance achieved.

28. Range. When gliding for range the all-important factor is the point of arrival on the ground.
In still air this is achieved by flying for minimum drag as explained earlier. It is easy to comprehend
that in a tail wind condition the net range will increase because the air mass in which the aircraft is
gliding would itself be travelling in the favourable direction giving the aircraft a higher ground speed
and hence longer range for the same ROD. Similarly, the effect of a head wind will be to decrease
the ground distance travelled because of reduced ground speed. The extent of this increment or
reduction in range would be proportional approximately to the ratio of the wind speed compared to the
TAS. But question arises that in the ‘windy’ situation, would the best L/D speed continue to give best
range or is it possible to extract more out of a favourable wind or reduce the ill effect of an
unfavourable wind? Some thought would reveal that it may be better to stay in air for longer by
reducing the ROD in a tail-wind situation thereby making use of the ‘push’ from the favourable tail
wind. However, this reduction in ROD will have to be achieved at a different (and lower) L/D ratio.
Therefore, care has to be taken that the advantage accrued by staying afloat longer is not undone by
the loss due to the reduced L/D ratio.

29. Similarly, in a head-wind situation, it may be better to reduce the exposure to the adverse
wind by increasing the rate of descent. Once again, the ensuing increment of speed at the new (and
lower) L/D should not offset the gain due to reduced exposure to adverse wind.

30. It is possible to find out the theoretically best gliding speeds despite head or tail winds in a
systematic manner instead of using approximation. Most convenient method is the graphical method
as explained below.

The Graphical Method of Finding Gliding Speeds for Maximum Range

31. The graphical method of finding the best


speed to glide for maximum range in the
presence of a head or tailwind is the similar to
that for finding the optimum range speed for a
piston aircraft. First let us see how to find out
the best speed in case of nil winds condition
from the TAS versus Power Required (i.e. Drag
x TAS) graph (Fig 16-7). Since we need to find
the point of minimum drag, a tangent from the
origin to the curve would represent the minimum
angle at origin or the minimum value of TAS X Fig 16-7: Effect of Winds on Gliding Range
Drag / TAS i.e. Minimum Drag.
FIS Book 1: Aerodynamics 258

32. Slight modification to the TAS versus Power Required graph would allow best gliding speed to
be determined accurately even for windy situation. The simple modification being that instead of
drawing the curve with respect to the TAS it may be drawn with respect to the ‘ground speed’. This
can be easily achieved by simply shifting the origin of the original graph along the x axis (P & Q) to the
appropriate side by an extent equivalent to the wind speed. Now when the tangent is drawn from the
new origin, the resultant speed (W & X) would be the theoretically best speed for glide in the given
wind condition.

Effect of Weight on Glide Performance

33. Effect in Nil Winds. Interestingly,


variation in the weight does not affect the gliding
angle, i.e. irrespective of the weight of the glider, it
would travel to the same point on the ground and
subtend the same gliding angle provided it
maintains the same angle of attack. For best
range conditions it means that the angle of attack
should correspond to the best L / D value. This
happens because (Fig 16-8) the higher lift required
due to increased weight is automatically provided Fig 16-8: Effect of Weight on Gliding
Range
by the increased weight since it is a component of
weight (W cos θ ) that provides the lift during the glide. Similarly, the increase in drag is also in equal
proportion to the increased lift since drag is equal to W sin θ, leaving the ratio between lift and drag
and hence L/D unchanged, provided the speed for glide is adjusted to maintain the required lift and
AOA. Accordingly, the ROD in the glide would vary directly with the speed since ROD = V sin θ. The
best EAS for the changed weight can be worked out by using the familiar 'lift’ equation and it will be
seen that the best EAS varies as the square root of the AUW. A simple method of estimating
changes in the EAS to compensate for changes in the AUW up to about 20% is to decrease or
increase the air speed by half the percentage change in the AUW. For example, a weight reduction of
10% necessitates a drop in air speed of 5%, an increase of weight of the same amount would entail a
5% increase in glide speed. Usually the exact distance covered during glide except in pure gliders is
not vitally important, therefore the gliding speed stated in the Aircrew Manual is a mean figure
applying to the lower weights of a particular aircraft and giving the best all-round performance in the
glide.

34. Effect in Winds. Where weight does affect the range is when there is a tailwind or
headwind component. The higher TAS for a heavier weight means higher ROD allows less time for
the wind to affect the aircraft and so it is better to have a heavier aircraft if gliding for range into wind,
converse is true in a tail wind. Consider an aircraft with a failed engine trying to glide to an airfield. If
this aircraft had the option of reducing its weight by dumping cargo or fuel, it would be advisable to
259 Climbing and Gliding

reduce its weight for the glide because even though it would glide the same distance irrespective of its
weight, it would have a lesser ROD giving the pilot some thinking time. However, if there was a head
wind, it may be beneficial, theoretically at least, to retain the excess weight since the higher ensuing
speed and hence the higher ROD would ensure that the detrimental effect of wind is offset to an
extent. In a tail wind, reducing the weight would have the benefit of not only more time but also
increased range because of staying afloat in tail wind for longer time (as discussed earlier in para 28 –
30). If minimum rate of descent is required, then the aircraft should be light. The lower drag requires
a less rapid expenditure of power, which is obtained from the aircraft’s potential energy (height).

35. To sum up the practical implications of most of the above factors, consider the following. If a
Kiran pilot experiences an engine failure away from base then irrespective of the altitude the pilot
needs to maintain a constant EAS of 130 Kts to ensure the optimum AOA and hence maximum
gliding range. If the pilot maintains the correct speed corresponding to the weight, the aircraft would
glide the same distance irrespective of the fuel (weight) of the aircraft. If a head wind were prevailing
then the Kiran would suffer a loss of glide range, however the loss of range in a heavier Kiran would
be comparatively smaller. If a tailwind was prevailing, the Kiran would gain on glide range, however,
a lighter Kiran would gain more on the glide range in such a situation. The range would not be
affected by changes in weight but the endurance would decrease with increase of weight and vice
versa. If two Kirans at same height but with different weights were to start a glide simultaneously,
then to maintain best glide conditions both aircraft would have to fly at the same angle of attack i.e. at
the best L / D value, the heavier Kiran though would have to maintain a higher EAS and ROD,
hence would cover the distance between the starting point and the touchdown in a shorter time. Both
would, however, cover the same distance in still air. While flying at endurance speeds, the endurance
speed of the heavier Kiran would be higher but its endurance would be lesser.
FIS Book 1: Aerodynamics 260
261

CHAPTER 17

MANOEUVRES

Introduction

1. Changes in the attitude of an aircraft in flight can take place in any one, or combination of, the
three major axes described below. During manoeuvres considerable forces are at work on the
airframe and these may be large enough to cause damage or even structural failure if the aircraft is
manoeuvred without consideration of the limits for which the airframe has been designed.

2. This chapter looks at the nature of the forces which affect limitations, the build-up of the
manoeuvre envelope and the theory of level turns as a specific manoeuvre. The Annex to this
chapter gives, in greater detail, the operating strength requirements, aircraft loads and operating
limitations.

Axes of Movement of an Aircraft

3. Lateral Axis. The lateral axis is a straight line through the CG, normal to the plane of
symmetry, rotation about which is termed pitching. This axis may also be known as the pitching or
looping axis. If any component of the forward flight velocity acts parallel to this axis the subsequent
motion is called sideslip or skid.

4. Longitudinal Axis. The longitudinal axis is a straight line through the CG, fore and aft in
the plane of symmetry, movement about which is known as rolling. This axis is sometimes called the
roll axis.

5. Normal Axis. The normal


axis is a straight line through the CG at
right angles to the longitudinal axis, in
the plane of symmetry, movement
about which is called yawing. This axis
can be referred to as the yawing axis.

6. Fixed Relationship. The


three axes are fixed relative to the
aircraft irrespective of its attitude. Fig
17-1 shows the major axes and the Fig 17-1: The Three Major Axes
possible movements about them.
FIS Book 1: Aerodynamics 262

Acceleration

7. Any aircraft or body in motion is subject to three laws of motion (Newton's laws) which state
that:

(a) All bodies tend to remain at rest or in a state of uniform motion in a straight line
unless acted upon by an external force, i.e. they have the property of inertia.

(b) To change the state of rest, or motion in a straight line, a force is required. To obtain
a given rate of change of motion and/or direction the force is proportional to the mass of the
body. It follows that for a given mass, the greater the rate of change of speed and/or direction
the greater is the force required.

(c) To every action there is an equal and opposite reaction.

8. Most manoeuvres involve changes in direction and speed, the degree of change depending
on the manoeuvre involved. Any change of direction and/or speed necessarily involves an
acceleration, which is often evident to the pilot as an apparent change in his weight. During an
acceleration the aircraft is not in equilibrium since an out-of-balance force is required to deflect it
continuously from a straight line.

9. While a body travels along a curved path it tries constantly to obey the first law and travel in a
straight line. To keep it turning a force is necessary to deflect it towards the centre of the turn. This
force is called centripetal force and its equal and opposite reaction is called centrifugal force.
Centripetal force can be provided in a number of ways. A weight, on the end of a piece of string that
is swung in a circle is subjected to centripetal force by the action of the string. If the string is released
the centripetal force is removed and the equal and opposite reaction (centrifugal force) disappears
simultaneously. The weight, conforming to the first law, then flies off in a straight line at a tangent to
the circle.

10. To clarify the difference


between centripetal and centrifugal
force, consider a body (Fig 17-2 A)
which moves along a series of straight
lines AB, BC, CD and DE, each inclined
to its neighbour at the same angle.
When it reaches B it is subjected to an
external force f1 acting at right angles to
AB, which alters its path to BC. At C a
force f2, acting at right angles to BC, Fig 17-2: Centripetal Force
263 Manoeuvres

alters the path to CD and so on. If the path from A to E consists of a greater number of shorter lines
(Fig 17-2 B) the moving body will be deflected at shorter intervals, and if the lines are infinitely short
the intermittent forces blend into one continuous force F and the path becomes the arc of a circle (Fig
17-2 C). F will act at right angles to the direction of motion, i.e. towards the centre of the arc, the
reaction to it, the centrifugal force, is equal in size and opposite in direction.

Gravity

11. The symbol g denotes the rate of acceleration of a body falling freely under the influence of its
own weight, i.e. the force of gravity. The acceleration is about 9.8 m/sec2 at the Earths’ surface when
measured in a vacuum so that there is no drag acting upon the body. The force of gravity varies with
the distance from the Earth's centre and therefore differs slightly at different points on the Earth's
surface, since the Earth is not a perfect sphere. The weight of a body of a stated density and volume
(mass) is proportional to the force of gravity and so varies slightly. For practical purposes g can be
considered to be constant at sea level, irrespective of the geographical location. As altitude is
increased, g decreases progressively, but in unaccelerated flight this effect is negligible at current
operating heights.

12. If any object has, for example, a total force acting upon it equal to five times its own weight it
will accelerate in the direction of the force at a rate five times greater than that due to gravity, namely
5 g. Although g is a unit of acceleration it is often used, inaccurately, to indicate the force required to
produce a given acceleration. Among pilots g is often used in this way to express the force
accompanying a manoeuvre in terms of a multiple of the static weight. For example, if a certain rate
of turn (i.e. acceleration) necessitates a centripetal force of three times the aircraft weight then the
turn is called a 3 g turn. Since the force is felt uniformly throughout the entire aircraft and its contents,
the crew also experiences this force and acceleration and feel it as an apparent increase in weight
which is proportional to the g. During straight and level flight an aircraft accelerometer shows 1 g, for
this is the normal force of gravity that is acting at all times on all objects. During inverted level flight
the instrument reads -1 g and the pilot's weight, acting vertically downwards, is supported by his
harness.

13. Calculation of Centripetal Force. The magnitude of the centripetal force during a given
turn is directly proportional to the mass of the body (its static weight) and the square of the speed, and
is inversely proportional to the radius of turn (Equation 1.5). It is calculated from the formula

F = W V2 kg
gr
Where W is the weight (kg)
V the TAS (m/s)
r the radius of turn (m)
g is a constant 9.8 m/s2.
FIS Book 1: Aerodynamics 264

THE MANOEUVRE ENVELOPE

General

14. The manoeuvre envelope (or V-n diagram) is a graphic representation of the operating limits
of an aircraft. The envelope is used:

(a) To lay down design requirements for a new aircraft.


(b) By manufacturers to illustrate the performance of their products.
(c) As a means of comparing the capabilities of different types.

The axes of the envelope are:

(d) On the Y-axis (vertical axis), load factor (n), positive and negative.
(e) On the X-axis (horizontal axis), EAS.

The Limits to the Basic Envelope (Theory)

15. One of the more obvious limits to the operation of an aircraft is the increase of the stalling
speed as the load factor is increased. From basic theory:

load factor (n) = _Lift__


Weight
Thus n = Lift at manoeuvre stalling speed (VM)
Lift at basic stalling speed (VB)
n = CL max ½ ρ VM2 S
CL max ½ ρ VB2 S
n = VM2 , so VM = VB √n
VB2

It should be noted from Fig 17-3 that flight


below the basic stalling speed is perfectly
feasible at load factors below one.

16. g Limits. Any aircraft is


designed to certain strength requirements.
Fighter and training aircraft need to be
stronger than transport aircraft. Extra
strength normally means an increase in
structural weight which in turn would need
larger engines. Most aircraft therefore are
designed to a bare minimum strength
Fig 17-3: Manoeuvre Stalling Speed
which depends on their role. Normally the
265 Manoeuvres

training and fighter aircraft are built to withstand about + 7 g in use, which is about the maximum that
an average human pilot can stand and still remain conscious. While the transport aircraft are built to
about +2 g. Both classes usually have
much lower negative g limits (e.g. about -3½
g and -½ g respectively). These values
usually have a 50% safety margin above
which permanent deformation and failure
may occur. However this does not mean
that lesser damage will not occur. Small
excesses may cause rivets to "pop" or
panels to come off, and as the inroads into
the safety margin increase so more severe
damage may occur. Since g = lift / weight,
the g limits need to be reduced if the
weight of the aircraft is increased in order to
Fig 17-4: g Boundary
retain the same safety margins (Fig 17-4).

17. EAS Limitation. An aircraft has a speed limit, (or maximum permissible diving speed) with
a "small" (usually 5 or 10%) safety factor. Exceeding this may lead to loss of access panels, failure of
the weakest structure (often the tailplane or canopy) or even control reversal. An aircraft in flight
above the limit speed may encounter the following:

(a) Critical gust (Refer annex to this chapter).


(b) Destructive flutter (Refer chapter on Flight Controls).
(c) Aileron reversal (Refer chapter on Flight Controls).
(d) Wing Surface Divergence
(d) Critical compressibility effects such as stability and control problems, damaging
buffet, etc. (Refer chapter on High Speed Aerodynamics).

18. The occurrence of any one of these items


could cause structural damage or failure of the
primary structure. A reasonable accounting of
these items is required during the design of an
aeroplane to prevent such occurrences in the
required operating regions. This limit completes
the basic manoeuvre envelope, with operation
outside the envelope either impossible or unsafe
(Fig 17-5). Whatever the resulting limit
airspeed happens to be, it deserves due
respect.
Fig 17-5: EAS Limit
FIS Book 1: Aerodynamics 266

Variations in CL max

19. The boundaries of the basic envelope constructed up to this point have been drawn on the
assumption that CL max remains constant. However, when plotted against Mach number, CL max is seen
to vary with changes of compressibility, RN and adverse pressure gradient.

20. Compressibility and its effects are dealt with fully in the chapter on Transonic And Supersonic
Aerodynamics, but for the purpose of this chapter it may be said that increasing the angle of attack
increases the local acceleration over the wing
so that the critical Mach number (MCRIT) is
reached at a progressively lower free stream
Mach number (MFS). The subsequent
shockwaves are nearer the leading edge and
as the flow behind the shock breaks away and
is turbulent, a greater part of the wing is
subject to separated flow and the loss of lift is
greater. Fig 17-6 shows typical variations at
different angles of attack with Mach number.
With moderately thick wings this reduction
could be at an extremely low MFS. Fig 17-6: Variation in CL with Mach Number

21. The increase in Reynolds number due


to speed only slightly affects the slope of CL
curve, i.e. when plotted against speed
(ignoring compressibility), CL is a straight line.
Its effect on CL max can however be quite
marked because it delays separation to a
higher angle of attack with the consequent
increase in CL max. When plotting CL max

against speed, theory would indicate that CL


max would increase until compressibility effects
become predominant, when it would decline
(Fig 17-7). As speed increases, the adverse
pressure gradient on the upper surface at high Fig 17-7: Effect of Increased RN on CL

angles of attack becomes stronger. This causes the boundary layer to slow down and thicken. The
separation point is moved forward and the CL max reduces.

22. For a variety of reasons the whole aircraft may be incapable of using the highest angle of
attack, so that the usable CL max is reduced. The reasons for this vary among types, but may include
the following:
267 Manoeuvres

(a) Reduction in effectiveness of the elevator, due to its operation behind a shockwave
on either the wing or tailplane.

(b) Buffet on the tailplane from the thick fluctuating wake of the wing, making it
impossible to accurately hold angles of attack near the stall. As a result a lower angle has to
be chosen.

(c) Wing tip stalling or the increase in downwash over the tailplane caused by the large
vortices of a swept-wing aircraft may lead to pitch-up, again limiting the usable angles of
attack.

23. What happens in practice depends very much on the wing in question but it is generally true
that the theoretical CL max cannot be used because control problems due to separated flow prevent the
required angle of attack being selected. Generally, the obtainable maximum value of CL for subsonic
and most transonic aircraft decreases from a very
low Mach number along the dashed line in Fig 17-6.
For supersonic aircraft the CL max may decline with
increasing Mach number in the subsonic range, but
usually recovers as the speed exceeds M 1.0.

The Modified Envelope due to Reduction in CL max

24. The reduction in CL max means that the wing


gives less lift than was expected at any particular
speed. Consequently the available load factor is
reduced and a new lift boundary curve must be Fig 17-8: Lift Boundary
produced (Fig 17-8).

25. As each EAS corresponds to a


higher Mach number at altitude, so the
CL max will decline further and new
curves must be drawn for each altitude,
although the Mach number at which the
maximum load factor occurs will remain
the same (Fig 17-9). It should be noted
that a lift boundary also exists for
negative load factors and that a different
lift boundary exists at different weights. Fig 17-9: Variation of Lift Boundary with Altitude
FIS Book 1: Aerodynamics 268

Structural Limitations

26. Mach Limit. Aircraft designed for subsonic or transonic flight usually have a compressibility
Mach number. If this is exceeded the position and size of the shockwaves make control difficult or
impossible and could even lead to
structural failure. At high angles of attack
the airflow is accelerated more and the
wave development will occur at a lower
Mach number. As a result, this limit
appears as a curved line (Fig 17-9). In the
complete envelope diagram, because the
EAS/Mach number relationship changes
with height, Mach number limits must be
drawn for each altitude. These are shown
in Fig 17-10. Fig 17-10: The Complete Flight Envelope

27. Rolling g Limit. A rolling g limit


is an additional limit imposed on the
operation of an aircraft when the roll
control is deflected. It is a lower figure
than the ordinary g limit because the wing
structure has to provide strength to
withstand the twisting forces caused by
the roll control deflection besides providing
strength to withstand the normal g. It may
also be imposed to avoid structural failure Fig 17-11: Rolling g Limit

due to divergence caused by inertia cross-


coupling (Fig 17-11).

28. Buffet Corners. If high air loads are combined with high loadings, the weakest part of the
structure is more likely to fail. In the case of the tailplane this will be further aggravated if it is buffeted
by the turbulent wake of the wings, possibly leading to fatigue failure.

29. Other Limits. Other limits which could be imposed include considerations for different
weights, fuel states, stores, CG positions, etc.

Information Available from a Manoeuvre Envelope

30. When all the limits are drawn together they show the available range of manoeuvre for the
aircraft at any height. The envelope shows:
269 Manoeuvres

(a) Basic stalling speed.


(b) Available load factor at any height and speed.
(c) Maximum EAS at any height.
(d) Stalling speed at any height and load factor.

These points are, of course, in addition to the Aircrew Manual limitations, which can be read directly
from the envelope:

(e) Maximum permitted load factor.


(f) Maximum EAS.
(g) Compressibility Mach number.
(h) Rolling g limits.

31. Manoeuvre Speed. There are two points of great importance on the V-n diagram (Fig
17-11). Firstly, the point of intersection of the negative limit load factor and line of
maximum negative lift capability (Point A). Any airspeed greater than the speed
corresponding to the intersection provides a negative lift capability sufficient to damage the
aircraft (Point B). Any airspeed less than that speed does not provide negative lift capability
sufficient to damage the aeroplane from excessive flight loads. The second point is the
intersection of the positive limit load factor and the line of maximum positive lift capability
(Point C). The airspeed at this point is the minimum airspeed at which the limit load can
be developed aerodynamically (Point D). Any airspeed greater than that provides a positive lift
capability sufficient to damage the aeroplane. Any airspeed less than that does not provide
positive lift capability sufficient to cause damage from excessive flight loads. The usual term given to
the speed at this point is called the "manoeuvre speed,'' since consideration of subsonic aerodynamics
would predict minimum usable turn radius to occur at this condition. The manoeuvre speed is a
valuable reference point since an aeroplane operating below this point cannot produce a damaging
positive flight load. Any combination of manoeuvre and gust (as explained in the annex to this
chapter) cannot create damage due to excess air load when the aeroplane is below the
manoeuvre speed.

32. Turbulence Speed. When it is impossible to avoid turbulent conditions and the
aeroplane must be subject to gusts, the flight condition must be properly controlled to minimize the
effect of turbulence. If possible, the aeroplane airspeed and power should be adjusted prior to
entry into turbulence to provide a stabilized attitude. Obviously, penetration of turbulence
should not be accomplished at an excess airspeed because of possible structural damage. On
the other hand, an excessively low speed should not be chosen to penetrate turbulence for the
gusts may cause stalling of the aircraft and difficulty of control. To select a proper penetration
airspeed the speed should not be excessively high or low. The "manoeuvre" speed is an
important reference point since it is the highest speed that can be taken to alleviate stall due to
FIS Book 1: Aerodynamics 270

gust and the lowest speed at which limit load factor can be developed aerodynamically. The
optimum penetration speed occurs at or very near the manoeuvre speed.

33. Limitations due to Power Unit.


The thrust boundary shows the maximum
altitude and/or speed range available at
any given load factor. Just as the
manoeuvre envelope shows the available
range of manoeuvre for the aircraft at any
height so the thrust boundary shows the
limitations due to the amount of thrust
available. A thrust boundary shows at
what speeds, altitudes and load factors
the thrust available from the engine is
equal to the drag produced by the whole
Fig 17-12: Thrust Boundary
aircraft when in straight and level flight or
in a level turn. Since the thrust varies with altitude and the drag varies with speed a typical boundary
is as shown in Fig 17-12.

TURNING

General

34. For an aircraft to turn, centripetal force is required to deflect it towards the centre of the turn.
By banking the aircraft and using the horizontal component of the now inclined lift force, the
necessary force is obtained to move the aircraft
along a curved path.

35. If the aircraft is banked, keeping the


angle of attack constant, then the vertical
component of the lift force will be too small to
balance the weight and the aircraft will start to
descend. Therefore, as the angle of bank
increases, the angle of attack must be increased
progressively by a backward movement of the
control column to bring about a greater total lift.
The vertical component is then large enough to
maintain level flight, while the horizontal
component is large enough to produce the
required centripetal force. Fig 17-13: Forces in a Turn
271 Manoeuvres

Effect of Weight

36. In a steady level turn, if thrust is ignored, then lift is providing a force to balance weight and a
centripetal force to turn the aircraft. If the same TAS and angle of bank can be obtained, the radius of
turn is basically independent of weight or aircraft type. However not all aircraft can reach the same
angle of bank at the same TAS.

The Minimum Radius Turn

37. From Fig 17-13 it can be seen that:

tan φ = CPF = WV2 = V2


W Wgr gr
Where φ is the angle of bank
V is the TAS in mps
r is in m
g is 9.8m/s2

Thus r = __V2__ (17.1)


g tan φ

However, not all aircraft can achieve the same angle of bank at the same TAS and so the
equation for radius may therefore usefully be broken down further:

The V in the formula is the manoeuvre stalling speed (VM).

_________
VM = VB √ load factor
= VB √ n
But n = _L_ = __1__ (from Fig 17-13)
W cos φ
Thus VM2 = VB2 __1__
cos φ

Substituting in equation (17.1)

r = __ VB2 X 1__
g tan φ cos φ
∴ r = __VB2__ (17.2)
g sin φ

At the basic stalling speed:

L = W = CL max ½ ρ VB2 S
∴VB2 = ___2W___
CL max ρ S
FIS Book 1: Aerodynamics 272

Substituting in equation (17.2)

r = ____1____ . _W_ . _1_ . _2_ (17.3)


CL max sin φ S ρ g

38. This means that for minimum radius:

(a) Wing loading W / S must be minimum.

(b) ρ must be maximum, this occurs at MSL..

(c) ____1_____ must be a minimum. _1_ reduced towards unity as φ increases


CL max sin φ sin φ
towards 90°. This requires an increase in speed to obtain the necessary increased lift. The
speed should be increased up to that speed where ____1_____ stops decreasing
CL max sin φ
because of a reducing CL max or simply
said, the maximum value of CL and
angle of bank must be obtained. Note
that this does not say maximum angle of
bank for the following reasons. The
angle of bank is increased to provide the
extra lift force which is needed to give
the centripetal force towards the centre
of the turn. To do this, when already at
critical angle of attack, the speed must
be increased. An increase in speed
may cause a reduction in CL max Thus
bank and CL max may not be at the
Fig 17-14: Theoretical Radius of Turn
maximum value at the same time.

39. If the aircraft is kept on the verge of the


stall to obtain CL max and the speed is steadily
increased from the basic stalling speed, the
radius will decrease as bank is increased until,
in theory, the minimum radius (equal to VB2 / g
since the maximum value of sin θ = 1) is
obtained at infinite speed (Fig 17-14). However,
the CL max changes with increasing Mach
number. Therefore if, at a given speed the CL is
less than expected, then less lift is generated
Fig 17-15: Variation in Minimum Radius
and as can be seen in Fig 17-13, the centripetal Turns with Speed and Altitude
273 Manoeuvres

force is reduced. Less bank can now be applied and so the minimum radius starts to increase as
speed is increased beyond a certain value. The overall result of this reduction in CL max on the
minimum radius is shown in Fig 17-15.

40. Altitude. Fig 17-15 shows that there is an increase in the minimum radius with
increase in altitude. This is commonly explained as being due to the EAS/TAS relationship. An
additional increase in minimum radius is also caused by the greater reduction in CLmax because the
Mach number is higher at altitude for a given TAS. In addition it should be remembered that thrust is
also reduced at altitude, although so far thrust has been ignored.

The Maximum Rate Turn

41. The rate of turn is given by the expression V / r Radians Per Second. From Fig 17-14:

sin φ = CPF = W . V2 . 1_ (17.4)


L g r L
Rate of turn (ω) = V
r
∴ ω = L sin φ g . (Substituting r from equation 17.4)
WV
But L = CL ½ ρ V2 S
∴ ω = CL ½ ρ V2 S sin φ g
WV
ω = (CL V Sin φ) . ρ . _g_ . _S_ (17.5)
2 W

42. This means that for a maximum rate turn:

(a) (CL V sin φ) must be a maximum. Sin φ increases as φ increases towards 90°. This
bank increase requires more speed, so
speed should be increased as long as CL
does not decrease faster than (V sin φ) is
increasing. OR simply said, the maximum
value of the product of angle of bank, speed
and CL value must be obtained. Again note
that this does not say maximum angle of
bank for the following reason. The angle of
bank is increased to provide the increased
lift force which is needed to give the
centripetal force towards the centre of the
turn. To do this when already at the critical Fig 17-16: Variation in Maximum Rate
Turns with Speed and Altitude
angle of attack the speed must be
FIS Book 1: Aerodynamics 274

increased. An increase in speed may cause a fall in the maximum value of CL (Fig 17-6).
Thus bank, speed and CL may not all be at maximum value at the same time.

(b) Wing loading W / S must be minimum.

(c) ρ must be maximum i.e. at MSL.

43. The speed for a maximum rate turn will be where a tangent from the origin just touches the
minimum radius curve (Fig 17-16). On the theoretical curve, with CL max constant, the tangent is at
infinite speed which is the same as the speed for minimum radius. However, when the practical
reduction in CL max with increasing Mach number is taken into account, the speed for a maximum rate
turn is found to be somewhat higher than that for minimum radius. Increasing altitude will cause the
rate of turn to decrease.

44. Maximum g Loading. The speed for maximum g loading is higher than the speed for a
maximum rate turn. If CL max were constant, the speed for minimum radius, maximum rate and
maximum g loading would be the same. However CL max decreases with increasing Mach number,
maximum g loading (or nmax) depends on the lift produced, the acceleration towards the centre of the
turn V2 / r is proportional to this lift (Fig 17-13). Therefore if the speed for maximum V2/ r is found this
will give nmax. In Fig 17-17, r / V is plotted
against V. The speed at which the tangent from
the origin just touches this curve will be that for
maximum V2 / r and therefore maximum g
loading. This is clearly at a higher speed than
the minimum r / V (or maximum V / r which is
maximum rate).

Effect of Thrust Fig 17-17: Variation of r/V with V


45. So far the effect of thrust on
level turns has been ignored. However,
the thrust, or lack of thrust, may be the
determining factor as to whether the
optimum speeds referred to in paras 36
to 42 can even be attained. This will
depend on the thrust boundary (para
31). Even in level flight it can be seen
clearly with some aircraft that a
component of thrust is acting in the
same direction as lift due to the Fig 17-18: Thrust Vectors
inclination of the thrust line from the
275 Manoeuvres

horizontal. This effect becomes more pronounced as the critical angle of attack is approached (Fig
17-18). In the minimum radius and maximum rate turns discussed, the aircraft is flown for CL max

which is obtained at the critical angle. The thrust component assists lift so that either less lift is
required from the wing (as simple turning theory requires for a particular turn) or the turn can be
improved beyond that indicated by simple theory. Just as lift was split into two components in Fig 17-
13, one to oppose weight and one to provide centripetal force, so the component of thrust that acts in
the same direction as lift can also be split into two similar components. This is the reason for the
remarkably small radius of turn of which some high performance aircraft like the Mig-29 and SU-30
(using thrust vectoring) are capable of. However, the reduction of thrust with increasing altitude will
cause a reduction in turning performance in addition to that caused by the TAS/EAS relationship and
the greater CL max reduction.

46. It should be realized that many aircraft do not have sufficient thrust to reach and sustain the
optimum speed for a minimum radius turn at constant altitude. In this case the speeds for minimum
radius, maximum rate and maximum g will all be the same at the maximum speed that can be
sustained at constant altitude.

47. Effect of Flap. The lowering


of flap produces more lift but also
more drag at any given EAS. A
smaller radius turn may be obtained
when flap is lowered provided the flap
limiting speed is not a critical factor
and provided that there is sufficient
thrust to overcome the extra drag.
The theoretical absolute minimum
radius will have been decreased since
it depends on VBf2 / g where VBf is
the new stalling speed with flap. Two
possible cases, one where lowering
flap is beneficial and one where it is
not, are illustrated in Fig 17-19. Fig 17-19: Comparison of Turning Radius With and
Without Flap

Nomogram of Turning Performance

48. Fig 17-20 is a nomogram applicable to any aircraft, from which considerable information can
be obtained on turning performance. Some examples of the use of the nomogram are given below.

49. If the TAS and altitude are known, then the Mach number can be determined. Assume a TAS
of 400 kt at 30,000 ft. The Mach number is found by placing a rule so that 400 kt (on the TAS scale)
FIS Book 1: Aerodynamics 276

and 30,000 ft (on the "height in


thousands of feet" scale) are in
alignment and then reading off the
Mach number from its scale. This
example is shown by the upper
dashed line on the nomogram and
gives a Mach number of 0.68M.
The example can equally well be
worked in reverse, i.e. the altitude
and Mach number can be used to
determine the TAS.

50. If the TAS and bank angle


are known then the rate of turn
can be found. Assume a TAS of
400 kt and an angle of bank of
60°, place a rule on the
nomogram to align these figures
and then read off the rate of turn,
in this case 284° per min, about
rate 1½. The bank scale also
shows the g realized in a
sustained turn at that angle of
bank. The Mach scale has no
significance in these examples.
Any two known factors can be Fig 17-20: Nomogram of Turning Performance

aligned to determine the unknown


third factor; e.g. to find the bank angle corresponding to a rate 1 turn at 500 kt TAS the rule is aligned
on these figures and the bank angle read off from its scale. The nomogram shows clearly that as
speed increases the angle of bank must be increased to maintain a constant rate of turn (See
examples in Table 17-1).

Table 17-1: Variation in Bank Angle with Speed for Same Rate of Turn

Rate of Turn Speed to Bank angle g


1 200 28° 1.15
1 400 48° 1.5
1 600 58° 2
277 Manoeuvres

SUMMARY

Manoeuvring Considerations

51. During a manoeuvre an aircraft is not in a state of equilibrium since an out-of-balance force is
required to deflect it continuously from a straight line. This force is called centripetal force.

52. The manoeuvre envelope is used:

(a) To specify design requirements for a new aircraft.


(b) By manufacturers to illustrate the performance of their products.
(c) As a means of comparing the capabilities of different types.

53. The axes of the manoeuvre envelope are load factor and velocity, and from it can be read:

(a) Basic stalling speed.


(b) Available load factor at any height and speed.
(c) Maximum EAS at any height.
(d) Stalling speed at any height and load factor.
(e) Maximum permitted load factor.
(f) Maximum EAS.
(g) Compressibility Mach number.
(h) Rolling g limits.

54. Variations in CL max are caused by:

(a) Compressibility effects:


(i) Separation behind shockwaves.
(ii) Reduction of control effectiveness behind shockwaves.
(iii) Buffet on the tailplane making it impossible to hold an accurate angle of
attack.

(b) Change in Reynolds number.

(c) Increase in the adverse pressure gradient with speed.

55. For minimum radius turns the following must be satisfied:

(a) Wing loading must be as low as possible.


(b) The air must be as dense as possible.
(c) The maximum value of the product of CL value and angle of bank must be obtained
(CL MAX Sinϕ).
FIS Book 1: Aerodynamics 278

56. For maximum rate turns the following must be satisfied:


(a) Wing loading must be as low as possible.
(b) The maximum value of the product of angle of bank, speed and CL value must be
obtained (V CL Sin ϕ).

In both cases the use of thrust will improve turning performance.

(c) The air must be as dense as possible.


279

ANNEX TO CHAPTER 17

OPERATING STRENGTH LIMITATIONS

Introduction

1. The weight of the structural components of an aircraft is an extremely important factor in the
development of an efficient aircraft configuration. In no other field of mechanical design is there such
necessary importance assigned to structural weight. The efficient aircraft and powerplant structure is
the zenith of highly refined minimum weight design. In order to obtain the required service life from his
aircraft, the pilot must understand, appreciate, and observe the operating strength limitations. Failure
to do so will incur excessive maintenance costs and a high incidence of failure during the service life
of an aircraft.

General Definitions and Structural Requirements

2. There are strength requirements which are common to all aircraft. In general, these
requirements can be separated into three particular areas. These are detailed in the following
paragraphs.

Static Strength

3. The static strength requirement is the consideration given to the effect of simple static
loads with none of the ramifications of the repetition or cyclic variation of loads. An important
reference point in the static strength requirement is the "limit load" condition. When the aircraft
is at the design configuration, there will be some maximum load, which would be anticipated
from the mission requirement of the aeroplane. For example, a fighter or attack type aircraft, at
the design configuration, may encounter a peak load factor of 7.5 in the accomplishment of its
mission. Of course, such an aircraft may be subject to load factors of 3, 4, 5, 6, 1, etc., but no more
than 7.5 should be required to accomplish the mission. Thus, the limit load condition is the
maximum load anticipated in normal operation of the aircraft. Various types of aircraft will
have different limit load factors according to the primary mission of the aircraft. Typical
values are tabulated below:

Type of aircraft Positive limit load factor


Fighter or attack 7.5
Trainer 7.5
Transport, patrol, antisubmarine 3.0 or 2.5

These examples are quite general and it is important to note that there may be variations
according to specific mission requirements.
FIS Book 1: Aerodynamics 280

4. Since the limit load is the maximum of the normally anticipated loads, the aircraft
structur e mus t w i ths tan d th is l oad w i th no i ll effects. Specifically, the primary structure
of the aircraft should experience no objectionable permanent deformation when subjected to
the limit load. In fact, the components must withstand this load with a positive margin. This
requirement implies that the aircraft should withstand successfully the limit load and then return
to the original unstressed shape when the load is removed. Obviously, if the aircraft is
subjected to some load, which is in excess of the limit load, the overstress may incur an
objectionable permanent deformation of the primary structure and require replacement of the
damaged parts.

5. Many different flight and ground load conditions must be considered to define the most
critical conditions for the structural components. In addition to positive lift flight, negative lift
flight must be considered. Also, the effect of flap and landing gear configuration, gross weight,
flight Mach number, symmetry of loading, CG. positions, etc., must be studied to account for all
possible sources of critical loads. To verify the capability of the structure, ground static tests
are conducted and flight demonstrations are required.

6. To provide for the rare instances of flight when a load greater than the limit is
required to prevent a disaster, an "ultimate factor of safety" is provided. Experience has
shown that an ultimate factor of safety of 1.5 is sufficient for piloted aircraft. Thus, the
aircraft must be capable of withstanding a load which is 1.5 times the design limit load. The
primary structure of the aircraft must withstand the "ultimate load" (1.5 times limit)
without failure. Permanent deformation may of course be expected with this "overstress"
but no actual failure of the major load-carrying components should take place at ultimate
load. Ground static tests are necessary to verify this capability of the structure.

7. An appreciation of
the static strength
requirements may be
obtained by inspection of
the basic properties of a
typical aircraft metal. Fig
17-21 illustrates the typical
static strength properties of
a metal sample by a plot of
applied stress versus
resulting strain. At low
values of stress the plot of
stress and strain is
Fig 17-21: Static Strength of Typical Aircraft Metal
essentially a straight line,
281 Operating Strength Limitations

i.e., the material in this range is elastic. A stress applied in this range incurs no permanent
deformation and the material returns to the original unstressed shape when the stress is
released. At higher values of stress the plot of stress versus strain develops a distinct curvature in
the strain direction and the material incurs disproportionate strains.

8. High levels of stress applied to the part and then released produce a permanent
deformation. Upon release of some high stress, the metal snaps back, but not all the way. The
stress defining the limit of tolerable permanent strain is the "yield stress" and stresses applied
above this point produce objectionable permanent deformation. The highest stress the material
can withstand is the "ultimate stress.'' Noticeable permanent deformation usually occurs in this
range, but the material does have the capability for withstanding one application of the ultimate
stress.

9. The relationship between the stress-strain diagram and operating strength limits should
be obvious. If the aircraft is subjected to a load greater than the limit, the yield stress may be
exceeded and objectionable permanent deformation may result. If the aircraft is subject to a
load greater than the ultimate load, failure is imminent.

10. Service Life. The various components of the aircraft and powerplant structure must be
capable of operating without failure or excessive deformation throughout the intended service
life. The repetition of various service loads can produce fatigue damage in the structure and
special attention must be given to prevent fatigue failure within the service life. Also, the
sustaining of various service loads can produce creep damage and special attention must be
given to prevent excessive deformation or creep failure within the service life. This is a
particular feature of components which are subjected to operation at high temperatures.

Fatigue Considerations

11. The fatigue strength requirement is the consideration given to the cumulative effect of
repeated or cyclic loads during service. While there is a vague relationship with the static strength,
repeated cyclic loads produce a completely separate effect. If a cyclic, tensile stress is applied to
a metal sample, the part is subject to a "fatigue" type loading. After a period of time, the cyclic
stressing will produce a minute crack at some critical location in the sample. With continued
application of the varying stress, the crack will enlarge and propagate into the cross section.
When the crack has progressed sufficiently, the remaining cross section is incapable of withstanding
the imposed stress and a sudden, final rupture occurs. In this fashion, a metal can fail at stresses
much lower than the static ultimate strength.

12. The time necessary to produce fatigue failure is related to the magnitude of the cyclic
stress. This relationship is typified by the graph of Fig 17-22. The fatigue strength of a material
FIS Book 1: Aerodynamics 282

can be demonstrated
by a plot of cyclic stress
versus cycles of stress
required to produce
fatigue failure. As
might be expected, a
very high stress level
requires relatively few
cycles to produce
fatigue failure. Moderate
stress levels require a
fairly large number of
cycles to produce failure
and a very low stress
may require nearly an Fig 17-22: Fatigue Strength of Typical Aircraft Metal
infinite number of
cycles to produce failure. The very certain implication is that the aircraft must be capable of
withstanding the gamut of service loads without producing fatigue failure of the primary
structure.

13. For each type of aircraft there is a probable spectrum of loads which the aircraft will
encounter. That is, various loads will be encountered with a frequency particular to the mission
profile. The fighter or attack type of aircraft usually experiences a predominance of manoeuvre
loads while the transport or patrol type usually encounters a predominance of gust loads. Since
fatigue damage is cumulative during cyclic stressing, the useful service life of the aircraft must be
anticipated to predict the gross effect of service loads. Then, the primary structure is required
to sustain the typical load spectrum through the anticipated service life without the occurrence of
fatigue failure. To prove this capability of the structure, various major components must be
subjected to an accelerated fatigue test to verify the resistance to repeated loads.

14. The design of a highly stressed or long life structure emphasizes the problems of fatigue.
Great care must be taken during design and manufacture to minimize stress concentrations which
enhance fatigue. When the aircraft enters service operation, care must be taken in the
maintenance of components to insure proper adjustment, torquing, inspection, etc., as proper
maintenance is a necessity for achieving full service life. Also, the structure must not be subjected
to a load spectrum more severe than was considered in design or fatigue failures may occur
within the anticipated service life. With this additional factor in mind, any pilot should have all the
more respect for the operating strength limits. “Recurring overstress causes a high rate of
fatigue damage”.
283 Operating Strength Limitations

Creep Considerations

15. By definition, creep is the structural deformation which occurs as a function of time. If a part is
subjected to a constant stress of sufficient magnitude, the part will continue to develop plastic strain and
deform with time. Eventually, failure can occur from the accumulation of creep damage. Creep
conditions are most critical at high stress and high temperature since both factors increase the rate of
creep damage. Any structure subject to creep conditions therefore should not encounter excessive
deformation or failure within the anticipated service life.

16. The high operating temperatures of gas turbine components furnish a critical environment for
creep conditions. The normal operating temperatures and stresses of gas turbine components create
considerable problems in design for service life. Thus, operating limitations deserve very serious respect
since excessive engine speed (RPM) or excessive turbine temperatures (JPT / TGT) will cause a large
increase in the rate of creep damage and lead to premature failure of components. Gas turbines require
high operating temperatures to achieve high performance and efficiency and short periods of excessive
temperatures can incur highly damaging creep rates.

17. Aeroplane structures can be subject to high temperatures due to aerodynamic heating at high
Mach numbers. Thus, very high speed aeroplanes can be subject to operating speed limitations due
to creep conditions.

Aeroelastic Effects

18. The requirement for structural stiffness and rigidity is the consideration given to the interaction
of aerodynamic forces and deflections of the structure. The aircraft and its components must have
sufficient stiffness to prevent or minimize aeroelastic influences in the normal flight range. Aileron
reversal, divergence, flutter, and vibration should not occur in the range of flight speeds which will be
normal operation for the aircraft.

19. It is important to distinguish between strength and stiffness. Strength is simply the
resistance to load while stiffness is the resistance to deflection or deformation. While Strength and
stiffness are related, it is necessary to appreciate that adequate structural strength does not
automatically provide adequate stiffness. Thus, special consideration is necessary to provide the
structural components with specific stiffness characteristics to prevent undesirable aeroelastic effects
during normal operation.

20. An obvious solution to the apparent problems of static strength, fatigue strength,
stiffness and rigidity would be to build the aeroplane like a product of an anvil works,
capable of withstanding all conceivable loads. However, high performance aeroplane configurations
cannot be developed with inefficient, lowly stressed structures. The effect of additional weight is
FIS Book 1: Aerodynamics 284

best illustrated by preliminary design study of a very long range, high altitude bomber. In the
preliminary phases of design, each additional kilogram of any weight would necessitate a
25-Kg increase in gross weight to maintain the same performance. An increase in the weight of
any item produce a chain reaction. More fuel, larger tanks, bigger engines, more fuel, heavier
landing gear, more fuel, etc. In the competitive sense of design, no additional structural weight can
be tolerated to provide more strength than is specified as necessary for the design mission
requirement.

AIRCRAFT LOADS AND OPERATING LIMITATIONS

Flight Loads: Manoeuvres and Gusts

21. The loads imposed on an aircraft in flight are the result of manoeuvres and gusts. The
manoeuvre loads may predominate in the design of fighter aeroplanes while gust loads may
predominate in the design of the large multiengine aircraft. The manoeuvre loads an aeroplane
may encounter depend in great part on the mission type of the aeroplane. However, the
maximum manoeuvreing capability is of interest because of the relationship with strength
limits.

22. The flight load factor is defined as the proportion between aeroplane lift and weight.

n = _L_
W
Where: n = load factor
L = Lift, Kg
W = Weight, Kg

Manoeuvreing Load Factors

23. The maximum lift attainable at any airspeed occurs when the aeroplane is at C L max. With
the use of the basic lift equation, this maximum lift is expressed as:

L max = CL max ½ ρ V2 S

Since maximum lift must be equal to the weight at the stall speed,

W = CL max ½ ρ VB2 S

If the effects of compressibility and viscosity on CL max are neglected for simplification, the maximum
load factor attainable is determined by the following relationship.

n max = L max = CL max ½ ρ V2 S


W CL max ½ ρ V B2 S
285 Operating Strength Limitations

2
n max = _V (17.5)
VB

24. Thus, if the aeroplane is flying at twice the stall speed and the angle of attack is
increased to obtain maximum lift, a maximum load factor of four will result. Therefore, any
aeroplane which has high speed performance may have the capability of high manoeuvreing load
factors. The aeroplane which is capable of flight speeds that are many times the stall speed
will require due consideration of the operating strength limits.

25. The structural design of the aircraft must consider the possibility of negative load factors
from manoeuvres. Since the pilot cannot comfortably tolerate large prolonged negative "g", the
aircraft need not be designed for negative load factors as great as the positive load factors.

26. The effect of aeroplane gross weight during manoeuvres must be appreciated because of the
particular relation to flight operating strength limitations. During flight, the pilot appreciates
the degree of a manoeuvre from the inertia forces produced by various load factors which the
aeroplane structure senses, the degree of a manoeuvre principally by the airloads involved. Thus,
the pilot recognizes load factor while the structure recognizes only load. To better understand this
relationship, consider an example aeroplane whose basic configuration gross weight is 10,000 Kg.
At this basic configuration assume a limit load factor for symmetrical flight of 5.6 and an ultimate
load factor of 8.4. If the aeroplane is operated at any other configuration, the load factor limits will
be altered. The following data illustrate this fact by tabulating the load factors required to
produce identical airloads at various gross weights.

Gross weight Kg Limit load factor Ultimate load factor


10000 (basic) 5.6 8.4
15000 (max takeoff) 3.73 5.6
6700 (min fuel) 8.4 12.6

27. As illustrated, at high gross weights above the basic configuration weight, the limit and
ultimate load factors may be seriously reduced. For the aeroplane shown, a 5-g manoeuvre
immediately after a high gross weight takeoff could be very near the “disaster regime”,
especially if turbulence is associated with the manoeuvre. In the same sense, this aeroplane at
very low operating weights below that of the basic configuration would experience greatly
increased limit and ultimate load factors.

28. Operation in this region of high load factors at low gross weight may create the impression that
the aeroplane has great excess strength capability. This effect must be understood and intelligently
appreciated since it is not uncommon to have a modern aeroplane configuration with more than 50 percent
of its gross weight as fuel.
FIS Book 1: Aerodynamics 286

Gust Load Factors

29. Gusts are associated


with the vertical and horizontal
velocity gradients in the
atmosphere. A horizontal gust
produces a change in dynamic
pressure on the aeroplane but Fig 17-23: Effect of Vertical Gust
causes relatively small and
unimportant changes in flight load factor. The more important gusts are the vertical gust, which cause
changes in angle of attack. This process is illustrated in Fig 17-23. The vectorial addition of the
gust velocity to the aeroplane velocity causes the change in angle of attack and change in lift. The
change in angle of attack at some flight condition causes a change in the flight load factor. The
increment change in load factor due to the vertical gust can be determined from the following
equation:

∆n = 0.115 m √ σ_ Ve (KU) (17.6)


(W / S)
Where: ∆n = change in load factor due to gust
M = lift curve slope, unit of CL per degree of α
σ = altitude density ratio
W/S = wing loading, psf
Ve = equivalent airspeed, knots
KU = equivalent sharp edged gust velocity ft. per sec.

30. As an example, consider the case of an aeroplane with a lift curve slope m = 0.08 and wing
loading, (W/S) = 60 psf. If this aeroplane were flying at sea level at 350 knots and encountered an
effective gust of 30 ft. per sec., the gust would produce a load factor increment of 1.61. This
increment would be added to the flight load factor of the aeroplane prior to the gust, e.g., if in
level flight before encountering the gust, a final load factor of 1.0+1.61 = 2.61 would result. As a
general requirement all aeroplanes must be capable of withstanding an approximate effective ±30
ft. per sec. gust when at maximum level flight speed for normal rated power. Such a gust intensity
has relatively low frequency of occurrence in ordinary flying operations.

31. The equation for gust load increment provides a basis for appreciating many of the
variables of flight. The gust load increment varies directly with the equivalent sharp edged
gust velocity, KU, since this factor effects the change in angle of attack. The highest
reasonable gust velocity that may be anticipated is an actual vertical velocity, U, of 50 ft per
sec. This value is tempered by the fact that the aeroplane does not effectively encounter the full
effect because of the response of the aeroplane and the gradient of the gust. A gust factor, K
287 Operating Strength Limitations

(usually on the order of 0.6), reduces the actual gust to the equivalent sharp edged gust velocity,
KU.

32. The properties of the aeroplane exert a powerful influence on the gust increment. The lift
curve slope, m, relates the sensitivity of the aeroplane to changes in angle of attack. An aircraft
with a straight, high aspect ratio wing would have a high lift curve slope and would be quite
sensitive to gusts. On the other hand, the low aspect ratio, swept wing aeroplane has a low lift
curve slope and is comparatively less sensitive to turbulence. The apparent effect of wing
loading, W/S, is at times misleading and is best understood by considering a particular aeroplane
encountering a fixed gust condition at various gross weights. If the aeroplane encounters the gust
at lower than ordinary gross weight, the accelerations due to the gust condition are higher. This
is explained by the fact that essentially the same lift change acts on the lighter mass.
The high accelerations and inertia forces magnify the impression of the magnitude of
turbulence. If this same aeroplane encounters the gust condition at higher than ordinary
gross weight, the accelerations due to the gust condition are lower, i.e., the same lift change acts
on the greater mass. Since the pilot primarily senses the degree of turbulence by the
resulting accelerations and inertia forces, this effect can produce a very misleading impression.

33. The effect of airspeed and altitude on the gust load factor is important from the
standpoint of flying operations. The effect of altitude is related by the term √σ which would mean
that an aeroplane flying at a given EAS at 40,000 ft. (σ = 0.25) would experience a gust load
factor increment only one-half as great as at sea level. This effect results because the true
airspeed is twice as great and only one-half the change in angle of attack occurs for a given gust
velocity. The effect of airspeed is illustrated by the linear variation of gust increment with
equivalent airspeed. Such a variation emphasizes the effect of gusts at high flight speeds and the
probability of structural damage at excessive speeds in turbulence.

34. The operation of any aircraft is subject to specific operating strength limitations. A
single large overstress may cause structural failure or damage severe enough to require costly
overhaul. Less severe overstress repeated for sufficient time will cause fatigue cracking and
require replacement of parts to prevent subsequent failure. Each aircraft type has strength
capability only specific to the mission requirement. Operating limitations must be given due
regard.

Landing and Ground Loads

35. The most critical loads on the landing gear occur at high gross weight and high rate of
descent at touchdown. Since the landing gear has requirements of static strength and fatigue
strength similar to any other component, overstress must be avoided to prevent failure and derive
the anticipated service life from the components.
FIS Book 1: Aerodynamics 288

36. The most significant function of the landing gear is to absorb the vertical energy of the
aircraft at touchdown. An aircraft at a given weight and rate of descent at touchdown has a
certain kinetic energy which must be dissipated in the shock absorbers of the landing gear. If the
energy were not absorbed at touchdown, the aircraft would bounce along similar to an
automobile with faulty shock absorbers. As the strut deflects on touchdown, oil is forced through
an orifice at high velocity and the energy of the aircraft is absorbed. To have an efficient strut the
orifice size must be controlled with a tapered pin to absorb the energy with the most uniform force
on the strut. The vertical landing loads resulting at touchdown can be simplified to an extent by
assuming the action of the strut to produce a uniformly accelerated motion of the aircraft. The
landing load factor for touchdown at a constant rate of descent can be expressed by the following
equation:

n = F/W
n = (ROD)2 (17.7)
2gs
Where: n = landing load factor, the ratio of the load in the strut, F, to the
weight, W
ROD = rate of descent, ft. per sec.
g = acceleration due to gravity (32 ft per sec2)
S = effective stroke of the strut, ft.

37. As an example, assume that an aircraft touches down at a constant rate of descent of 18 ft
per sec and the effective stroke of the strut is 18 inches (1.5 ft). The landing load factor for the
condition would be 3.37. The average force would be 3.37 times the weight of the aircraft. (NOTE:
there is no specific correlation between the landing load factor and the indication of a cockpit
mounted g meter. The response of the instrument, its mounting, and the onset of landing loads
usually prevent direct correlation.)

38. This simplified equation points out two important facts. The effective stroke of the strut
should be large to minimize the loads since a greater distance of travel reduces the force
necessary to do the work of arresting the vertical descent of the aircraft. This should
emphasize the necessity of proper maintenance of the struts. An additional fact illustrated is that
the landing load factor varies as the square of the touchdown rate of descent. Therefore, a 20
percent higher rate of descent increases the landing load factor by 44 percent. This fact should
emphasize the need for proper landing technique to prevent a hard landing and over-stress of the
landing gear components and associated structure.

39. The effect of landing gross weight is twofold. A higher gross weight at some landing
load factor produces a higher force on the landing gear. The higher gross weight requires a
higher approach speed and, if the same glide path is used, a higher rate of descent results. In
addition to the principal vertical loads on the landing gear, there are varied side loads, wheel
289 Operating Strength Limitations

spin up and spring back loads, etc., all of which tend to be more critical at high gross weight,
high touchdown ground speed, and high rate of descent.

40. The function of the landing gear as a shock absorbing device has an important application
when a forced landing must be accomplished on an unprepared surface. If the terrain is rough
and the landing gear is not extended, initial contact will be made with relatively solid structure
and whatever energy is absorbed will be accompanied by high vertical accelerations. These high
vertical accelerations encountered with a gear-up landing on an unprepared surface are the
source of a very incapacitating type injury, (vertical compression fracture of the vertebrae). Unless
some peculiarity of the configuration makes it inadvisable, it is generally recommended that the
landing gear be down for forced landing on an unprepared surface. (NOTE: for those prone to
forget, it is also recommended that the gear be down for landing on prepared surfaces.)

Effect of Overstress on Service Life

41. Accumulated periods of overstress can create a very detrimental effect on the useful
service life of any structural component. This fact is certain and irreversible. Thus, the operation of
the aeroplane, powerplant, and various systems must be limited to design values to prevent failure or
excessive maintenance costs early in the anticipated service life. The operating limitations
presented in the pilots notes must be adhered to in a very strict fashion.

42. In many cases of modern aircraft structures it is very difficult to appreciate the effect of a
moderate overstress. This feature is due in great part to the inherent strength of the materials
used in modern aircraft construction. As a general airframe static strength requirement, the primary
structure must not experience objectionable permanent deformation at limit load or failure at 150
percent of limit load (ultimate load is 1.5 times limit load). To satisfy each part of the
requirement, limit load must not exceed the yield stress and ultimate load must not exceed the
ultimate stress capability of the parts.

43. Many of the high strength materials used in aircraft construction have stress-strain diagrams
typical of Fig 17-24. One feature of these materials is that the yield point is at some stress
much greater than two-thirds of the ultimate stress. Thus, the critical design condition is the
ultimate load. If 150 percent of limit load corresponds to ultimate stress of the material, 100 percent
of limit load corresponds to a stress much lower than the yield stress. Because of the inherent
properties of the high strength material and the ultimate factor of safety of 1.5, the limit load
condition is rarely the critical design point and usually possesses a large positive margin of static
strength. This fact alone implies that the structure must be grossly overstressed to produce
damage easily visible to the naked eye. This lack of immediate visible damage with "overstress"
makes it quite difficult to recognize or appreciate the long range effect.
FIS Book 1: Aerodynamics 290

44. A reference point provided on the stress strain diagram of Fig 17-24 is a stress termed the
"endurance limit." If the operating cyclic stresses never exceed this "endurance limit" an infinite (or in
some cases "near infinite") number of cycles can be withstood without fatigue failure. No significant
fatigue damage accrues from stresses below the endurance limit but the value of this endurance
limit is approximately 30 to 50 percent of the yield strength for the light alloys used in aircraft
construction. The rate of
fatigue damage caused by
stresses only slightly above the
endurance limit is insignificant.
Even stresses near the limit
load do not cause a significant
accumulation of fatigue damage
if the frequency of application is
reasonable and within the
intended mission requirement.
However, stresses above the
limit load, and especially
stresses well above the limit
load, create a very rapid rate
Fig 17-24: Typical Stress Strain Diagram for a High
of fatigue damage. Strength Aluminum Alloy

45. A puzzling situation then exists. "Over-stress" is difficult to recognize because of the
inherent high yield strength and low ductility of typical aircraft metals. These same over-stresses
cause high rate of fatigue damage and create premature failure of parts in service. The effect of
accumulated overstress is the formation and propagation of fatigue cracks. While it is sure that
fatigue crack always will be formed before final failure of a part, accumulated overstress is most
severe and fatigue provoking at the inevitable stress concentrations. Hence, disassembly and
detailed inspection is both costly and time-consuming. To prevent in-service failures of a basically
sound structure, the part must be properly maintained and operated within the design "envelope."

46. The operation of any aircraft and powerplant must be conducted within the operating
limitations prescribed in the flight handbook. No hearsay or rumors can be substituted for the
accepted data presented in the aircraft handbook. All of the various static strength, service
life, and aero elastic effects must be given proper respect. An aeroplane can be overstressed
with the possibility that no immediate damage is apparent. A powerplant may be operated
past the specified time, speed, or temperature limits without immediate apparent damage. In
that case, the cumulative effect will tell at some later time when in-service failure occurs and
maintenance costs increases.
291

CHAPTER 18

RANGE AND ENDURANCE

Introduction

I. By range and endurance we imply respectively the greatest distance the aircraft can travel
and the longest time it can remain in flight, for a given quantity of fuel. This problem clearly involves,
in addition to the knowledge of the drag and power characteristics of the aircraft and its power plant, a
consideration of the fuel consumption of the engine(s). Piston and jet engines have different
properties in connection with fuel consumption and for this reason the conditions giving optimum
performance in these respects are quite different.

2. This chapter examines the factors affecting the performance of turbo-jet and piston engine
aircraft with particular reference to fuel consumption and gives the theory underlying the technique of
range flying and endurance flying at subsonic speeds. The practical techniques often differ from the
theoretical in that, simplification is adopted, accepting any small loss in performance. However, unless
the theory is understood the pilot cannot deal efficiently with unexpected circumstances.

Definitions

3. The efficient operation of an aircraft requires that it be flown in a manner which achieves the
delivery of the maximum weight (payload) over the maximum range using minimum fuel. Certain
terms which are used to determine and assess the efficiency are:

(a) Gross Fuel Consumption (GFC). It is the amount of fuel consumed in unit time. Its
unit is N / hr.

GFC = Fuel Consumed (18.1)


Time

(b) Specific Fuel Consumption (SFC). It is the amount of fuel used in unit time to
produce unit thrust, in other words for jets it is the rate of fuel consumption per unit thrust. For
pistons however, it is defined as the amount of fuel used in unit time to produce unit brake
power. SFC is denoted by 'c' and is best when its value is least.

(c) Specific Endurance. Flying for endurance implies flying the aircraft for as
long as possible on the fuel available. In other words, it involves the factor of maximum time
on a given fuel. To be more precise:

Specific Endurance = ___Flight Time___ = t


Quantity of Fuel M
FIS Book 1: Aerodynamics 292

(d) Specific Air Range (SAR). It is the air distance flown per unit quantity of fuel. In SI
units SAR will have unit of Air Metres / Newtons of fuel. However, it is still common to
express SAR in Metres / kg of fuel.

4. For a given engine if c is the fuel flow per unit time to produce unit thrust and if the total thrust
of the engine is T then the total fuel consumed per unit time is T X c which by definition is the GFC.

∴ GFC = SFC X Thrust for jets. (18.2)


and GFC = SFC X Brake Power for pistons. (18.3)

PISTON RANGE

Introduction

5. The power output from a piston aero-engine is derived from burning fuel - air mixture in the
cylinders. This power is transmitted to a common crankshaft and then, through a reduction gear, to
the propeller. In addition to driving the propeller, the crankshaft also drives a centrifugal supercharger
(if fitted), which raises the pressure of the fuel / air mixture fed to the engine, and numerous
accessories such as generators and pumps. The brake horsepower (bhp) developed at the crankshaft
is converted by the propeller to thrust horse power (thp) and thus the piston engine/propeller
contribution produces thrust indirectly by accelerating a large mass of air through the action of a
propeller.

ENGINE CONSIDERATION

Power

6. The rate of fuel consumption of a piston engine is, over a wide range of operating conditions,
approximately proportional to the brake power of the engine. Thus, provided that the propeller
efficiency does not vary much, which implies that the speed is not too low, we may assume that the
rate of fuel consumption is proportional to the power supplied to the aircraft by the engine and
propeller unit.

7. As already described, the output of piston engines is measured in terms of power (bhp), the
amount of which depends basically upon the weight of fuel consumed. This weight is largely
dependent upon:

(a) The pressure of the fuel/air mixture which is fed to the engines. This is measured in
2
Kg / cm and is known as manifold air pressure (MAP).

(b) The rate at which this mixture flows into the engines. This rate is governed by the
293 Range and Endurance

crankshaft speed (rpm) selected.

In addition to the fuel/air mixture strength, the ambient temperature and the action of the super
charger influence power output.

8. MAP and RPM. Cruising power output can be obtained by using a variety of MAP and rpm
settings as follows:

(a) Low MAP and high rpm.


(b) Medium MAP and medium rpm.
(c) High MAP and low rpm.

9. High MAP and low rpm is used whenever possible for reasons of economy. Maximum MAP is
determined by the pressure above which rich mixture has to be used, the minimum rpm is determined
by the engine speed below which rough running, due to an incorrect (too weak) mixture strength takes
place. Extra power is obtained by increasing rpm and less power is obtained by throttling the engine
thereby reducing MAP. Any tendency for the engine to slow down is transmitted to the propeller
constant speed unit (CSU), which adjusts the pitch of the blades, whilst the engine speed is
maintained constant at the rpm selected.

10. Mixture Strength. As already stated, the mixture strength (fuel/air ratio by weight) has to be
enriched (more fuel) at high MAP and rpm settings. This enrichment is necessary to cool the burning
charge and thus avoid detonation, which is a spontaneous explosive and very inefficient combustion.
Although some of the extra weight of charge does contribute to extra power, most of the fuel is
exhausted unburned and is therefore wasted. In cruising flight an efficient, weak mixture should
always be used (approximately15 parts air to one part fuel by weight).

11. Altitude. The power output achieved at altitude is influenced by the ambient air pressure
and temperature, together with the combined action of the supercharger and the carburetor automatic
manifold pressure control, formerly known as the automatic boost control, which prevents the
supercharger from over pressurizing the engine

12. Supercharger. A supercharger usually has the following characteristics:

(a) It is engine-driven at a speed proportional to the rpm selected.


(b) At a particular rpm it is able to raise the pressure between the eye and the periphery
of the centrifugal flow compressor by a set ratio.
(c) The pressure at the eye of the compressor is regulated by the throttle butterfly valve
so that a constant pressure may be delivered to the manifold.
(d) It is often fitted with an optional second gear for use in maintaining power at high
altitudes.
FIS Book 1: Aerodynamics 294

13. Full Throttle Height. At low altitudes, where the ambient air pressures is comparatively
high, the throttle butterfly valve is automatically closed to reduce the pressure drawn to a level which
the engine can accept without detonation. As altitude increases the pressure sensitive automatic
manifold pressure control allows the throttle butterfly valve to open progressively so as to maintain the
pressure on either side of the supercharger despite a reduction in the ambient pressure. When the
butterfly valve has opened completely the engine is operating at full throttle for this rpm setting and
the engine is said to be at full throttle height for the selected MAP and rpm settings. Above this
altitude the MAP can no longer be maintained and the power output falls. However, an increase in
rpm or the selection of a higher supercharger gear would raise the full throttle height, thus maintaining
power at higher altitudes.

14. Fig 18-1 shows how power output,


with maximum weak mixture MAP selected,
varies with altitude when using both
maximum and minimum cruising rpm
settings in both high and low supercharger
gear. The slight increase in power output
from sea level upto full throttle height is due
to the following factors:

(a) The decreasing ambient


pressure against which the engine
has to exhaust the spent charge,
improves the scavenging of the Fig 18-1: Supercharger Gear Ratio

cylinders and in effect lets in more charge into the cylinder.

(b) The decreasing temperature of the charge entering the cylinders causes an increase in
the weight of the same volume of charge being drawn in.

15. The combined effect of these factors is to improve the volumetric efficiency of the engine i.e.
increase the weight of fuel consumed. Since power output ideally should be related to changes in
density i.e. the mass flow of air hence amount of fuel going into the cylinders, the power output
improves with gain in altitude. Non-standard temperature however causes changes in density. In
colder than standard temperatures at a particular pressure altitude, density is increased and therefore
power is also greater. Thus, particular power outputs demand increased rpm in hot air and reduced
rpm in cold air.

Specific Fuel Consumption

16. If the power output of an engine were always constant, then its efficiency could be measured
purely by observing its gross fuel consumption (GFC), which is the fuel used per unit time. However,
295 Range and Endurance

when the power output varies considerably, as for example, when the weight, and hence the drag, of
an aircraft reduces in flight, then another term has to be found which relates the power output to the
fuel used. The term for this ratio is specific fuel consumption (SFC). With piston aero-engines it is the
ratio between power output and gross fuel consumption i.e. SFC = _GFC_ .
Power

17. One useful guide to engine efficiency is the percentage of total power which has to be used
within the engine instead of being employed in propelling the air. This percentage should also be kept
to the minimum. The SFC is measured in Kg fuel per hour per bhp and is best when at a minimum.
The factors which affect power output and their influence on the SFC are as follows:

(a) MAP and RPM. As stated in para 8, with piston engines, there is a choice of MAP
and rpm settings which can provide the same power output. High MAP and low rpm gives
minimum SFC because by using low rpm:

(i) Friction and inertia losses are low.


(ii) Generators, pumps, etc. absorb minimum useful power.
(iii) Supercharger friction and heat losses are low.
(iv) Volumetric efficiency is better.

(b) Altitude. As altitude increases upto full throttle height, the SFC improves (reduces)
for the three following reasons:

(i) Progressively less throttling of the engine is needed and the work done by
the supercharger is used more and more effectively.
(ii) Progressively less useful power output is wasted in exhausting the spent
mixture against the ambient air pressure.
(iii) The colder air contributes to a greater rise in temperature within the engine,
which is a measure of its thermal efficiency.

As altitude increases above the maximum weak mixture MAP fuIl throttle height, the power
required can only be achieved by increasing rpm or by putting the supercharger into a higher
gear. In either case, the SFC increases (deteriorates) because more useful power is
absorbed within the engine.

(c) Temperature. In cold air, at a particular altitude, the rpm may be reduced to
provide the power required because:

(i) The power available from the engine is greater.


(ii) The power required to propel the airframe in cold air is less.

The SFC will therefore be reduced in cold air because the power absorbed (wasted) at the
lower rpm is reduced and the thermal efficiency is increased.
FIS Book 1: Aerodynamics 296

(d) Carburettor and Intake. In order to prevent icing in the carburettor, heating of
the air intake may be used. However, the use of intake heating entails:

(i) A loss of power at a given MAP and rpm because of the reduced density and
hence weight of the charge drawn in.
(ii) A loss of the ram effect (which without intake heating, raises the intake air
pressure and supplements the action of the supercharger and thereby increases full
throttle height-approximately 1,000 ft at a TAS of 170 knots)
(iii) Enrichment of the mixture and a consequent increase in SFC. This increase
can be minimized if the carburettor is correctly compensated.

(e) Air Filter. Use of an air filter to eliminate the induction of dust into the engine
entails a loss of the (slight) increase in MAP due to ram effect. The consequent increase in
rpm required to maintain the power output therefore increases SFC.

Summary of Engine Considerations

18. A piston aero-engine is able to provide the most efficient power output required in flight under
the following conditions:

(a) Not more than maximum weak mixture MAP is set.


(b) Minimum rpm used.
(c) Supercharger in low gear.
(d) Aircraft at full throttle height.
(e) Carburettor air intake "COLD" and air filter "OUT".
(f) Power output is enhanced in cold air and engine efficiency (SFC) is improved.

AIRFRAME CONSIDERATIONS

Specific Air Range

19. An aircraft should generally be flown in such a way that a maximum weight of aircraft and
load is delivered over a maximum distance using minimum fuel. Towards this the Specific Air Range
(SAR) is defined as air distance travelled per unit quantity of fuel i.e. air nautical miles per kg (pound
or gallon) of fuel or air metres / Newton. Thus:

SAR = distance travelled


fuel used
Multiplying and dividing by "time",
SAR = distance travelled X __Time__
Time fuel used
Since Distance/ Time = TAS and from equation (18.1), GFC = Fuel Consumed / Time
297 Range and Endurance

SAR = TAS (18.4)


GFC

20. The fuel used by a piston engine depends upon the power output of the engine and the
efficiency with which it produces that power. From equation (18.3), since GFC = SFC X Power,
substituting it in equation (18.4):

SAR = _TAS _ X __1__ (18.5)


Power SFC

Therefore, at a particular weight, in still air, a piston engine aircraft should be flown for a maximum
product of airframe efficiency TAS / Power and engine efficiency 1 / SFC

TAS I Power Ratio

21. The power required to propel an aircraft in level flight depends upon speed, altitude and
weight. The optimum values of the TAS / Power obtainable will be considered as will the optimum
value of the ground speed / power ratio, taking into account head and tail winds. The following terms
will be used:

(a) Thrust. Thrust is the force exerted by the engines on the airframe to overcome
drag and is measured in Newtons.

(b) Work. Work is said to be done when any force displaces a body and the amount of
work done is the product of the force exerted and the distance through which the body is
moved. In the case of an aircraft engine:

Work done = Thrust X Distance. (18.6)

(c) Power. Power is the rate at which work is done. In the case of an aircraft engine:

Power required = Thrust X Distance = Thrust X TAS


Time
Since in level flight Thrust = Drag
Power required = Drag x TAS. (18.7)

22. Fig 18-2 shows a graph of TAS against power required which has been evolved from a TAS /
Drag curve by multiplying each value of drag by the appropriate TAS and converting to power
required. Note that the speed corresponding to the lowest point on the TAS / Power curve, known as
the minimum power speed (VIMP), is lower than the minimum drag speed (VIMD). Also the maximum
value of the TAS / power ratio is found at a point where the tangent from the origin touches the curve.
This speed coincides with the V IMD for the following reason:
FIS Book 1: Aerodynamics 298

Power required is a product of drag and TAS


(equation 18.7), thus

_TAS_ = ___TAS___ = __1__


Power Drag x TAS Drag

i.e. maximum TAS / power occurs at the


minimum drag or maximum 1 / Drag speed.

23. Effect of Altitude. An aircraft flying at


VIMD experiences a constant drag at any altitude
Fig 18-2: Variation of Drag and Power with
since the ratio of TAS / Power = 1 / Drag and Velocity (TAS)
the drag will remain constant for a constant IAS.
This ignores compressibility, which is unlikely to
affect piston engine aircraft airframe
performance. However, as altitude increases, the
TAS for a given IAS increases, but the power
required also increases by the same amount
(power required = Drag x TAS) Fig 18-3 shows
how power required increases with altitude, whilst
airframe efficiency remain unchanged.

Effect of Winds Fig 18-3: Effect of Altitude on Power


Required and Velocity (TAS)
24. Range should properly be an expression
of ground distance divided by fuel used. Thus, in
place of TAS / Power, ground speed / Power
should be studied. In Fig 18-4 TAS / Power
curve has been given three horizontal axes.

(a) In still air, a tangent from the


origin O indicates a minimum drag
speed V and the specific air range is
OV / AV.

(b) With a 75 kmph tail wind, the


ground speed / Power ratio at this speed
Fig 18-4: Effect of Wind on Range
V would be PV / AV. To obtain a
maximum ground speed / Power ratio in these conditions a slightly slower speed ‘W’ is flown,
where PB is the tangent from P and the curve, and PW / WB represents the maximum
possible efficiency.
299 Range and Endurance

(c) Similarly, with a 75 kmph headwind the ground speed / Power ratio is maximum at a
slightly higher speed ‘X’ where QC is the tangent from the false origin Q to the curve, and
QX / XC represents the maximum efficiency.

25. The effects of the head and tail winds on range speed are small and depend upon their
numerical relationship to the TAS. Unless a headwind exceeds 25% of TAS or tailwind 33% of TAS,
air speed adjustments are not made.

26. Recommended Range Speed (RRS). In practice, aircraft are flown for range at a speed
slightly faster than VIMD for two reasons.

(a) The variation of TAS / Power at speeds near VIMD is negligible and a slightly higher
speed will allow faster flights without undue loss of range.

(b) When flying at VIMD any turbulence or manoeuvres will cause a loss of lift and height.
This height can only be regained by the application of more power and consequent fuel
wastage. A higher speed ensures a margin for such events.

Piston engine aircraft are therefore normally flown at the recommended range speed, which is some
10% higher than VIMD.

27. Effect of Weight. As VIMD is proportional to the square root of all up weight so also is RRS.
The power required depends upon the drag
and the TAS. Drag depends upon the Lift /
Drag ratio at the angle of attack represented
by this airframe speed and the weight. The
TAS depends upon the value of the IAS,
altitude and temperature. It is possible to
predict the power required by an airframe
flown at RRS for a whole spectrum of weight
and altitudes. The graph at Fig 18-5 can be
used in conjunction with the power available
curves to decide the best altitude to fly at
Fig 18-5: Power Required at RRS at Various
various aircraft weights. Altitudes and Weights

Summary of Airframe Considerations

28. Piston engine airframe efficiency can be summarized as follows:

(a) By virtue of their propulsive nature, such airframes should be flown for a maximum
value of TAS / Power which is equal to 1 / Drag.
FIS Book 1: Aerodynamics 300

(b) In practice, such airframes should be flown at a speed 10% higher than VIMD i.e. the
recommended range speed.

(c) This speed should be increased slightly for headwinds in excess of 25% of TAS and
decreased slightly for tailwinds in excess of 33% of TAS.

(d) The altitude to fly is immaterial in so far as the airframe is concerned.

(e) As weight reduces, RRS reduces proportionally to the all-up weight. The power
required therefore, at this speed, is a function of weight, Lift / Drag ratio (a constant for a
given α) and TAS (which increases with altitude).

ENGINE AND AIRFRAME COMBINATION

29. For maximum specific air range an aircraft must achieve the largest TAS / GFC ratio. As GFC
is dependent upon engine efficiency and power developed, (from equation 18.5) for a piston engine
aircraft.

SAR = _TAS_ X __1__


Power SFC

30. Speed for Maximum Range. For a maximum value of TAS/Power it has been seen that, in
practice airframes should be flown at the recommended range speed, which is an IAS proportional to
the square root of the all-up weight, small adjustments being made for large head or tail winds. This is
the overriding consideration regarding range speed.

Summary

31. Piston engine aircraft cruise control procedures are as follows:

(a) Such aircraft are normally flown at a constant altitude dictated by their design and
operation of the supercharger, as there is very little advantage of relatively small change in
weight during flight.

(b) At a constant altitude these aircraft may be flown for maximum range:
(i) At a reducing IAS (RRS), but this is navigationally complicated and slow.
(ii) At a constant speed, which eases navigation and gives near maximum range
in slightly less time.
(iii) At constant power which results in an increasing air speed. This method is
most suitable for short distances and is the fastest of three.
301 Range and Endurance

DERIVATION FOR PISTON RANGE PARAMETER EXPRESSION

32. Consider an aircraft flying distance d in time t, in the process consumes M quantity of fuel.

By definition we know that SAR is defined as the air distance flown per unit quantity of fuel.

SAR = d Multiplying and dividing by t which is the time taken


M
SAR = d/t
M/t

If, time is unit time d / t = Speed or V = TAS and the total amount of fuel consumed in unit time =GFC

∴ SAR = __V__
GFC
for pistons GFC = SFC x Brake Power (equation 18.3)
∴ SAR = _______V________ (18.7)
SFC X Brake Power

But from basic definition, propeller efficiency (η) = Thrust Power


Brake Power
∴ Brake Power = Thrust Power / η

On substituting Brake Power in equation (18.7) we get

SAR = V = __ V X η__
SFC X TP/η SFC X TP

Also Thrust Power = T X V = D X V (∴ T = D in level flight)

Or ____V_____ = 1
Thrust Power D
∴ SAR = ___η____ (18.8)
SFC X D

In level flight L = W Therefore, dividing the numerator and denominator by L / W we get equation
(18.8) as

SAR = L X ___η___
W SFC X D
SAR = L X __1_ X 1 X η
D SFC W
SAR = CL X __1__ X 1 X η (18.9)
CD SFC W

33. Thus for maximum Range a piston engine aircraft is flown under conditions of maximum L / D
or maximum CL / CD with minimum weight, SFC and maximum propeller efficiency.
FIS Book 1: Aerodynamics 302

34. Total Range. For the total range of an aircraft, SAR has to be multiplied with the available
fuel. However, it should be noted that as fuel is consumed in level flight, weight also reduces which
causes an increase in range. In every large transport aircraft where fuel constitutes a very big
proportion of weight the effect of changes in weight is significant and the specific range expression
has to be integrated with respect to weight appropriately to correctly calculate range. Finally the effect
of wind will have to be taken into consideration to get the actual distance covered.

PISTON ENDURANCE

Introduction

35. To keep an aircraft airborne for as long as possible on the fuel available, the IAS should be
that at which the engine consumes least fuel. Endurance (measured in hours) is obtained by dividing
the fuel available by the gross fuel consumption (GFC). Since the GFC is assumed to be proportional
to the power output, maximum fuel economy is obtained by using the level-flight speed corresponding
to the lowest power output from the engine.

36. This speed is determined by the aerodynamic characteristics of the airframe, the specific fuel
consumption (SFC) of the engine at the power output required, the propeller efficiency and the all-up
weight (AUW). In practice, the speed may be modified by handling considerations. The actual
endurance, being a function of AUW, depends on the altitude, the outside air temperature and the
external condition of the aircraft itself. Wind has no effect on endurance.

Endurance Speed

37. By gradually reducing the power during level flight, a speed and power is eventually reached
at which the aircraft can just maintain a height at which there is no margin of control for handling
purposes. If the total drag and TAS are known, this minimum power may easily be calculated, since
Power = total Drag x TAS. Assuming the fuel consumption to be proportional to the power output, the
speed for minimum power gives maximum endurance.

38. It is interesting to note that at this low speed


the drag is greater than when flying at range speed,
which is higher, while the power output is less. This is
because the power depends on two factors (drag and
air speed), which alters when power is reduced, but
not in proportion. It can be seen from Fig 18-6 that an
appreciable reduction can be made in the speed for
minimum drag without incurring any large increase in
Fig 18-6: Total Drag vs. (EAS)
drag. The overall effect is a reduction in the product of
303 Range and Endurance

the two factors although drag has increased and speed decreased.

39. In practice, flight at the speed for minimum power is not practicable since all ability to
manoeuvre disappears and engine handling problems arise. For these reasons a margin of speed is
added to improve the manoeuvrability, resulting in a recommended endurance speed, which should
be maintained within plus or minus five knots. Endurance speed may be defined as the level flight
speed produced by the use of minimum power, plus a margin for control and manoeuvre. It is stated
in the Aircrew Manual for each type of piston engine aircraft.

40. From the power required curve for a


typical aircraft (Fig 18-7) the best endurance
speed would appear to be 200 kmph. The
recommended endurance speed would be some
20 Kmph above this speed.

41. The recommended IAS for endurance is


considerably higher than the minimum possible
flight speed, primarily because of the large rate of
drag increase at the lower speeds (Fig 18-7).
Flight at these speeds would entail the use of
higher power than that for maximum endurance,
and at full power, level flight is possible on some Fig 18-7: Deviation of Endurance Speed.
aircraft at speeds well below the power-off stalling
speed (Basic Stalling Speed).

42. Effect of Altitude on Endurance Speed.


The effect of increasing altitude on the curve in Fig
18-8 would be to move the power required curve
upwards and to the right, thus increasing the TAS
corresponding to minimum power, this increase in
the TAS is obtained automatically by maintaining
the same recommended IAS regardless of the
altitude. Therefore, at a given AUW the same IAS
should be used, irrespective of the altitude. Fig 18-8: Variation of Power Required with
Altitude

43. Effect of Weight on Endurance Speed. This is the only operational consideration, which
may require the use of a speed other than that recommended. The angle of attack at the speed for
minimum power is constant, irrespective of the weight. In consequence, the variation in lift with a
changing AUW is met by a corresponding variation in speed. Aircrew Manuals sometimes indicate the
speeds to be used for varies conditions of AUW but a useful rule for estimating the change in IAS
FIS Book 1: Aerodynamics 304

necessary to compensate for a change in AUW up to 20% is to vary the speed by half the percentage
variation in weight.

Factors Affecting Endurance

44. Variation in AUW. Since speed changes are obtained by varying the power, it follows that
less power is required for the reduced IAS associated with a lower weight and that fuel consumption
will be proportional to the weight of the aircraft. A 10% reduction in AUW increases endurance by
about 7%.

45. Effect of Altitude. In level flight, maximum endurance reduces with increasing altitude.
Assuming the fuel used per hour to be proportional to the power used, we know that Power Used =
Total Drag X TAS. For a given IAS the drag is same at all altitudes, but the corresponding TAS
increases with height. It follows that the power required for endurance speed increases with altitude,
as does the fuel consumption and endurance is therefore reduced. Although maximum endurance is
obtained at sea level, a safety limit is imposed in practice. If flying in cloud, the local safety height
must be observed and in clear conditions a height of a least 1000 ft above ground is recommended.

46. Effect of Additional Drag. Since the recommended speed must be maintained to provide
adequate control and manoeuvrability, more power must be used to counteract extra drag arising from
the carriage of external stores. Endurance will be improved by any measures taken to ensure that the
aircraft is aerodynamically clean, however, with a large increase in drag, the best IAS is obtainable
only at a higher power setting, thus reducing the endurance.

47. Effect of Temperature Variation. At high atmospheric temperatures the air density, the
weight of the induction charge, and therefore the power output, are all reduced. To maintain a given
lAS, an increase in MAP and / or rpm is necessary, and there may be a considerable loss in
endurance.

Flying for Maximum Endurance

48. Maximum fuel economy is obtained only if full consideration is given to the following factors:

(a) MAP and RPM. While the ideal combination for fuel economy is high MAP and low
rpm, it may be impossible to maintain the correct endurance speed without reducing the
throttle opening. In very low atmospheric temperatures, sustained cruising at low power may
give rise to engine temperature problems. The pilot should make the choice of settings in the
light of power plant circumstances.

(b) Mixture Strength. Wherever possible, weak mixture should be used as the excess
fuel in rich mixture is expended only in reducing combustion temperatures. Cruising in high
305 Range and Endurance

atmospheric temperature may cause overheating in weak mixture, which may be alleviated by
the periodic use of rich mixture at appropriate intervals.

(c) Supercharger Gear. At the low altitudes at which the aircraft is flown for
endurance, low gear is invariably in use. If it becomes necessary to fly at high altitude, the
best fuel economy is obtained by using full throttle for the MAP required, with rpm reduced to
give the required speed. Low gear usually provides enough power even at the highest
altitudes.

(d) Carburettor Air Intake Shutter. Unless in icing conditions, cold air should be used
since preheating of the intake gases results in reduced power output and impaired fuel
economy. The degree to which endurance is affected depends on the type of carburettor and
instructions on the use of the intake shutter are included in Aircrew Manuals.

(e) Air Filter. Better fuel economy results if the air filter is out of operation in flight,
since least restriction is caused to the airflow into the engine. However, when at the low
altitude selected for endurance, sustained cruising with the filter out of operation may cause
excessive engine wear and it may be advisable to bring the filter in. This applies especially in
tropical countries where dust may extend to some height into the atmosphere.

49. Technique for Endurance Flying. The technique of flying for endurance may be
summarized as:

(a) Fly at the lowest, safe, practicable altitude.


(b) Adjust the throttle and rpm to give the recommended speed for the AUW and
maintain this speed within five knots.
(c) Use weak mixture, carburettor cold air and air filter out of operation.
(d) Trim the aircraft correctly and reduce all possible drag such as that caused by
hatches, flaps radiator shutters, etc.
(e) Jettison any excess load possible if
maximum endurance is required.

50. Comparison of Range and Endurance Speeds.


Since endurance implies minimum power and range
implies minimum drag, comparison of the two speeds is
not easy. However, the endurance speed may be
identified by referring to Fig 18-9, which plots power
required against TAS. The speed requiring minimum
powder is found at the base of the power curve. The range
speed for this aircraft would be that giving the highest ratio Fig 18-9: Relationship between
Range and Endurance speed
FIS Book 1: Aerodynamics 306

of speed / Power and may be found by drawing a tangent from the origin of the graph to the power
curve. The range speed occurs at the point of tangency. Range and endurance speeds differ for each
type of aircraft, but in general the endurance speed is about 80% of the range speed.

Derivation for Piston Endurance Expression

51. It is known that whereas range flying is concerned with the distance flown on a given fuel,
endurance flying means flying for as long as possible on the fuel available. In other words it involves
the factor of max time on a given fuel. To be precise:

Specific Endurance = __Flight time__ = _t_


Quantity of fuel M

On dividing the numerator and denominator by time we get,

Specific Endurance = __Flight time / time = t/t


Quantity of fuel / time M/t
Specific Endurance = _____________1____________ = __1_
Quantity of fuel used per unit time GFC
Therefore, Specific Endurance = __1__ i.e. reciprocal of fuel flow.
GFC

For maximum Specific Endurance 1 / GFC is to be maximum, GFC is to be minimum.

1 / GFC maximum implies 1 / (SFC X BP) maximum ( from equation 18.3 )

We also know that Propeller efficiency (η) = TP or BP = TP


BP η
On substituting the value of BP we get

Specific Endurance = ___1____ = _1__ X _η_


SFC x TP/η SFC TP
Also TP = DXV

On substituting we get

Specific Endurance = _1_ X η X __1__ (18.10)


SFC DxV

Since L = W & T = D in level flight and using the basic equation of L =W = CL½ ρv2S &
D = T = CD ½ ρv2S

½
Therefore D = CD X W And V = _2W_
CL CL ρS
307 Range and Endurance

We can rewrite equation (18.10) as

½
Specific Endurance = __1__ X η X CL X 1 X CL ρ S
SFC CD W 2W

½
Or Specific endurance = __1__ X η X CL 3/2 X _1_ X ρS (18.11)
SFC CD W3/2 2

52. Thus for Maximum Specific Endurance, the aircraft must be presented at an AOA that gives
the best value of CL3/2 / CD the speed corresponding to this is of Minimum power required. The SFC
must be minimum, ac must be as light as possible and flown at the lowest practicable height to take
advantage of maximum density.

53. For the total endurance of an aircraft Specific Endurance has to be multiplied with the
available fuel. However it should be noted that as fuel is consumed in level flight, weight also reduces
which causes an increase in Specific Endurance. In very large transport aircraft a very big proportion
of weight is made up of fuel and the specific endurance expression has to be integrated with respect
to weight appropriately to calculate endurance.

JET RANGE

54. To obtain the maximum range from turbo-jet aircraft they must be operated under known and
precise conditions of altitude, speed and weight. Any deviation from these conditions can only result in
performance penalties. The operating techniques for turbo-jet aircraft differ from, and are more
exacting than those used for piston-engine aircraft. Experience of operating piston-engine aircraft for
range is of little value in the operation of turbo-jets. Since range flying is of paramount importance, it
is essential that the basic theory and methods of flight planning and cruise control are understood.

55. From equation (18.2) we know that SAR = TAS / GFC.

Since gross fuel consumption (GFC) is the product of specific fuel consumption (SFC) and
thrust, the SAR can be expressed as:

SAR = TAS / (SFC X Thrust)

In level flight, thrust equals drag and therefore the above expression for jets can be equally
well expressed as:

SAR = ___TAS____ (18.12)


SFC X Drag
FIS Book 1: Aerodynamics 308

Since at high altitudes the turbo-jet is working at high rpm, the SFC can be regarded, for all practical
purposes, as constant. SAR therefore varies according to the variation in the TAS / Drag ratio and this
is the principal factor that governs the range. If the TAS / Drag ratio is high then greatest range is
achieved and vice versa.

Basic Techniques

56. Irrespective of the technique used, it is always necessary to fly at a certain lAS. The actual
lAS used depends on which technique is in use. From equation (18.12), it has been shown that the
range of a turbo-jet aircraft depends on the TAS / Drag ratio. To keep the ratio at its maximum the
aircraft must be flown at the optimum angle of attack during the whole cruising period. Since the AUW
reduces during flight, the lift must be adjusted to balance the reduced weight. This can only be done
by reducing the lAS, if the angle of attack is to remain at the best setting.

57. There are only two principal techniques for cruising for maximum range:

(a) Constant altitude. Reducing the lAS as the weight reduces, maintaining a constant
altitude.
(b) Climbing cruise. Gradually climbing as the weight reduces and reducing lAS in this
way.

These two techniques must be considered in detail, ignoring initially, compressibility effects.

Constant Altitude Technique

58. Consider a typical level flight drag curve


plotted against TAS for a given AUW and altitude.
The TAS / Drag ratio is at a maximum at that speed
at which the tangent from the origin just touches the
drag curve. At any other point the proportion of drag
is larger and the ratio falls, even though the drag is
least at the lowest point of the curve. The ratio of the
corresponding TAS to the minimum drag is less than
that found at the point of tangency from the origin (Fig
18-10).

59. It can be shown mathematically that this Fig 18-10: Variation of Total Drag with
Speed
speed is about 1.32 times the TAS corresponding to
the lowest point of the curve (refer page 93 and 421 of the book “Aerodynamics” by LJ Clancy). Flight
at this speed means that a particular angle of attack and hence lift coefficient must be used. Thus for
a given constant weight the same lAS would apply at all altitudes.
309 Range and Endurance

60. So far the speed problem is fairly straightforward but the first difficulty arises when the
question of variation in weight is considered. Apart from war load, fuel alone may account for 50 per
cent of the AUW at take-off. Obviously advantage can be taken of the reduction in weight as this fuel
is used. To obtain the best TAS / Drag ratio, it is necessary to maintain the same angle of attack.
Therefore to reduce the lift necessary for the decreasing weight and keep the angle of attack
constant, a low lAS must be set and this is done by reducing rpm (thrust). Lower IAS results in a lower
drag and since thrust equals drag in level flight, the overall effect is an improvement in the overall
TAS / Drag ratio and thus range performance. (Even though the TAS is reduced because of the lower
lAS, at the reduced power, the drag falls faster since it varies as the square of the IAS).

Climbing Cruise Technique

61. Although a good range performance is obtained by using the method described above, this is
not the best result. Maximum range will result when the drag is reduced without decreasing the TAS.
The TAS can be kept constant at a reduced IAS onIy by an increase in altitude. Therefore the best
result is obtained if the aircraft is allowed to climb slowly at constant rpm as the weight reduces so
that the TAS remains constant while at the same time the IAS reduces in proportion to the increased
altitude. Also since VIMD is proportional to t he square root of aircraft weight, to ensure maximum
performance as weight reduces, altitude must be changed i.e. ___Weight____ ratio must remain
constant. relative pressure

62. However, when using the climbing cruise technique, a small difference arises. Because of the
slowly changing altitude the best TAS is no longer 1.32 times VIMD, but is 1.19 times the VIMD. For
simplicity the mathematics of this change is omitted (described on page 455 and 456 of
“Aerodynamics” by LJ Clancy). Using the climbing cruise technique at this slightly lower speed, gains
in range of 15 to 20 percent are obtained when compared with the constant altitude technique. On a
long flight such an advantage cannot be foregone. The constant altitude technique is therefore
suitable only for shorter flights.

Effect of Compressibility

63. Compressibility effects govern the final speed selected, because they have an important
effect on the TAS / Drag ratio. Above the critical Mach number (Mcrit) compressibility effects increase
rapidly to cause a fall in the TAS / Drag ratio (because of the higher drag at the same TAS) and
therefore a limit is set on the indicated Mach no (IMN) for economical cruising (Fig 18-10).

64. Consider an aircraft operating in the stratosphere where the temperature is theoretically
constant. At the constant temperature, the speed of sound is therefore constant and a fixed IMN
implies a constant TAS. Therefore a cruising IMN limit (refer para 63) means that the cruising TAS
has a limit. Since now the maximum TAS has a limit, the highest possible value of TAS /Drag (and
hence the range of a particular aircraft) has a limit, which is realized when the drag is least. Under
FIS Book 1: Aerodynamics 310

stratospheric conditions, therefore the TAS for cruising is that for minimum drag, which gives the
absolute maximum range provided that the aircraft can be handled without undue fatigue at this low
lAS for long periods at altitudes near the ceiling. It is assumed that sufficient thrust is available at
these altitudes to maintain the min drag speed.

65. Since Mcrit depends upon the angle of attack the overall picture is now further complicated.
Therefore at high angles of attack the compressibility drag rise starts at a lower IMN than at smaller
angles of attack, i.e. Mcrit is lower. While flying at the speed for minimum drag, a certain angle of
attack will be used, which has a certain critical mach number. If the IAS is now increased slightly, a
lower angle of attack will be needed. This lower angle of attack will have a higher MCRIT. In this way a
higher TAS is now achieved before the compressibility drag rise commences. The higher TAS is more
significant than the slight rise in drag due to the higher IAS and, provided that the increase in speed
from that of minimum drag is kept small, an improved TAS / Drag ratio occurs.

66. In practice, therefore, the best speed under stratospheric conditions is not exactly the speed
for minimum drag but a some what higher one, the exact figure depending entirely on the
aerodynamics of the aircraft concerned. In the remainder of this chapter the variation between types
will, for simplicity, be ignored but should be borne in mind when the minimum drag speed is
mentioned. For a full understanding of this technique it is necessary to look more closely into the
effects of altitude.

Effect of Altitude

67. First, the influence of altitude on the TAS / Drag ratio must be considered when the aircraft is
at a given weight in level flight, at a constant angle of attack at various altitudes. With angle of attack
and weight constant, the IAS and drag are unchanged but the same IAS gives an increasing TAS as
altitude increases.

68. If there were no compressibility effects, as altitude was increased at a constant IAS the drag
at 40,000 feet would still be same, but the TAS would be about double the IAS. Thus the TAS / Drag
ratio would have doubled and would continue to increase with altitude. (This principle emphasizes the
importance of high altitude in obtaining a high SAR from the aerodynamic aspect alone. No matter-
how low i.e. favourable the SFC may be at sea level, the very low TAS / Drag ratio at this height
absolutely precludes a high SAR).

69. Because of compressibility effects, the rate of gain in the TAS / Drag ratio is not sustained. At
a constant IAS a higher altitude means an increase in IMN. As soon as the IMN reaches Mcrit any
further increase in altitude at the same IAS results in Mcrit being exceeded. The drag immediately
starts rising at a greater rate and the TAS / Drag ratio decays rapidly. To understand these points
more clearly, refer to Fig 18-11 which shows three typical curves of drag versus TAS in level flight for
311 Range and Endurance

an aircraft at a given weight flying at sea level,


30,000 feet, and 45,000 feet. Also shown is the
0.8 M line for the standard temperature at these
heights which is the Mcrit for the aircraft under
consideration. Note that an increase in altitude
causes the drag curve to shift right, the minimum
drag remaining same at different altitudes since
the α and the IAS is constant but the TAS is
changing with altitude.

70. As before, the angle, which the tangent to


each curve makes with the horizontal, is a
measure of the TAS / Drag ratio. The smaller the
angle, the higher the ratio and the greater the
SAR. Note first that an increase in altitude causes
the angle to reduce. It can also be seen that the
best speed to fly at any particular height is 1.32 X Fig 18-11: Effect of Altitude on Drag.
the minimum drag speed (IAS or TAS) until the
drag curves have moved so far to the right that compressibility causes the drag curve to rise steeply.
For this aircraft this occurs at 45,000 feet and owing to the steep drag rise the best speed is now only
1.1 times the minimum drag speed. As the drag curve moves still further to the right with increase of
altitude, towards the limit imposed by Mcrit, the point of tangency approaches more closely the
minimum drag speed and the angle becomes still smaller.

71. The considerations of paras, 65 to 68 apply to cruising at 1.32 or 1.19 times the minimum
drag speed, the only difference being in the altitude at which the TAS/Drag ratio starts to decay. Since
1.19 is a lower speed (IAS) than 1.32, then at the lower IAS Mcrit will not be reached until a higher
altitude, therefore the TAS/Drag ratio continues to increase to this higher altitude.

72. Consider now two identical aircraft, A and B, flying at these speeds under optimum
conditions. A is flying at 1.32 X minimum drag speed at a constant altitude such that the IMN is Mcrit. B
is flying more slowly at 1.19 times the minimum drag speed on a climbing cruise but a higher altitude,
so that the IMN is also Mcrit in the stratosphere. In such a case both aircraft have almost the same
TAS but B, flying at the lower speed, will have about 9 percent less drag. Thus the TAS / Drag ratio
and SAR of B is about 9 per cent greater. As time passes and A stays at 1.32 x minimum drag speed
by reducing thrust as the weight reduces (para 57), the TAS decreases and the mach number falls
below Mcrit. This decreases the TAS / Drag ratio by an amount which increases with time of flight.
Thus the SAR of B continues to exceed that of A by a factor which increases with time.
FIS Book 1: Aerodynamics 312

73. It has been shown in para 64 that the best TAS / Drag ratio is obtained by flying at the
minimum drag speed. It can now be seen that this is true only if the aircraft has the power to fly at a
higher altitude than A or B. Consider C flying at minimum drag speed and therefore higher than B at
an IMN equal to Mcrit. All three aircraft have almost the same TAS (limited by Mcrit) but C has the
lowest possible drag, being about 8 per cent less than that of B and about 17 percent less than that of
A at the start of the cruise. The SAR is increased in the same proportion. This benefit is obtained only
by flying at the minimum drag speed at the higher altitude.

74. If B slowed down to the minimum drag speed at its own altitude, the TAS / Drag ratio would
decrease. By maintaining the IMN (and TAS) and climbing to the same altitude as C, the IAS would
then automatically fall to that for minimum drag. In this condition the TAS is at its maximum (at the
Mcrit limit) and the drag the least. This then is the absolute maximum TAS / Drag ratio obtainable from
the aircraft and is the condition for maximum range.

75. Little m. The ratio of the cruising IAS to the minimum drag IAS has been given the symbol
m. When speaking of this ratio 'little m' is generally used, to avoid confusion with Mach Number (M
No.). If m = 2 the cruising IAS is twice the minimum drag IAS and if m = 1.32 the cruising IAS would
be 1.32 times the minimum drag IAS and so on.

76. The ideal conditions for range are therefore obtained when m=1 at such an altitude that the
IMN is Mcrit. In practice some departure from the ideal conditions is often necessary for the following
main reasons:

(a) The aircraft may be tiring to fly, from the handling viewpoint, for long periods at the
low IAS.
(b) There may not be sufficient thrust available at this altitude to give the required speed.

77. If the ideal conditions cannot be


achieved, the best result is obtained by
flying at the highest possible altitude that
can be reached, keeping the IMN at Mcrit.
Here, because the altitude is lower, the
corresponding IAS is greater than that for
minimum drag, i.e. m is greater than 1. In
fact the lower the altitude at which it is
necessary to cruise, still at Mcrit, the
higher is the lAS, the greater does m
become until it reaches 1.19 if a climbing
cruise is in use, or 1.32 if the constant
Fig 18-12: Comparison of the Two Climbing
altitude cruise has been chosen. These Cruises
conditions are shown in Fig 18-12 and 18-13.
313 Range and Endurance

78. If it were necessary to use a climbing cruise


at an altitude lower than B the speed would be m=
1.19, and the lower the altitude the more would the
IMN fall below Mcrit. If for some reason the climbing
cruise was not used, the speed required for a
constant altitude flight would be m=1.32 up to that
altitude where the IAS corresponds to Mcrit at the
start of the flight, i.e. the condition of A in Fig 18-13.
For level flight at a higher altitude the speed would
still be Mcrit to begin with and m would reduce
progressively if the level flight altitude was
increased until a speed of m=1 was reached at the
maximum altitude. Fig 18-13: Comparison of Ranges

Effect of Weight

79. It has been shown (paras 57 to 61) how important it is to exploit the reduction in weight that
occurs during the flight and that this is best done by using the climbing cruise technique (constant
IMN, angle of attack and m). Thus during the time that the weight is reducing the aircraft gains height
and the IAS reduces.

80. The weight, altitude, and IAS are all closely related during a climbing cruise. Fig 18-13
illustrates such a relationship for a cruise at m= 1 and shows the increase in altitude (and reduction in
IAS) with decreasing weight.

81. Temperature Considerations. At a given pressure altitude a change in temperature has no


effect on the IAS and angle of attack required for the flight condition. Since the IAS and angle of
attack are unaltered the drag remains the same. The only difference is that the TAS changes because
of the effect of temperature on the air density. However, Mcrit is unaffected because the temperature
change also changes the speed of sound in the same ratio as it changes the TAS. Aerodynamically,
therefore, the only change caused, e.g. by a rise in air temperature, is to increase the TAS, giving a
small gain in the TAS / Drag ratio. A reduction in temperature causes a reduction in TAS and TAS /
Drag ratio.

Summary of Aerodynamic Considerations

82. From the discussion above, certain conditions are required if a high SAR is to be achieved
from the aerodynamic point of view. They are: -

(a) A high TAS/Drag ratio.


(b) IAS as close to minimum drag IAS as acceptable.
FIS Book 1: Aerodynamics 314

(c) TAS at Mcrit (never above).


(d) Reduction in weight must be turned to advantage (climbing cruise adopted).
(e) High altitude.

JET RANGE ENGINE CONSIDERATIONS

83. To get the best results, the engine factors too have to be considered. By combining the
aerodynamic & engine factors the overall range flying technique is evolved. A high SAR (air distance
per unit of fuel) is equally important for range to both turbo-jet and piston-engine aircraft. However, the
turbo-jet is much more sensitive to flight conditions than the piston engine and to understand the
changes in the cruising technique, which are necessary for turbo-jet aircraft, it is necessary to
examine the variable factors that affect the turbo- jet.

Thrust

84. The thrust (N) of a turbo-jet engine depends directly on two factors:

(a) The mass flow (Kg) of air through the engine per unit time.
(b) The rearward acceleration imparted to this mass on passing through the engine and
subsequently expanding in the atmosphere.

85. The two factors are governed by: -

(a) Entry speed of air into the intake.


(b) Atmospheric Temperature (Absolute).
(c) Engine size (capacity of compressor).
(d) Atmospheric pressure.
(e) Engine (compressor) rpm.

86. Gross Fuel Consumption. The lower graph of Fig l8-14 shows the variation of GFC (Kg
fuel / hour) against altitude and rpm for a typical engine in the standard atmosphere. The upper graph
of Fig l8-14, which plots thrust against altitude and rpm, can be seen to follow the same trend as the
lower graph. Therefore fuel, consumption can be said to be proportional to thrust.

87. The Effect of Temperature on GFC. At any given pressure altitude, a constant thrust can
only be maintained in conditions of changing temperature, by correcting the TAS and rpm. A
reduction in temperature causes the thrust to increase and hence the rpm should be reduced. The
ratio of rpm (N) to the square root of relative temperature (t) should be constant i.e. N/√ t is constant.
Later, when discussing the aircraft as a whole, it will be seen that temperature changes have no effect
on the SAR. However, temperature does influence the take-off and climb. An increase in air
temperature at constant rpm reduces the thrust and take-off rpm During the climb, if rpm is increased
315 Range and Endurance

to maintain the thrust the GFC is raised and if


the rpm is kept constant, the reduced thrust
increases the time to height. In both instances
more fuel is used to climb to any particular
height.

Specific Fuel Consumption

88. Specific fuel consumption is defined


as the amount of fuel used per unit time to
produce one unit of thrust. The higher the
thermal efficiency of the engine, the lower the
SFC at a given TAS. The biggest single factor
affecting the SFC is the rpm. Fig 18-15 shows
sea level static curves for an axial flow
compressor engine. It can be seen that the
SFC reduces (improves) rapidly with
Fig 18-14: Variation of Thrust and GFC with
increased rpm until a point is reached where it
Altitude at Various RPM.
becomes a minimum. This curve also shows
that the SFC curve is quite flat over a fairly
wide band of cruising rpm. Provided that the
engine is operated within this band of engine
speeds the SFC can be taken to be virtually
constant.

89. Although rpm has the largest single


effect on SFC, other factors affect the final
figure. At a constant TAS and rpm a decrease
in temperature causes the SFC to reduce
owing to the improved thermal efficiency, Fig 18-15: Static Performance
since the engine is working between wider temperature limits. Thus under these conditions the SFC
decreases up to the tropopause and thereafter stays constant at higher altitudes. At constant rpm and
altitude, any variation in TAS has little effect on the thrust, but the engine BPC is sensitive to the
increased compressor intake pressure and causes an increase in the fuel flow. The increase in fuel
consumption with no increase in thrust causes a higher SFC.

Summary of Effects

90. Changes in TAS have little effect on the thrust of a turbo-jet engine (the reasons for this are
not within the scope of this chapter). However the thrust horse-power (given by thrust X TAS) will of
FIS Book 1: Aerodynamics 316

course increase. The performance of the turbo-jet is affected by three main variables:

(a) RPM. Increased rpm:


(i) Increases thrust at constant altitude and temperature.
(ii) Reduces SFC rapidly at first but more slowly at the higher end of the rpm
range. The highest engine speeds increase SFC slightly.
(iii) Increases GFC.

(b) Altitude. Increased altitude:


(i) Reduces thrust at constant rpm.
(ii) Reduces GFC because of (i).
(iii) Reduces SFC up to the tropopause.

(c) Temperature. Increased temperature:


(i) Reduces thrust at given rpm and altitude.
(ii) Increases SFC.

91. When flying for range the best engine cruising conditions are:

(a) High rpm for lowest SFC.


(b) High altitude, so that:
(i) Thrust output at high rpm is just sufficient to obtain the desired speed and so
reduce the GFC at this rpm.
(ii) SFC is kept low.

92. The Aircraft as a Whole. To obtain an estimate of the range performance of a complete
aircraft, the engine and airframe characteristics must be combined. In general at high altitudes, the
aerodynamic considerations are the most important, the engines being simply slaves to propel the
aircraft at its most efficient speed.

Speed for Range

93. As far as the range speed at high altitude is concerned, the engine characteristics, with one
exception have no effect on the airframe considerations. The exception is the instance when the
engines cannot provide enough thrust to realize the optimum conditions of m and Mcrit. In this case the
cruising altitude must be reduced, thus raising the value of m. However this state of affairs is unlikely
to occur on long-range turbo-jet aircraft, since they are designed with this feature in mind so that just
sufficient power is available to realize the optimum speeds at the designed cruising altitude.

94. However development has resulted in the selected engines giving a higher thrust than that
envisaged, so that the reserve of power is available at the designed operating altitude. The higher
317 Range and Endurance

thrust gives no improvement in cruising altitude or speed because Mcrit and the speed for minimum
drag are dictated by purely aerodynamic reasons, and are closely related to a narrow band of
altitudes, departure from which brings a reduction of performance. The power reserve does give one
advantage in that it can be used to ensure that optimum conditions are always available despite large
variations in temperature, although if too much reserve is in hand, the more severely throttled
engines, operating at too low rpm, would realize a poor SFC and have an adverse effect on range.

Altitude

95. Climbing Cruise in Tropospheric Conditions. In the troposphere, owing to the


temperature lapse rate, the TAS decreases as altitude is gained at constant IMN and therefore the
drag is decreasing faster in proportion to the air density. At constant rpm the thrust is reducing less
rapidly than in proportion to the density since the temperature is reducing as height is increased.
Therefore in the troposphere some adjustment of the rpm may be necessary. Since the time spent at
these altitudes should not normally be long, these adjustments are not large. Since practically the
whole cruise is done in the stratosphere the cruising technique is simple. The pilot merely flies at the
correct IMN and checks that the rpm is constant at the appropriate figure. A navigational advantage of
this technique is that the TAS remains the same while in the stratosphere at a constant IMN.

96. Climbing Cruise in Stratospheric


Conditions. During a climbing cruise in
the stratosphere at constant Mcrit and
constant m, the drag is decreasing in
proportion to the reduction in density. At
constant rpm the thrust is decreasing in the
same proportion. Thus having set the
correct rpm (so that the thrust equals drag)
all that needs to be done is to check that
the rpm remains constant so that as
altitude is gained the thrust continues to
equal the drag (Fig 18-16).

97. Temperature. It was stated that,


for a change in temperature, the rpm has to
be corrected to give the same thrust and
Fig. 18-16: Effect of Altitude on Engine
the TAS decreases in a climb at constant Performance
IMN. For a rise in temperature, higher rpm is necessary for a given thrust and the fuel consumption
rises. Provided that the thrust required can be obtained within the rpm limitations, the increased
consumption is balanced by the airframe advantage gained from the higher temperature and thus the
SAR is unchanged.
FIS Book 1: Aerodynamics 318

Effect of Wind

98. The proceeding paragraphs have considered range flying in still air and it is important to note
that whenever range is mentioned in Aircrew Manuals, it invariably refers to still-air range. A head
wind will have a detrimental effect, conversely a tail wind will increase specific range. It is therefore
sound practice to examine winds at various altitudes while flight planning to determine the best height
to fly.

99. The basic requirements for maximum range in turbojet aircraft have been shown to be
minimum SFC and the best TAS / Drag ratio. Broadly speaking, to fulfil these requirements and
achieve the best still-air range, jet aircraft normally fly as high and as fast as possible, the range
speed being governed by that at which compressibility effects start to degrade the TAS / Drag ratio.
However, in some low performance jet aircraft there is only sufficient power to fly at low speeds at
medium and high altitudes and compressibility effects are of no consequence. In such aircraft it is
possible to increase speed by an appreciable amount without any significant decrease in still-air
range, and therefore in a strong head wind, an increase in speed will result in a reduction of overall
fuel consumption against a given ground distance flown. If a strong head wind component is
encountered in any jet aircraft the best solution lies in seeking a flight level where the effect of the
head wind is minimized, bearing in mind the other factors affecting range. In high performance
aircraft, increasing speed into a head wind is unlikely to be of benefit since the TAS / Drag ratio will be
adversely affected by compressibility effects.

100. It has been shown that for aerodynamic reasons alone, increased altitude brings increased
SAR and at the same time engine efficiency is improved from the higher rpm needed. The lower
temperature encountered also helps to increase engine efficiency with a negligible loss in airframe
performance.

Derivation for Jet Range Parameter Expression

101. Consider an aircraft flying distance d in time t and in the process consumes M quantity of fuel.
Now by definition we know that Specific Air Range or SAR is defined as the air distance flown per unit
quantity of fuel.

SAR = _d_ (18.13)


M

If t is the time taken we get on multiplying and dividing equation (18.13) by t

SAR = d/t
M/t
If time taken is unit time then
d/t = speed or V = True Air Speed (TAS)
319 Range and Endurance

and total amount of fuel consumed in unit time = Gross Fuel Consumption (GFC)

Therefore, SAR = _TAS_ (18.14)


GFC

But for jets GFC = SFC X Thrust from equation (18.2)


In level flight Thrust = Drag

Therefore (18.14) becomes SAR = ___TAS____


SFC X Drag
Or SAR = TAS X _1_ (18.15)
Drag SFC

102. Thus to get maximum range, which implies maximum SAR, aircraft should be flown at a
speed such that TAS / Drag is maximum and SFC is minimum. Best TAS / Drag is obtained from the
Total drag curve. The AOA which is maintained consequently corresponding to the maximum value of
√CL /CD as can be seen from below.

In level flight L = W & T = D and using the basic equation of


T = D = CD½ ρv2S
L = W = CL½ ρv2S

Dividing Drag by Weight

we get D = CD x W
CL
Also we can see that V = ___√ W____
(CL½ ρv2S) ½

On substituting the values of D and V in (18.15)

½ ½ ½ ½
SAR = W X 1 X 1 X 1 X CL X 1 X 1
CL (1/2) ρ S CD W SFC

½ ½ ½
SAR = √ CL X 2 X 1 X 1 X 1 (18.16)
CD ρ S W SFC

Equation (18.16) shows the various factors on which SAR is dependent upon.
103. For the total range of an aircraft, SAR has to be multiplied with the available fuel. However, it
should be noted that as fuel is consumed in level flight, weight also reduces which causes an increase
in Specific Air Range. In very large transport aircraft, where a very big proportion of weight is that of
fuel, the effect of change in weight is significant and the range expression has to be integrated with
respect to weight appropriately to correctly calculate range. The effects of wind-encountered will have
FIS Book 1: Aerodynamics 320

to be taken into account to know the distance covered on ground.

104. This paragraph presents the most simple, but at the same time, most efficient means of flying
turbo jet aircraft for maximum range in various roles. The most commonly used procedures are
studied under various headings, each one being more efficient than its predecessor.

(a) Low Altitude. At low level, the efficiency of the engine (rpm within the optimum
band) outweighs the efficiency of the airframe (1.32 x VIMD). Normally a high IAS,
approximately 2.0 x VIMD, is flown using rpm which are reduced as weight decreases. Range
is about 30 - 40 % of maximum.

(b) Medium Altitude. At medium altitude, a reducing IAS (1.32 x VIMD) is flown using
reducing rpm contained within the lower half of the optimum band. Range is about 50-70% of
maximum.

(c) High Altitude. Maximum ranges are achieved at high altitudes. The methods
adopted could be:

(i) Constant Altitude. To take maximum advantage of decreasing weight, an


optimum Mach number (set by compressibility) is initially flown at an altitude where
near maximum cruising rpm are required. As weight reduces, so the speed and rpm
are reduced to maintain a constant angle of attack. The advantage to the TAS / Drag
ratio is that although the TAS is reduced, because the angle of attack is being
maintained at a constant, drag is reducing as the square of the speed, thus resulting
in an overall increase in the TAS / Drag ratio. The required decrease in rpm will
have little effect on the 1 / SFC ratio, as the engine will still be operating within the
optimum rpm range. However, on long flights this method does lead to problems in
flight planning because of the changing speed, and so an alternative and more
practical method is to maintain the originally selected Mach number by slightly
reducing rpm as weight decreases. An improvement in the TAS / Drag ratio is thus
obtained as TAS is now constant, and drag is reducing as weight decreases. Range
is about 80-90% of maximum.

(ii) Stepped Climb. An optimum Mach number is flown at an altitude


approximately 1,000 ft above the ideal starting altitude set by the optimum weight /
relative pressure value. As the fuel is consumed and the aircraft weight reduces, the
theoretical ideal cruise climb altitude starts to increase, but in this method, the level
cruise is flown by reducing rpm until the aircraft is 3,000 ft below the ideal cruise
climb altitude. A climb is then made to the next semi-circular height 1,000ft above the
ideal altitude, and so on. Range is about 95 % of maximum.
321 Range and Endurance

(iii) Cruise Climb. An optimum Mach number is flown starting at an altitude


determined by the optimum weight / relative pressure, using an rpm setting
determined by the optimum N / √ t value. Range is maximum. However, in very hot
conditions, a lower weight / relative pressure value (lower altitude) may have to be
accepted with a lower N / √ t value (lower rpm) to avoid exceeding maximum cruising
rpm.

(iv) Fast Cruise. Turbojet aircraft should never be flown at slower speeds than
those outlined above except when endurance is required. Frequently, faster speeds
are flown in the interests
of time economy at the
expense of fuel
economy. The
performance penalties
are obviously
progressive.

105. Summary. In Fig 18-17 the


profile of each type of cruise procedure
is shown with an indication of the range
achieved. Fig 18-17: Profiles of Turbojet Aircraft Procedures

FLYING FOR ENDURANCE: TURBO-JET AIRCRAFT

Introduction

106. Flying for endurance implies flying in conditions which realize the minimum fuel flow and so
the longest possible time in the air on the fuel available. The need for maximum endurance arises
less frequently than that for maximum range, but circumstances can arise which require aircraft to
loiter or stand-off for varying periods of time during which fuel must be conserved. Whereas range
flying is more closely concerned with specific fuel consumption (SFC) and air nautical miles per
kilogram of fuel, endurance flying is concerned with the gross fuel consumption (GFC), i.e. the weight
of fuel consumed per hour.

Principle

107. Broadly, since fuel flow is proportional to thrust, fuel flow is least when thrust is least.
Therefore maximum level flight endurance is obtained when the aircraft is flying at the IAS for
minimum drag (VIMD), because in level flight thrust is equal to drag.
FIS Book 1: Aerodynamics 322

108. Maximum endurance is obtained at an altitude which is governed by engine considerations.


Although for a given set of conditions the IAS for minimum drag remains virtually constant at all
altitudes, the engine efficiency varies with altitude and is lowest i.e. poorest at the lowest altitudes
where rpm must be severely reduced to provide the low thrust required.

109. To obtain the required amount of thrust most economically, the engine must be run at
maximum continuous rpm. Therefore maximum endurance is obtained by flying at such an altitude
that, with the engine(s) running at or near optimum cruising rpm, just enough thrust is provided to
realize the speed for minimum drag, or MCRIT, whichever is the lower. Above the optimum altitude
little, if any, additional benefit is obtained, and in some cases there may be a slight reduction because
burner efficiency decreases at or about the highest altitude at which level flight is possible at VIMD. In
general, optimum endurance is obtained by remaining between 20,000 ft and the tropopause at the
recommended IAS and appropriate rpm. The greater the Power/weight ratio of the aircraft, the
greater will be the optimum height. With aircraft having high Power/weight ratios, maximum
endurance is obtained at the tropopause.

110. Altitude should only be changed to that for maximum endurance if the aircraft is above or near
the endurance ceiling, otherwise if the aircraft is climbed from a much lower altitude, a considerably
higher fuel flow will be required on the climb thereby reducing overall endurance. On engines having
variable swirl vanes, the consumption increases markedly if the rpm are so low that the swirl vanes
are closed. If the altitude is low enough to cause the swirl vanes to close at the rpm required for VIMD,
the aircraft should either be climbed to the lowest altitude at which VIMD can be obtained with the swirl
vanes open, or the rpm increased to the point at which the vanes open, accepting the higher IAS.

111. Effect of Weight. Drag and thrust at the optimum IAS are functions of the all-up weight.
The lower the weight the lower the thrust and fuel flow. Endurance varies inversely as the weight and
not as the square root of the weight as in range flying because in pure endurance flying the TAS has
no importance.

112. Effect of Temperature. In general, the lower the ambient air temperature, the higher the
endurance, due to increased thermal efficiency, and vice versa. However, the effect is not marked
unless the temperature differs considerably from the standard temperature for the altitude. In any
case, the captain can do nothing but accept the difference, since any set of circumstances requiring
flying for endurance usually ties the aircraft to a particular area and height band.
113. Twin and Multi-Engine Aircraft. When flying for endurance in twin or multi-engine aircraft
at medium and low altitude, endurance can be improved by shutting down one or more engines. In
this way the live engine(s) can be run at rpm closer to the optimum for the thrust required to fly at
VIMD, thus improving GFC. Provided that the correct number of engines are used for the height,
altitude has virtually no effect on the endurance achieved.
323 Range and Endurance

114. Use of the Fuel Flow meter. The fuel flow meter is a useful aid when flying for endurance.
If the endurance speed is unknown, the throttle(s) should be set at the rpm which give the lowest
indicated rate of fuel flow in level flight for the particular altitude.

115. Conclusions. Maximum endurance is achieved by flying at an altitude where optimum rpm
give the minimum drag speed. It will rarely pay to climb to a higher altitude unless the commencing
altitude is very low. In any event, the instruction or need to fly for endurance may preclude this. At the
lower altitudes, maximum endurance may be obtained either by flaming-out engines to use optimum
rpm on the remainder, or by using near-optimum rpm to give VIMD. It should be remembered that:

(a) The importance of flying at VIMD outweighs engine considerations, always assuming
that an engine, or engines, will be stopped in the low level case.
(b) At lower altitudes there will be a slight decrease in endurance due to the higher
ambient temperature reducing engine thermal efficiency.

Derivation for Jet Endurance Parameter Expression

116. It is known that whereas range flying is concerned with the distance flown on a given fuel,
endurance flying means flying the ac for as long as possible on the fuel available, in other words it
involves the factor of maximum time on a given fuel. To be precise

Specific Endurance = ___Flight time___ = _t_


quantity of fuel M

On dividing the numerator and denominator by time we get

Specific endurance = _Flight time/time__ = t/t


quantity of fuel/time M/t
Specific endurance = ____________1_____________ = _1_
Quantity of fuel used per unit time. GFC

Thus Specific endurance it is the reciprocal of fuel flow.

For Maximum Specific Endurance 1 / GFC should be maximum or GFC should be minimum.

__1__ Maximum implies _____1____ maximum (from (18.2))


GFC SFC X thrust

Specific Endurance = _____1_____ maximum (in level flight Thrust = Drag)


SFC X Drag

On multiplying this by L / W since L = W in level flight, we get.

Specific Endurance = L X 1 X _1_.


W SFC Drag
FIS Book 1: Aerodynamics 324

Specific Endurance = L X 1 X 1.
D SFC W
Specific Endurance = CL X _1_ X 1 (18.17)
CD SFC W

Therefore for maximum endurance, jet aircraft is flown under conditions of maximum L/D or maximum
CL/CD with minimum SFC, and aircraft as light as possible.

117. Total Endurance. For the total endurance of an aircraft Specific Endurance bas to be
multiplied with the available fuel. However it should be noted that as fuel is consumed in level flight
weight also reduces a Specific Endurance. In very large transport aircraft where a very big proportion
of weight is made up of fuel, the specific endurance expression has to be integrated with respect to
weight appropriately, to calculate endurance.
325

CHAPTER 19

TRANSONIC AND SUPERSONIC AERODYNAMICS

Introduction

1. Aircraft flying at speeds well below the speed of sound send out pressure disturbances, or
waves, in all directions. This enables an approaching aircraft to be heard and, more important for the
aircraft, the air is warned of its approach. This warning gives the air time to divide and this allows the
passage of the aircraft with minimum disturbance.

2. If air was incompressible, the speed at which pressure disturbances travelled would be
infinite, therefore the disturbance created by an aircraft would be felt everywhere instantaneously,
regardless of the aircraft’s speed, also the flow pattern around the aircraft would be independent of its
speed. However, air is compressible and a change of density and temperature accompanies a
change of pressure. Because of this, the speed of propagation of the pressure wave has a finite
value - the speed of sound.

3. As the speed of an aircraft is increased, there is a decrease in the distance ahead of the
influence of the advancing pressure waves. There is also a change in the flow and pressure patterns
around the aircraft and this will ultimately change its manoeuvrability, stability and control
characteristics. To understand the changes in these patterns it is helpful to examine the changes that
take place when a source of small pressure waves is moved through the air. The distinguishing
feature of small pressure waves is that they travel at the same speed (i.e. sonic speed) as they
radiate from their source. Sound waves are audible pressure waves and, of course, also travel at the
speed of sound.

Definitions

4. Definitions of terms, which will be used throughout this chapter, are as follows:

(a) Speed of Sound (a). The speed at which a very small pressure disturbance is
propagated in a fluid under specified conditions.

(b) Mach Number (M). M = the speed of an object or flow = V


the speed of sound in the same a
part of the atmosphere

Notes: 1. Both V and a may vary.


2. M is ratio and has no units.
FIS Book 1: Aerodynamics 326

(c) Free Stream Mach Number (MFS). This is the Mach number of the flow sufficiently
remote from an aircraft to be unaffected by it.

Notes: 1. M = V, therefore M = __ TAS_________


a local speed of sound
2. MFS is sometimes called flight Mach Number.
3. Ignoring small instrument errors, MFS is the true Mach number of an
aircraft as shown on the mach meter.

(d) Local Mach Number (ML). When an aircraft flies at a certain MFS, the flow is
accelerated in some places and slowed down in others. The speed of sound also changes
because the temperature around the aircraft changes. Hence, ML may be higher than, the
same as, or lower than MFS.

(e) Critical Mach Number (MCRIT). As MFS increases, so do some of the local Mach
numbers. That MFS at which any ML has reached unity is called the critical Mach number. As
will be seen later, MCRIT for an aircraft or wing varies with angle of attack. It also marks the
lower limit of a speed band wherein ML may be either subsonic or supersonic. This band is
known as the transonic range.

(f) Detachment Mach Number (MDET). If the leading edge of an aerofoil has no
leading edge radius, there will be a MFS at which the bow shockwave attaches to the sharp
leading edge (or detaches if the aircraft is decelerating from supersonic speed). This value of
MFS is MDET. At MDET all values of ML are supersonic except in the lower part of the boundary
layer. In practice this rarely happens since wings have a significant radius of the leading
edge, and behind the shock wave immediately in front of the leading edge there will be a
small area of subsonic flow. Nevertheless MDET can still usefully be defined as that MFS
above which there is only a small movement of the bow shockwave with an increase in
speed. MDET can therefore be used to indicate the upper limit of the transonic range.

(g) Critical Drag Rise Mach Number (MCDR). MCDR is that MFS at which, because of
shockwaves, the CD for a given angle of attack has increased significantly. Different criteria
are often used, i.e. 0.002 rise in CD, 20% rise in CD etc.

5. Definitions of Flow. Fig 19-1 illustrates the flow speed ranges. Changes in airflow occur at
ML = 1.0 and the boundary between each region of flow is that MFS that produces an ML appropriate to
that region.

Notes: 1. The subsonic region has been subdivided at M = 0.4 since below this Mach
number errors in dynamic pressure, assuming incompressibility, are small. However,
327 Transonic and Supersonic AD

compressibility effects can be present even at M = 0.4, whether they are or not
depends on wing section and angle of attack.
2. The actual values of MCRIT and MDET depend on individual aircraft and angle
of attack.
3. A third definition of the transonic range is that range of MFS during which
shockwaves form and move significantly.

Fig 19-1: Flow Speed Ranges

Wave Propagation

6. The two properties of air which determine the speed of sound within it are resilience and
density. Resilience describes how readily the air regains its original state having been disturbed. If
one leg of a vibrating tuning fork is considered, as the leg moves sideways, the air adjacent to it is
compressed and the molecules are crowded together. The air, recovering from its compression,
expands and compresses the air adjacent to it and so a pressure wave is propagated.

7. As each particle of air is compressed, the individual speed of the crowded molecules is
increased and this affects the expansion and compression of the particles surrounding it. The more
the air that can be compressed, the higher will be the increased speed of the molecules, the more
rapid will be the expansion and so the faster the wave will travel. Temperature is a measure of
molecular speed, and therefore increasing molecular speed suggests a rise in temperature in the air.
However, this is not so because the situation is adiabatic, i.e. there is no overall rise in temperature
and the temperature change in the production of small pressure waves is infinitesimal. But if the
temperature of the air is increased as a whole, then the added increase in molecular speed as the air
is compressed is sufficient to cause an increase in the rate of expansion, and therefore an increase in
the speed of propagation of the pressure wave, i.e. the speed of sound has been increased by an
increase in temperature.

8. The relationship of mach no in terms of pressure and density is a2 = γP


ρ
FIS Book 1: Aerodynamics 328

where γ is the ratio of the specific heats = (SH Constant Pressure) = 1.4 for air.
(SH Constant Volume)
This can be rewritten as a2 = γ R T where R is the gas constant and T is in °K.
Therefore a = √ (γ R T)
Substituting for the two constants, a = 39 √ T kt
∴ it can be seen that a ∝ √T (19.1)

9. It should be noted that, in the transmission of


pressure waves, air is not physically displaced. It merely
vibrates about a mean position, as does the prong of a
tuning fork. The wave motion can be seen from a graph
showing pressure variations against distance from a
source at an instance of time (Fig 19-2).

10. The size of the wave is measured by its Fig 19-2: Pressure Waves from a
Stationary Source
amplitude, i.e. the maximum change of pressure from
static. The amplitude of a sound wave is approximately
one millionth of an atmosphere. In free air the waves
weaken with distance from source because the energy in
the wave is spread over an ever-increasing surface. The
amplitude decreases rapidly at first, and then more
gradually until the wave becomes too weak to measure.

Pressure Waves from a Moving Source

11. Source Moving at Subsonic Speed. Fig 19-3


shows the pattern produced at a given instant in time.
Fig 19-3: Pattern from a Subsonic
The sound waves emanate in all directions relative to the Source
source, although they are closer together ahead of the
source than behind it. The waves maintain their
separation and there is no tendency for them to bunch.

12. Source Moving at Sonic Speed. A different


pattern is produced when the source is moving at the
speed of sound (Fig 19-4). The waves cannot move
ahead of the source and they bunch up and form a Mach
wave ahead of which the air is quite unaware of the
existence of the sound source. The Mach wave is not a
Fig 19-4: Pattern from a Sonic
shockwave, it is merely a line dividing areas where the Source
source can be heard and areas where it cannot. As the
329 Transonic and Supersonic AD

Mach wave is at right angles to the direction of movement of the source the wave is called a Normal
Mach wave.

13. Source Moving at Supersonic Speed. At supersonic speeds yet another pattern is
produced (Fig 19-5), again with a boundary beyond which no wave can pass, i.e. the limit of influence
of the source. This boundary is called an
oblique Mach wave and the angle it
makes with the flight path is called the
Mach angle (m). From Fig 19-5 it can be
seen in triangle AOB that sin m = a / V,
but as M =V / a (paragraph 4b), sin m =
1 / M. In the air this pattern is three-
dimensional so the boundary becomes a
surface called the Mach cone. To
distinguish between the regions where the
air is affected by the source and where it
is not, the terms Mach after-cone and
Mach fore-cone respectively are used. Fig 19-5: Pattern from a Supersonic Source

14. Mach Waves. It should be noted that a flow passing through a Mach wave is affected
infinitesimally since the pressure change across the wave is infinitely small. A Mach wave, or line, is
a line along which a pressure disturbance is felt and the significance of Mach lines will be appreciated
when expansions in supersonic flow are considered later.

15. Large Pressure Waves. In para 3 it was stated that small pressure waves travel at uniform
speed. When large pressure waves are produced (say, by a wing) there is a significant temperature
increase at the source. This increase reduces with distance from the source and changes the speed
of wave propagation as the waves radiate from the source. Due to this temperature rise, there is an
increase in the local speed of sound and the initial speed of propagation of the pressure wave is at
that higher sonic speed.

SHOCKWAVES

Formation of a Bow Shockwave

16. The pressure pattern from an aerofoil section at some subsonic speeds is well known. In
particular, the region close to the leading edge will be experiencing pressure higher than atmospheric.
Since the pressure and temperature are increased, then the speed of sound in that region will be
higher than in the undisturbed flow some distance ahead of the wing. Using the same reasoning, by
beginning at the leading edge where the pressure is highest, then moving up-stream ahead of the
FIS Book 1: Aerodynamics 330

wing, pressure and temperature will reduce until finally


a point is reached where they have reduced to the free
stream static values. Between this point and the
leading edge, the speed of sound will be higher than
that of the free stream because of the compression and
consequent temperature rise. The symbol
conventionally given to this locally increased speed of
sound is a'. If a typical curve of pressurization were
plotted from its static value ahead of the leading edge,
the result would be as shown in Fig 19-6. Fig 19-6: Pressure Variations from a
Static Source

17. When a wing or any object is placed in a flow, the flow around the object is deflected. If the
deflection is such that the air is compressed, then the deflection and speed of the flow will determine
the compression and temperature rise and, consequently, the local increase in the speed of sound.
Consider an infinitely thin flat plate. If it is at 0° angle of attack, there will be no pressure changes
round it. If it is set at a positive angle of attack, the pressure rise underneath will increase the
temperature such that the increase in speed of propagation of the pressure waves ahead of the plate
is a' (where a is the speed of sound in the undisturbed flow). The initial value of a' depends on the
shape and speed of the source, the blunter the source, the greater the increase in the local speed of
sound.

18. Considering now a wing at an MFS < 1.0, the


pressure disturbances from the wing will propagate
forward initially at some value of a', which will
decrease with distance until its value is equal to a.
Since the aircraft is subsonic, i.e. its speed is less
than a, the disturbances will continue to progress up-
stream until they finally die out. They will not bunch
and no Mach wave will form. This is shown in Fig
Fig 19-7: Pressure Variations from a
19-7. Subsonic Source

19. When MFS = 1.0, the pressure waves propagate initially at a value of a', greater than MFS,
resulting from the heated high-pressure region ahead of the wing. They can, therefore, escape
forward away from the advancing wing for as long as the temperature and pressure remain higher
than ambient temperature and P0 (the free stream static pressure) respectively at which stage the
disturbance cannot make any further progress up-stream. Thus, a Mach wave will form at this point
well ahead of the aircraft and the blunter the wing leading edge (more disturbance) the further away it
will be. The location of the Mach wave is shown on Fig 19-8. Note that it is still a Mach wave - not a
shock wave - because pressure is P0 at this point.
331 Transonic and Supersonic AD

20. An increase in MFS above M 1.0


results in an increase in the initial a', but at
some distance ahead of the aircraft the
reducing value of a' will equal the forward
speed of the aircraft, i.e. a' will equal MFS.
The pressure waves will therefore be unable
to penetrate further up-stream, as shown in
Fig 19-9. The pressure waves will bunch
and form a shockwave called the bow
shockwave. This wave is the forward limit of
influence of the aeroplane and the air ahead Fig 19-8: Pressure Variations from a Source at
of it receives no warning of the aircraft’s the Speed of Sound
approach.

21. As far as an observer on the ground


is concerned, there is no aircraft noise
ahead of the bow shockwave, but as the
shockwave passes the observer there is an
intense increase in noise, the sonic "bang",
and then the normal aircraft sound is heard.
The second "bang" comes from the
shockwave formed around the wings and
Fig 19-9: Pressure Variations from a Supersonic
fuselage of the aircraft and will be explained Source
in a later paragraph.

22. If the aircraft continues to accelerate, the pressure waves reach their forward limit of travel
even closer to the source. Once again the bow shockwave will be positioned where the decreasing a'
is equal to the MFS and will be more intense. Eventually, when the MFS reaches the same value as the
initial a' of the disturbance, the bow shockwave attaches to the leading edge. This is at the
detachment Mach number (MDET) and any further increase in speed results in the shockwave
becoming more oblique. This stage will only be reached if the leading edge is very sharp, since the
value of a' is a function of speed and shape, and increases with speed. Where there is a rounded
leading edge, an increase in speed will give rise
to increased stagnation temperatures and the
shock will be "held off".

23. The disturbances, which cause an


aerofoil to generate shockwaves, originate at the
points where the airflow suddenly changes
direction - notably at the leading and trailing Fig 19-10: Deflection Angle
FIS Book 1: Aerodynamics 332

edges. The angle through which the airflow changes direction at those points is known as the
deflection angle (Fig 19-10). Below is a table of detachment Mach numbers for particular deflection
angles with regard to the bow wave.

Deflection Angle MDET


10° 1.41
20° 1.83
30° 2.00

24. Comparing the case V > a (small


source) and V > a' (large source), it is
found that because a' is greater than a, the
wave angle (W) is always greater than the
Mach angle (m), as shown in Fig 19-11.
Also from Fig 19-11,

Sin W = a’ Fig 19-11: Wave Angle Greater than Mach Angle


V

Multiplying by a/a,

Sin W = a’ X a = a’ X 1 from which W can be calculated


a V a M

25. To summarize:

(a) MFS < 1.0 - no bow shockwave.


(b) MFS > 1.0 < a' - detached normal shockwave.
(c) MFS > a' - oblique attached shockwave.
(d) The blunter the source, the higher MDET will be.

Wing Surface Shockwaves

26. The bow shockwave is not the first shockwave to form. Long before the aircraft as a whole
has reached Mach 1, some portions of the flow around the aircraft have exceeded the speed of
sound. When the first local Mach number (ML) equals 1.0 somewhere around the aerofoil, the aircraft
has reached its critical Mach number (MCRIT) and shockwaves will begin to form.

(a) Conditions for Shockwave Generation. Fig 19-12a shows the local Mach
numbers around a wing at a speed in excess of MCRIT (MFS = 0.8 in the example) at the
instant in time before any shockwave has had time to develop. Where the top and bottom
airflows meet at the trailing edge, they are deflected violently (each through a different
333 Transonic and Supersonic AD

deflection angle) to follow a common path. At the point where they meet there will be another
region of high pressure similar to the region of high pressure at the leading edge. Beneath
the wing, since the highest ML = 0.95, the pressure waves travelling at sonic speed are able
to escape forwards. Similarly,
above the wing nearer the
trailing edge, in the flow where
ML = 0.9 and 0.95, pressure
waves can progress forwards
but they are soon brought to
rest at the point where ML =
1.0.
Fig 19-12(a): Local Mach Nos

(b) Top Shockwave. Fig


19-12b shows the situation an
instant later when only a top
shockwave forms. Shockwave
generation is complicated
somewhat by the presence of
the boundary layer, part of
which must always be Fig 19-12(b): Top Shockwave
subsonic. This enables the pressure rise to be communicated forward of the shock,
thickening the boundary layer and giving rise to a further compressive corner ahead of the
main shockwave known as the lambda foot. This interaction between the shockwave and the
boundary layer is dealt with in more detail in the chapter on Design for High Speed Flight.

(c) Bottom Shockwave.


As MFS is increased the top
shockwave will grow steadily
stronger and start to move
rearwards, eventually attaching
to the trailing edge. At some
stage during this movement, Fig 19-12(c): Formation of Bottom Shockwave

some local Mach numbers


below the wing will reach unity
and a bottom shockwave will
form (Fig 19-12c). This will be
weaker than the top shockwave
because a wing at a positive
Fig 19-12(d): Bottom Shockwave Moves
angle of attack will feature a to Trailing Edge
smaller deflection angle for the
FIS Book 1: Aerodynamics 334

bottom shockwave. This smaller deflection angle also means that the bottom shockwave will
move rearwards faster than the top shockwave and will attach to the trailing edge before it
(Fig 19-12d).

(d) Progression to MDET. As MFS is increased further, both surface shockwaves


become attached to the trailing edge (Fig 19-12e) and, as speed is increased even further,
become more and more
oblique. The bow wave forms
once MFS exceeds 1.0 and
moves towards attachment but,
for a blunt leading edge, does
not quite achieve this stage as
Fig 19-12(e): Both Wing Shockwaves at the Trailing
explained in para 22 and Edge
illustrated in Fig 19-12(f).
Conversely, the sharper the
leading edge the sooner will
the weaker shockwave become
attached (at MDET) and any
further increase in Mach
number will cause the normal Fig 19-12(f): Bow Wave Just Detached
shockwave to become oblique.

Acceleration of a Supersonic Flow

27. To examine the changes that take place


when supersonic flow is accelerated, consider the
flow through a venturi tube, as in Fig 19-13. The
up-stream end is connected to a reservoir
containing air, the total pressure P0 and
temperature T0 of which are maintained constant.
The down-stream end is connected to another
tank in which the pressure may be varied. If no
pressure difference exists along the duct there is
no flow. If the pressure P is decreased slightly to
P1, then a low speed flow results, producing a
pressure distribution as shown. Gradual lowering
of the pressure P will lead to larger flow speeds
until at pressure P4 there is a throat pressure of
0.528P0 and a flow speed of M = 1.0. If the Fig 19-13: Supersonic Flow Acceleration
pressure P is decreased even further to P5, the in a Duct
335 Transonic and Supersonic AD

change in pressure will be transmitted by a pressure wave, which will travel at Mach 1.0 up-stream
from the downstream end of the venturi. This can reach the throat but there it will meet the flow
travelling downstream at the same speed and so it will be unable to pass through the throat. The flow
up-stream of the throat will therefore be unaware of any changes in pressure downstream of the
throat, and will remain unchanged no matter how large or small are further reductions in P. This
produces several important results:

(a) M = 1.0 is the highest Mach number that can be obtained at the throat.
(b) At M = 1.0 the pressure will always be 52.8% of the total head pressure. (In this case
THP is P0 because the situation started with no flow.)
(c) Maximum mass flow is obtained with M = 1.0 at the throat. The flow at the throat is
sonic and so mass flow is independent of the down-stream pressure.
(d) The streamlines are parallel at the throat.

28. The reduction in pressure is, however, felt up to the throat and the adverse pressure gradient
is removed in a small section of the duct. Now a sonic flow is encountering diverging conditions in a
cross-sectional area of no pressure gradient. What happens to the flow can be best explained by
considering the small section of the duct, which has just started to diverge from the throat.

29. Fig 19-14 shows a very small section of a venturi


at, and just beyond, the throat through which there is a
steady flow. A small, very thin disc of air lying between
XX and YY travelling at sonic speed will be considered
as it moves to the position X1X1 and Y1Y1. There is no
pressure difference at the moment between A and B.

30. In seeing how velocity and Mach number will


change, three separate stages are considered:

(a) On moving out of the throat, because of Fig 19-14: Supersonic Flow in a
Divergent Duct
inertia, the disc faces will want to maintain their
velocity. The duct has, however, diverged and the disc will expand to fill the available space.
In doing so the density, and with it the temperature will decrease, this increases the Mach
number.

(b) In comparison with X1X1, the greater expansion, and therefore the greater density
change is felt at position Y1Y1. Thus there will be a pressure difference across the disc and
the disc will accelerate. If the velocity increases, Mach number increases.

(c) In a steady flow the mass flow is constant and the same number of discs will pass a
point in unit time. However, the velocity has increased so there has been an apparent
FIS Book 1: Aerodynamics 336

increase in mass flow. To keep the mass flow constant the thickness of the disc must
increase and this means a further drop in density and temperature. As a result Mach number
again increases.

Notes: 1. Because of the density changes, a large increase in Mach number can be
achieved for a relatively small increase in velocity.
2. The shape of the duct controls the expansion and thus the acceleration.
3. This effect could be produced in subsonic flow if conditions of no pressure
gradient could be achieved from A to B.

Thus we have sonic flow expanding and accelerating in conditions of divergence. Streamlines of the
flow will also diverge.

31. Referring back to the whole venturi tube in Fig 19-13, a reduction of pressure to P6 means
that the area of no pressure gradient will be increased and the flow will be free to expand and
accelerate along the curve to PC. This curve is the plot of flow if the speed is supersonic for the whole
of the venturi’s diverging length, and its shape is entirely dependent on the shape of the diverging part
of the duct (See para 30, note 2).

32. At some distance down the duct the now supersonic flow will encounter an adverse pressure
gradient. The decreasing pressure of the expanding supersonic flow is equal to the pressure in the
duct and the flow will be forced to decelerate. This deceleration takes place through a shockwave.

33. Considering a curved wing surface as part of a duct the important fact to appreciate is that the
flow will be parallel to the surface and the change in surface ahead of the flow will be sensed as an
area of divergence. When an ML = 1.0 is reached the flow will be free to accelerate within the limits of
the shape of the camber of wing and until it encounters no more divergence or a strong adverse
pressure gradient.

34. The flow through a duct can be summarized as follows:

Flow Duct or Stream lines


Divergent Convergent
Subsonic P↑ ρ↑ V↓ P↓ρ↓V↑
Supersonic P↓ ρ ↓ V ↑ P ↑ ρ ↑ V↓

35. Flow through a duct has been considered but, to understand the reason behind the design of
a fully supersonic wing section, it is necessary to examine in rather more detail how a supersonic flow
negotiates sharp corners.
337 Transonic and Supersonic AD

Expansive Corners

36. As its name implies, an expansive


corner is a convex corner, which allows the
flow to expand and thus accelerate. Any
corner can be considered to consist of a series
of infinitely small changes in surface angle and
Fig 19-15 shows a single corner broken down
into a series of greatly exaggerated steps.

37. Consider the two streamlines in the Fig 19-15: Supersonic Flow Round a
supersonic flow M1 as they reach the corner. Convex Corner
The one adjacent to the surface will sense the
change in surface as soon as it gets to the corner, the flow will expand, accelerate and, associated
with the decrease in density, the pressure will decrease. This pressure disturbance will be felt along a
Mach line appropriate to the flow M1. The streamline further away from the surface will be "unaware"
of the corner until it reaches the Mach line originating at the corner. This Mach line is a boundary
between relatively high and low pressure, therefore a pressure gradient is felt across it such that the
streamline is accelerated slightly and turned through an angle equal to the change in surface angle. It
now continues parallel to the new surface at a higher Mach number (V increased, a decreased
because of lower temperature). The streamlines are now farther apart. The same process will be
repeated at the next change in surface angle with one important difference i.e. the Mach angle of the
flow M2 will be smaller.

38. The process outlined above will take place at a


corner through an infinite number of Mach lines and,
because the lines "fan" out, the expansion and
acceleration is smooth. The region within which the
expansion takes place is limited by the Mach lines
appropriate to the speed of the flow ahead and behind
the corner and this is called a Prandtl-Meyer
Expansion (Fig 19-16). Fig 19-16: Prandtl-Meyer Expansion

39. The decrease in pressure, i.e. a favourable pressure gradient, round an expansive corner in
supersonic flow allows an attached boundary layer to be maintained. This is exactly the opposite to
what happens in a subsonic flow where an adverse pressure gradient would exist causing the
boundary layer to thicken and break away. Consequently there is no objection to such corners on
essentially supersonic aircraft and design features are permitted which would be quite unacceptable
in subsonic aircraft.
FIS Book 1: Aerodynamics 338

40. An example of expansive


corners and associated velocity and
Mach number increases is given in Fig
19-17. The initial flow is Mach = 1.0 in
standard conditions. The theoretical
maximum angle of expansion for sonic
flow is 130.5°. The vectors indicate true
velocity of the flow. It should be noted
that to increase the final speed the
temperature of the initial flow would Fig 19-17: Mach Number Increased Round
have to be increased. Expansive Corners

41. Compressive Corners. A compressive corner causes supersonic streamlines to converge,


creating a shockwave whose formation can be visualized by treating the corner as a source of
pressure waves creating a bow wave. The shockwave exhibits all the characteristics of a bow wave.
It first forms as a detached normal shockwave, moves back to the source with an increase in speed,
attaches and becomes oblique. Fig 19-18 illustrates a compressive corner obstructing supersonic
flow above MDET.

Fig 19-18: Supersonic Flow at a Compressive Corner

NATURE OF A SHOCKWAVE

42. Physical Effects of a Shockwave. A shockwave is a very narrow region, 1/400 mm thick,
within which the air is in a high state of compression. The molecules within the shock are continually
being replaced by those from the free flow ahead of the shock. Mach number can be considered as
being the ratio of directed energy (DE) to random energy (RE) within the flow. The velocity is a
measure of DE whilst RE is related to temperature. On encountering a shockwave, a flow is violently
compressed such that some DE is converted to RE (i.e. velocity is decreased, temperature is
increased), the Mach number is reduced, and pressure and density are increased. The total energy
of the flow remains constant, but there has been an increase in entropy because of the temperature
increase. This process is irreversible. This is significant because this energy is lost to the aircraft and
339 Transonic and Supersonic AD

represents drag. It is also the reason why designers try to ensure correct engine intake design for a
particular Mach number, the intention being to achieve the deceleration smoothly and with minimum
loss of heat energy through as many weak shocks as possible. The stronger the shock, the greater
the conversion to heat energy. M = V / a, therefore M2 = V2 / a2. V2 is proportional to the kinetic energy
and a2 is proportional to T (equation 19.1) therefore m2 is proportional to directed energy / random
energy. Therefore, if DE decreases and RE increases, M2 decreases.

43. Normal Shockwave. A normal


shockwave is one perpendicular to the direction
of flow. It is always detached from its source,
behind it the direction of flow is unchanged and
the Mach number is always subsonic (Fig 19-19).

44. Oblique Shockwave. The flow behind


an oblique shockwave has its Mach number
reduced, but if the wave angle (W) is less than Fig 19-19: Flow through a Normal
Shockwave
about 70° it will still be supersonic. The direction
of flow behind an oblique shockwave is turned through an angle equal to the deflection angle of the
obstruction giving rise to the shockwave.

45. Reduction in Flow Speed Behind Oblique Shockwave. Fig 19-20 shows why the
reduction in flow speed behind an oblique
shockwave is less than that behind a normal
shockwave. The actual approach flow has been
resolved into two vectors thus:

(a) X is the component normal to the


shockwave and is therefore affected by it.
(b) Y is the component parallel to the
Fig 19-20: Flow through an Oblique
shockwave and is therefore unaffected by it. Shockwave

Once the flow has passed through the shockwave:

(c) Component X has been reduced to a subsonic value, X'.


(d) Component Y is unaffected.
(e) The flow Mach number has been reduced and travels parallel to the new surface.

With the foregoing in mind two very important statements can be made about the pressures
experienced round an aircraft above MDET:

(f) All forward-facing surfaces experience greater than atmospheric pressure.


(g) All rearward-facing surfaces experience less than atmospheric pressure.
FIS Book 1: Aerodynamics 340

THE EFFECTS OF COMPRESSIBILITY ON LIFT

Introduction

46. In considering the effects of compressibility on lift it is necessary to start at a speed where
they become significant and see how they vary with an increasing Mach number. The lower limit will
be determined by section and angle of attack. This naturally leads to the consideration of three quite
distinct speed ranges:

(a) Fully subsonic flow.


(b) Fully supersonic flow.
(c) The transonic region within which there is a mixture of both.

In the case of a and b the flow behaviour is basically simple and relatively easy to predict. In the
transonic range, however, the change from stable subsonic flow to stable supersonic flow is governed
by the formation and movement of shockwaves. Unless the wing is of a suitable design this will lead
to undesirable lift and drag characteristics with consequent problems in stability and control. The CL
curve shown in Fig 19-21 is for a small symmetrical wing of 12% thickness chord ratio set at a
constant angle of attack of +2°.

Fig 19-21: Variation of CL with Mach Number at a Constant Angle of Attack

47. It is apparent from the graph that a section producing such a curve is totally unsuitable for use
in the transonic range. Even so, it is instructive to examine the reasons for its shape and see how a
thin wing will eliminate the valley at point C, which causes most of the undesirable handling
characteristics.
341 Transonic and Supersonic AD

Subsonic Rise in CL

48. Consider two identical ducts each as


shown in Fig 19-22 with identical flows in the
parallel section before the throat. The flows
have the same velocity and density but flow 1
is considered to be incompressible and flow 2
compressible.

49. Since the initial flows are the same


and the ducts have the same cross-sectional
area, then the mass flow of both will be
exactly the same and, since we are
Fig 19-22: Comparison of Flows in a Venturi
considering stream tubes, they must remain
constant, i.e. ρ X V X A = K. For the purpose of this discussion give these symbols values: let ρ = 4
units and V = 10 units, and then compare the flows in the parallel area (6 units) and throat area (3
units) sections. In the parallel section the mass flow of both flows will be 4 X 10 X 6 = 240 units.

50. Now compare the flows at the throat:

(a) In the incompressible flow, with density unchanged:


4XVX3 = 240,
∴ V = 20
(b) In the compressible flow a density change will occur at the throat associated with the
pressure decrease. Assuming the density reduces to 2 units, then:
2XVX3 = 240,
∴ V = 40.

51. An increase in velocity is always accompanied by a decrease in pressure and since the
velocity increase in the compressible flow is greater than in the incompressible flow for the same duct
(and the same wing), the pressure will be lower and thus the lift will be greater for a wing in a
compressible flow. At low speed, where air can be considered incompressible, lift is proportional to
V2, and CL can be assumed to remain constant for the same angle of attack. At moderately high
speeds where density changes are becoming significant, lift increases at a rate higher than indicated
by V2. In other words CL increases for the same angle of attack.

52. Another factor which, together with the density change outlined above, affects CL is the
amount of warning the air gets of the wing’s approach, or how far ahead of the wing the streamlines
start to adjust. In Fig 19-23 the low subsonic streamline is flowing to the stagnation point. As the
speed increases, compressibility effects increase and the reduced warning causes the flow
FIS Book 1: Aerodynamics 342

displacement to start closer to the wing, this


is shown as the subsonic compressible flow.
This effectively increases the angle of attack
of the wing and so increases CL. As the
speed increases the air cannot be displaced
fast enough and so the stagnation point
moves forward as shown in the faster
compressible flow. This forward movement
causes a reduction in up-wash before the
wing which results in a loss of lift. When the
speed increases to MDET the stagnation point
Fig 19-23: Streamline Variation with Speed
is theoretically at the leading edge and there
is no up-wash at all, this is the supersonic flow.

53. A mathematical expression has been derived known as the Prandtl-Glauert Factor,
representing the increase in CL due to compressibility. A simplified expression is:

Prandtl-Glauert Factor = 1 .
2
√(1-M )

i.e., the CL may be calculated by this factor such that

CL(compressible) = CL(Incompressible) (19.2)


2
√(1-M )

This would indicate that CL would be infinite at M = 1.0 and so it needs qualification, In fact it is limited
to relatively thin wings at small angles of attack with no shockwaves present.

Transonic Variations in CL

54. In considering the CL` curve, the shockwave development and movement in the transonic
range, five significant speeds are selected: the points A, B, C, D and E in Fig 19-21.

55. In analysing the flow a percentage total head pressure, or stagnation pressure (PS), diagram
is used as a convenient method of showing instantaneous values of speed and pressure at any point
on the wing for any given MFS. The method of interpreting a PS diagram is as follows.

56. In Fig 19-24, 25, 26, 27 and 28 the pressure at the forward stagnation point, where the flow is
brought to rest, is given the value of unity. This pressure is the sum of the dynamic and static
pressures, and is known as the total head or stagnation pressure and it is always greater than
atmospheric pressure. On either side of the stagnation point the flow accelerates causing a rise in ML
and an associated fall in pressure. When the flow has accelerated to a value equal to MFS the
343 Transonic and Supersonic AD

pressure will again be atmospheric and a further acceleration above MFS will cause the local pressure
to fall below atmospheric.

Notes:

1. Pressures are plotted as percentages of the PS value on the left-hand axis. The
value decreases from the origin.
2. ML is plotted on the right-hand axis, increasing from 0 at the origin.
3. For every ML there is a corresponding pressure for the flow. For ML = 1.0 percentage
PS is 52.8% (see para 27b).
4. The horizontal axis represents the wing chord.
5. The position of the centre of pressure and variations in CL at different MFS may also
be judged from the diagrams.
6. Where the pressure in the flow has been reduced to static pressure, the Mach
number of the flow is the free stream Mach number.
7. The percentage total head pressures are shown as follows:
Upper surface -- blue line
Lower surface – green line
8. The area between the lines is a measure of the amount of lift being delivered. If the
upper surface pressure is lower than the lower surface pressure this will be positive lift, and
vice versa.

57. At point A in Fig 19-21 the MFS is 0.75 and the flow accelerates rapidly away from the
stagnation point along both upper and lower surfaces, giving a sharp drop in pressure through and
below the atmospheric figure. Values of ML greater than 1.0 are found between 4% and 25% chord
(Fig 19-24), thus the wing is above its MCRIT.

58. When ML reaches 1.0, the streamlines,


which previously were converging during
acceleration, are now parallel. Any further
movement across a convex surface will cause the
streamlines to diverge, thus giving a supersonic
acceleration which would theoretically continue
right to the trailing edge. However, this does not
happen in practice because of the "whiskering
effect".

59. The "whiskers" are, in effect, weak shock


waves through each of which the flow is retarded
Fig 19-24: Pressure Pattern at Point A
slightly, the cumulative effect being to prevent
further supersonic acceleration and eventually to decelerate the flow smoothly back to a subsonic
FIS Book 1: Aerodynamics 344

value. Their origin can best be imagined by considering the compressive corner formed at the trailing
edge. The separated flow leaving the trailing edge is sensed by the streamline flow as a deflection
angle and, since the wake is turbulent, the value of the deflection angle will continually be changing,
as will the amplitude of the pressure waves originating there. The longer ones will progress further
into the supersonic flow before coming to rest than will the smaller ones. As MFS is increased the
Mach waves are swept together and they then form a sudden pressure discontinuity, i.e. a
shockwave.

60. Summarizing the situation at MFS 0.75, the wing is travelling at just above MCRIT, giving the
following conditions over the top surface of the wing:

(a) A supersonic bubble exists and as yet there is no shockwave.


(b) CL has risen by 60% of its low speed value for the same angle of attack.
(c) The centre of pressure is about 20% chord.

Over the bottom surface the flow speed is still subsonic.

61. At Point B in Fig 19-21 the MFS is 0.81. The


top shockwave has strengthened and is located at
approximately 70% chord, there is still no shockwave
on the under surface (Fig 19-25). In comparing this PS
diagram with the one for MFS = 0.75 it is significant to
note that whereas behind the shockwave on the rear
part of the wing there is no real change in the pressure
differential between upper and lower surfaces, ahead
of it and behind the 40% chord position the pressure
differential has increased considerably due to the
supersonic acceleration up to the shockwave. This
effectively increases the value of CL to approximately
double its incompressible value and causes the centre
of pressure to move rearwards to approximately 30% Fig 19-25: Pressure Pattern at Point B

chord.

62. The shape of the shockwave near the surface (Lambda foot) is caused by the pressure rise
behind the shockwave being communicated forward via the subsonic part of the boundary layer. This
effectively thickens the boundary layer, presenting a compressive corner to the flow. A series of
oblique shockwaves appear which merge into the main normal shockwave. The boundary layer
separates from the base of the shockwave causing a thick turbulent wake.

63. Summarizing the situation at MFS = 0.81, the following conditions exist:
345 Transonic and Supersonic AD

(a) There is a shockwave at about 70% chord.


(b) CL has risen to 100% of low subsonic value.
(c) The centre of pressure is about 30% chord.
(d) The flow under the bottom surface is just sonic.

64. Between MFS = 0.81 and MFS = 0.89 a


shockwave has formed on the under surface and
moved to the trailing edge while the upper surface
shockwave has remained virtually stationary (Fig 19-
26). The reason for the differing behaviour of the
shockwaves is the effect that each has on the
boundary layer. On the relatively thick wing under
consideration the top shockwave has a large pressure
rise across it, producing a very strong adverse
pressure gradient in this region. Since part of the
boundary layer is subsonic, the increased pressure is
felt ahead of the shockwave and the adverse pressure
Fig 19-26: Pressure Pattern at Point C
gradient is strong enough to cause a marked
thickening and separation of the boundary layer. The thickening of the boundary layer gives rise to a
compressive corner which tends to anchor the shockwave to this point on the wing. On the under
surface the acceleration is much less because of the angle of attack of its section, this results in a
weaker shockwave. The pressure rise across it is insufficient to cause significant boundary layer
separation and the shockwave moves quickly to the trailing edge. Such an arrangement of
shockwaves leads to the pressure distribution as shown, where it can be seen that the wing behind
the upper shockwave is producing negative lift. In computing lift, and thus CL, this negative area has
to be subtracted from the positive lift-producing area. CL has reduced to a value 30% below its
incompressible value and the centre of pressure has moved forward to approximately 15% chord.

65. Referring back to Fig 19-21, the slope of the curve between points B and C is governed by
the relative movement between the upper and lower shockwaves. The lower shockwave moves
faster and, from the time it moves aft of the upper one, a small area of negative lift develops and the
CL starts to decrease. This situation progresses until the lower shockwave reaches the trailing edge,
by which time the negative lift area is at its maximum and the CL at its minimum. With any further
increase in MFS the upper shockwave moves aft and thereby reduces the negative lift area.

66. Summarizing the situation at MFS = 0.89, the following conditions exist:

(a) The bottom shockwave is at the trailing edge.


(b) CL has reduced to about 70% of its low subsonic value.
(c) The centre of pressure is about 15% chord.
FIS Book 1: Aerodynamics 346

67. At point D in Fig 19-21 the MFS is 0.98.


Eventually the increase in ML will overcome the
resistance to rearward travel and the top surface
shockwave will be forced back to the trailing edge
(shown as MFS = 0.98 in Fig 19-27). The area of
negative lift has been replaced by an orthodox
pressure differential over the entire chord of the wing
and the turbulent wake has been reduced to its low
subsonic thickness.

68. Summarizing the situation at MFS = 0.98, the


following conditions exist:

Fig 19-27: Pressure Pattern at Point D


(a) Both shockwaves are at the trailing
edge.
(b) The turbulent wake is at its low speed thickness.
(c) CL is about 10% above its basic value.
(d) The centre of pressure has moved back to about 45% chord. This is the rearward
movement of the centre of pressure experienced by all aircraft going through the transonic
range.

69. At point E in Fig 19-21 the MFS is 1.4. Above


M = 1.0 the bow shockwave forms and at M = 1.4 (Fig
19-28) is almost attached to the leading edge (had the
leading edge been very sharp, the bow wave would
have attached at MDET). The whole of the wing is
producing lift and the centre of pressure is in the
approximate mid-chord position. CL has reduced to a
value of 30% less than its incompressible value. The
reasons for this are:

(a) The loss of circulation due to the


stagnation point moving to the most forward
Fig 19-28: Pressure Pattern at Point E
point on the leading edge (para 52).
(b) There has been a loss of pressure energy through the bow shockwave.

70. Summarizing the situation at MFS = 1.4, the following conditions exist:

(a) CL is 70% of its low subsonic value.


(b) The centre of pressure is about 50% chord.
347 Transonic and Supersonic AD

71. The effect of changing angle of attack on shockwaves at constant MFS is now considered. An
increase in angle of attack will cause the minimum pressure point to move forward along the top
surface. Lower pressures and a higher ML will occur for a given MFS. The effects are as follows:

(a) The top surface shockwave forms at a lower MFS and moves forward as angle of
attack is increased.

(b) Separation affects a greater length of chord.

(c) The influence of the shockwave on the boundary layer is more severe. Lower
pressures are reached as a result of local acceleration and give higher adverse pressure
gradients with the shockwave. This pressure gradient affects the thickness of the boundary
layer under the shockwave.

(d) The position and intensity of the two shockwaves alter considerably as angle of attack
changes. This erratic movement has a large effect on the total lift and drag of the aerofoil and
on the centre of pressure position. These changes give rise to transonic stability and control
problems in manoeuvre.

The consequences of varying the angle of attack are confined to the transonic range. While the bow
shockwave remains attached, there will be little movement of the centre of pressure. The limiting
factor when manoeuvring supersonically is usually determined by the amount of tailplane movement
available.

Supersonic Fall in CL

72. In fully supersonic flow (above MDET) for thin wings, regardless of camber, Ackeret’s Theory
states that CL = ___4α___ . Substituting this in the lift formula:
√(M2 –1)
Lift = ___4α___ ½ ρ V 2S
√(M2 –1)
But a2 = γ P, therefore re-arranging for ρ and substituting:
ρ
Lift = __4α__ ½ γ P V 2 S
√(M2 –1) a2
= __4α__ ½ γ P M2 S
√(M2 –1)
Lift = 2α γPS __ M2__ (19.2)
√(M2 –1)

If α is a constant then Lift is now purely a function of ____M2___ and theoretically lift becomes
√(M2 –1)
FIS Book 1: Aerodynamics 348

infinite at M = 1.0, decreases to a minimum when M = 1.4 and then slowly increases as MFS is
increased. In practice, any decrease in lift between MDET and MFS = 1.4 would be masked by trim
changes produced in passing through the transonic range.

73. The practical result of this is that


supersonically the lift curve slope becomes
progressively more gentle with an increase in
MFS (Fig 19-29).

Variations in Maximum Value of CL

74. In the derivation of the lift formula, CL


was shown to be dependent on Mach number
and Reynolds number (RN), as the aircraft
speed increased, so did RN. Therefore with Fig 19-29: Slope of CL Curve
an increase in speed it would appear that the
maximum value of CL would increase (Figs
19-30 and 31). An increase in flow speed
across a wing causes an increase in the
adverse pressure gradient, therefore the
transition point moves forward. The turbulent
boundary layer has more energy in its flow
and will therefore remain attached to the wing
longer, moving the separation point aft. The
wing can therefore be taken to a higher angle
of attack and a higher value of CL before it Fig 19-30: Effect of an Increase in Reynolds
reaches its critical angle of attack. However, Number

with a further increase in speed the mach


number increases and so do compressibility
problems.

75. Another factor to be considered is that


with a further increase in speed the adverse
pressure gradient will be even greater,
causing the separation point to move forward.
This, therefore, lowers the maximum value of
CL.

76. The theoretical result is that there is Fig 19-31: Theoretical Variation
an increase in CL max for an increase in speed
349 Transonic and Supersonic AD

until the effects of compressibility and adverse pressure gradient are felt. In practice the value of CL
max depends on the effectiveness of the aircraft’s elevators
as to whether the pilot is able to use the increased value of
CL max. Generally the usable value of CL max decays with a
speed increase, as shown in Fig 19-32.

Compressive Lift

77. Beyond M = 3.0, shockwaves are the predominant


feature of the relative airflow and since they cannot be
avoided, they may be deliberately cultivated to provide lift. Fig 19-32: Practical Variation
of CL max
A vehicle designed to utilize compressive lift is called a

Fig 19-33: Compressive Lift

wave rider and, in its simplest form, is a semi-ogive fitted with wings. The large under-body creates a
semi-conical shock and at design MFS the shape of the wings is such that the shock lies along the
leading edge. Fig 19-33 shows that the under-surfaces
experience a high pressure, thus providing lift.

Vortex Lift

78. Due to the spillage of air around the wing tip


caused by the mixing of the relatively high pressure air
below the wing and the relatively low pressure air above
the wing, a vortex is formed. The strength of this vortex
varies as the angle of attack. Normally this vortex is left
behind but on certain aircraft it is used to produce lift.
Fig 19-34: Slender Delta Aerofoil
FIS Book 1: Aerodynamics 350

79. On slender delta aircraft at subsonic speeds


the vortex is arranged so that it lies on top of the wing
and has its origin at or near the wing root. As the
core of a vortex is a region of low pressure, there is a
vortex core lying along the top of the slender delta
wing (Fig 19-34) creating a greater pressure change
between the top and bottom surface of the wing than
there would be if the vortices were attached to the
wing tips. The vortex system also delays separation
from the rear of the wing. Fig 19-35: Effect of Vortices on CL

80. The stalling angle for this type of aerofoil can be very large (40° or more) but the lift is
accompanied by very high drag and a low lift/drag ratio. Also any irregularity in the vortex formation
will cause adverse stability problems. However in moderate angles of attack the value of CL is
increased as shown in Fig 19-35.

THE EFFECTS OF COMPRESSIBILITY ON DRAG

Introduction

81. As the speed of an aircraft increases some factors affecting drag assume greater significance
and some arise due to shockwaves and compressibility. Those that assume greater significance are:

(a) Interference drag.


(b) Trim drag.

These can be reduced to a minimum or virtually eliminated by aircraft design. However, the drag that
arises due to the formation of shockwaves is called wave drag and although this cannot be eliminated
it can be minimized.

Wave Drag

82. Sources of Wave Drag. Wave drag arises from two sources: energy drag and boundary
layer separation.

83. Energy Drag. Energy drag stems from the irreversible nature of the changes which occur
as a flow crosses a shockwave. Energy has to be used to provide the temperature rise across the
shockwave and this energy loss is drag on the aircraft. The more oblique the shockwaves are, the
less energy they absorb, but because they become more extensive laterally and affect more air, the
energy drag rises progressively as MFS increases.
351 Transonic and Supersonic AD

84. Boundary Layer Separation. In certain


stages of shockwave movement there is
considerable flow separation, as shown in Figs 19-
25 and 19-26. This turbulence represents energy
lost to the flow and contributes to the drag. As MFS
increases through the transonic range the
shockwaves move to the trailing edge and the
separation decreases, hence the drag decreases.

85. Effect of Wave Drag. Taken together,


these low sources of drag modify the drag curve as
shown in Fig 19-36. The change in drag Fig 19-36: Effect of Wave Drag
characteristics is also shown by the CD curve for a
section at a constant angle of attack. The hump in
the curve in Fig 19-37 is caused by:

(a) The drag directly associated with


the trailing edge shocks arising from the
non-isentropic compression process
(energy loss) (see para 42).
(b) Separation of the boundary layer.
(c) The formation of the bow
shockwave above M = 1.0. Fig 19-37: Variation of CD with Mach
Number at a Constant AOA
86. Analysis of Drag Due to Shockwaves.
The mechanics of how a shockwave creates a drag force on a wing can be seen from Figs 19-38 to
19-40 where a symmetrical bi-convex is shown at three speeds. (A pressure coefficient is a non-
dimensional method of showing pressure acting at a point. Its vectorial quantity is (P-P0) / q and it is
drawn at right angles to the surface. A positive coefficient indicates a pressure greater than static and
points towards the surface.)

(a) Below MCRIT. Below MCRIT the pressure coefficients can be split into two
components relative to the flight path. The sum of all rearward-acting components minus all
the forward-acting components gives a measure of the pressure drag. At low subsonic
speeds a rough rule of thumb gives the percentage of the zero-lift drag, which is pressure
drag, as being equal to the thickness chord ratio expressed as a percentage, e.g. for 10% t/c
wing the pressure drag would be 10% of the zero lift drag, the remaining 90% being surface
friction drag. Compressibility only affects pressure drag and, as with CL, the pressure drag
coefficient can be found by applying Glauert’s factor, but the overall effect on CD up to MCRIT is
very small. The important point to note from Fig-19-38 is the pressure recovery which is
FIS Book 1: Aerodynamics 352

taking place between the point of maximum thickness and the trailing edge, i.e. flow velocity
is decreasing and pressure increasing. Fig 19-24 also shows this occurring.

(b) Just below MFS = 1.0. Fig 19-


39 shows the pressure distribution at
MFS just below M = 1.0. Again,
summing the horizontal components of
the pressure coefficients will give a
measure of the pressure drag. A
comparison of Figs 19-38 and 19-39
shows an obvious increase in the drag
coefficient. The points to note are:

(i) Instead of decelerating


after the point of maximum
Fig 19-38: Subsonic Pressure Coefficients
thickness, the flow continues to
accelerate supersonically until
it reaches the shockwaves,
thus producing lower pressures
on the rearward-facing
surfaces.

(ii) These lower pressures


will increase the drag force on
the wing and it is most
important to realise that there is
a direct relationship between Fig 19-39: Transonic Pressure Coefficients

the pressures ahead of the shockwave and the strength of the shockwave itself. The
stronger the shockwave, the greater the energy loss, but also the lower the pressure
ahead of the shockwave. This is how a shockwave exerts a drag force on a body.

(c) Above MDET. Through the


leading edge shock the flow is
decelerated and deflected such that
behind the shockwave the flow is
tangential to the surface and still
supersonic (Fig 19-40). The flow now
accelerates all the way to the trailing
edge and the pressure change is such
that up to mid-chord the wing surface Fig 19-40: Supersonic Pressure Coefficients
353 Transonic and Supersonic AD

experiences pressures greater than atmospheric and from 50%-100% chord the pressures
are less than atmospheric. Therefore all the wing surface is experiencing a drag force (which
is not so subsonically). This force is dependent on the strength of the shockwave which in
turn determines the energy loss.

Above MDET the CD settles at approximately 1.5 times its low subsonic value and decreases slowly as
MFS is increased.

87. Reduction of Wave Drag. To reduce wave drag the shockwaves must be as weak as
possible, therefore wings must have a sharp leading edge as well as a thin section to keep the
deflection angle to a minimum and so produce a weak bow shockwave. The thin wing will have a
reduced camber thus lowering the pressure drop over the wing. The adverse pressure gradient
across the wing shockwaves will be smaller and the strength of the shockwaves will be reduced.
Fuselages can be treated in a similar manner. For a given minimum cross-section, an increase in
length, within reason, will reduce the wave drag.

Interference Drag

88. Where the flows over two or more parts of an aircraft meet, and these flows are at dissimilar
speeds, there is a certain amount of mixing. The faster flows slow down, the slower flows speed up
and this all takes place in an area of turbulence, this is interference drag.

89. High speed aircraft tend to have low aspect ratio wings, and to produce lift the flow over these
wings is usually much faster than the flow over other parts of the aircraft when compared with the flow
speed of slower aircraft. When these two flows meet, usually at the wing root/fuselage junction, then
due to the speed gradient of the two flows, there can be a lot of interference drag produced.

90. If the aircraft is flying at a speed where shockwaves are forming with their associated
increase in drag, the combination of wave drag, unstable wave formation at the wing/fuselage junction
and the increased interference drag could have a marked effect on the performance of the aircraft.
On a poorly-designed aircraft interference drag is by far the largest component of total drag around
MFS = 1.0.

Trim Drag

91. Trim drag is that increment in drag which results from the aerodynamic trimming of an aircraft.
Unless the thrust, drag, lift and weight forces are so arranged that there is no resultant pitching
moment, an aircraft will require a trimming force arranged about the centre of gravity so as to provide
the necessary balancing moment to maintain equilibrium. In producing this force an aircraft
experiences an increase in drag. This extra drag, trim drag, is not new but has only recently achieved
FIS Book 1: Aerodynamics 354

prominence because of the large change in the


position of the centre of pressure which takes place
through the transonic range, necessitating extremely
large trimming forces (Fig 19-41).

92. If equilibrium is to be achieved, all forces must


balance, i.e. the horizontal forces (thrust and drag)
must be equal, and so must the vertical forces. Since
LT is acting downwards, in order to maintain level
flight, lift must be made equal to W + LT. To produce
this extra lift a higher angle of attack is required which,
in turn, will mean more drag.
Fig 19-41: Balancing the Shift of the
Centre of Pressure
93. Trim drag then has two sources, and it will
readily be seen that it can be present to some degree at any speed. Subsonically, because the
movement of the centre of pressure with speed is relatively small, trim drag can be minimized by
suitable design. Consider a supersonic aircraft of conventional design with static longitudinal stability
subsonically. As it accelerates through the transonic range there will be a steady rearwards
movement of the centre of pressure. This gives a strong nose-down pitching moment requiring very
large trimming forces and resulting in a large increase in drag. It is because of this rearward shift of
the centre of pressure associated with supersonic flight that trim drag has become such a problem.

94. With careful design this rearward shift of the centre of pressure can be minimized but not, as
far as is known at the moment, eliminated. For example, the Concord restricts the movement of the
CP to something like 8% of the chord. In terms of feet, this is still a considerable distance, and so the
CG is also moved rearwards as the aircraft accelerates, by pumping fuel aft. However, this is only
one method; the subject is discussed more fully in the Chapter on design for high speed flight.

THE EFFECTS OF INCREASING MACH NUMBER ON STABILITY

Introduction

95. Certain aspects of stability have already been discussed. The purpose of the following
paragraphs is to cover, very briefly, the main effects of compressibility on an aircraft’s stability.

Transonic Longitudinal Stability

96. Most aircraft which operate in the transonic speed range experience a nose-down pitch with
an increase in speed. Although there are two causes of this the relative importance of each cannot be
stated without reference to specific aircraft. The two causes are:
355 Transonic and Supersonic AD

(a) Rearward Movement of the Centre of Pressure. This increases the longitudinal
stability. Because of the supersonic acceleration at the higher speed, pressure continues to
decrease past the 50% chord point, thus increasing the amount of lift produced by the mid-
chord part of the wing.

(b) Modification to the Airflow Over


the Tailplane. Most tailplanes work in a
region of downwash from the mainplane;
in the case of swept-wing aircraft the
downwash can be considerable. The
formation of shockwaves on the main
plane modifies the flow such that the
downwash is reduced and this will pitch
Fig 19-42: Elimination of Downwash on
the aircraft nose-down (Fig 19-42).
the Tailplane

97. The effects on an aircraft’s handling characteristics of nose-down pitch with an increase in
speed are twofold:

(a) At some Mach number an aircraft will become unstable with respect to speed, i.e. an
increase in Mach number will necessitate a rearward movement of the control column. This
is potentially dangerous since, if the pilot’s attention is diverted, a small increase in Mach
number will give a nose-down pitch which will give a further increase in Mach number. This
in turn leads to a further increase in the nose-down pitching moment. The designer’s answer
to this problem is to ensure positive stability by use of a Mach trimmer. This is a device,
sensitive to Mach number, which simply deflect the elevator/tailplane by an amount greater
than that just to compensate for the trim change. This ensures that the aircraft has positive
stability and that the pilot will have to trim nose-down as speed increases.

(b) The requirement for a large up-deflection of the elevator/tailplane just to maintain
straight and level flight reduces the amount of control deflection available for manoeuvring.
This is especially significant on tailless deltas and in one instance is the deciding factor in
limiting the maximum speed.

98. The graph in Fig 19-43 illustrates the


magnitude of trim changes experienced in the
transonic speed range. With reference to the curve
for the Harrier GR 7, a marked tailplane trim change
occurs when accelerating through M = 0.9. There is
a smooth reversal of trim direction above M = 0.8 in
the Hawk. Note that in the Tornado transition Fig 19-43: Trim Changes with M No.
FIS Book 1: Aerodynamics 356

problems have been almost designed out throughout the range - the largest trim change occurring in
the region M = 0.8.

Supersonic Longitudinal Stability

99. The rearwards movement of the centre of pressure in the transonic region continues as the
aircraft accelerates into fully supersonic flight. Thus all aircraft experience a marked increase in
longitudinal stability with a consequent increase in the size of trimming forces and reduction in their
manoeuvring capability. The adverse effect of having to produce large trimming forces has already
been covered.

100. There are several possible methods of reducing either the rearward movement, or the effects
of rearward movement, of the centre of pressure. These are discussed in the next chapter.

Transonic Lateral Stability

101. Disturbances in the rolling plane are often experienced in transonic flight. On certain aircraft
one wing starts to drop when MCRIT is exceeded. This tendency occurs due to the difference in lift
from the two wings because in most aircraft the shockwaves do not form at identical Mach numbers
and positions on each wing.

102. Design features which normally provide lateral stability may possibly have the reverse effect
and aggravate the wing-drop. In a sideslip the down-going wing gains lift from a greater effective
aspect ratio (sweepback) or a greater angle of attack (dihedral), each method produces a greater
local acceleration of the flow. This higher local Mach number may promote or intensify an already
existing shockwave on the lower wing, possibly causing the wing to drop further.

Supersonic Lateral Stability

103. Lateral stability depends upon the


lower wing developing lift once the aircraft is
sideslipping. CL decreases in supersonic
flight and thus the correcting force is
reduced, dihedral and sweepback are
consequently less effective. An additional
effect is also worth considering, Fig 19-44
illustrates a rectangular wing at supersonic
speed. Only within the Mach cones
originating at A and B will there be a three- Fig 19-44: Mack Cones on a Supersonic Wing

dimensional flow, ie the Mach cones are


357 Transonic and Supersonic AD

bounded by the Mach lines and air flowing through these Mach lines will be aware of and respond to
the pressure difference from lower to upper surface. In the triangles ADX and BCY the average
pressure is approximately half that of the two-dimensional flow in area ABYX and so the lifting
efficiency is reduced. So is the drag, but surface friction drag is not reduced and so the lift/drag ratio
is adversely effected. This is the reason why the "tip triangles" on some wings are cut off, e.g. in
some missiles.

104. Consider now the flow pattern


with the wing side slipping Fig 19-45
shows a wing-drop to starboard.
Because of the changed relative airflow
the Mach cones affect different areas of
the wing, the lower wing is producing
much less lift and, all other things being
equal, a strong destabilising rolling
moment is set up. Again there is a Fig 19-45: Mack Cones Displaced by Sideslip
requirement for features giving lateral
stability, usually in the form of an autostabilization system.

Directional Stability

105. The trend towards rear-mounted engines, and consequently an aft CG, has meant a
decreased arm about which the fin can act. Also the supersonic decrease in CL for a given angle of
attack caused by sideslip means a reduction in fin effectiveness.

106. Subsonically, the fuselage side-force in a sideslip acts in front of the CG and the vertical tail
surfaces are able to over-come this Stable
Fin
destabilizing situation, as shown in Fig 19- Force
46. But in supersonic flight the fuselage
side-force moves forward. As long as the
aircraft is in balanced flight there is no
CG CP
problem, but if the relative airflow is off the
RAF due
longitudinal axis a destabilizing force at the
to Sideslip
nose results. This is caused by asymmetry
in the strength of the two shockwaves which
produces a pressure gradient across the
nose. A strong shockwave has a greater FIG 19-46 Subsonic Directional Stability
pressure rise across it compared to a weak shockwave.
FIS Book 1: Aerodynamics 358

107. The nose force illustrated in Fig 19-47


is tending to prevent the nose being turned
into the relative airflow and is therefore Weak Shock Wave
destabilizing. This force increases with speed Strong Shock
Small Wave
and has a longer moment arm than the fin.
Pressure
Thus above about M = 2.0 the moment Rise
becomes greater than that of the fin and
Large
rudder and the aircraft becomes directionally Pressure
Rise
unstable (Fig 19-48). The point of application
of this force is difficult to define but it is Pressure Gradient
Force
located at that part of the fuselage where the
cross-sectional area is increasing. Fig 19-47: Formation of a Sideways Nose

108. One answer to this problem is to fit


longer fins and to increase their number, but
obviously there is a limit if only for wave drag PRESSURE
GRADIENT
considerations. A better method is to fit yaw FORCE
dampers that sense any out-of-balance CG
condition before the pilot notices it and will
correct the situation by automatically
applying the correct amount of rudder.

Fig 19-48: Instability Above M = 2.0

INERTIA CROSS-COUPLING

109. One of the problems associated with the shape and distribution of weight in modern aircraft is
that of inertial cross-coupling which imposes limitations on manoeuvrability both in combat and in
weapon delivery. There are three particular cases which should be clearly understood.

(a) Yaw divergence.


(b) Pitch divergence.
(c) Autorotation.

110. Up to the end of the Second World War


aircraft behaviour in manoeuvre was determined
almost exclusively by aerodynamic characteristics.
In more recent times, however, fuselages have
grown longer and heavier in relation to the wings.
This distribution of weight can be shown as in Fig Fig 19-49: Distribution of Weight in a
Supersonic Aircraft
19-49. The dumb-bells represent the moments of
359 Transonic and Supersonic AD

inertia which are a product of the weight and the moment arm. The moment of inertia of the fuselage
is usually referred to as B, whilst that of the wings is A. As an example, ignoring the effect of tip/drop
tank fuel, the B/A ratio of the Spitfire was 4/3, while that of Jaguar 7/1.

Yaw Divergence

111. Yaw divergence is the most common form of inertial cross-coupling encountered in practice.
It results when high rates of roll are applied during pull-up manoeuvres in aircraft with high B/A ratios,
good longitudinal static stability and poor directional stability. If the rate of roll is high then a
surprisingly small rate of pitch can be sufficient to start the sequence.

112. A simple and practical Pitch

explanation of the cycle of


Roll to Left
events starts with the aircraft
being rolled in a pull-up
manoeuvre from level flight.
Because of its longitudinal
moment of inertia the fuselage
could be likened to the wheel of
a gyroscope, (Fig 19-50). As
Fig 19-50: Explanation of Yaw Divergence
the aircraft is pitching nose-up,
the wheel is rotating clockwise viewed from the port side. If the aircraft is now rolled to the left it has
the effect of applying a torque to the top of the gyro which, when translated through 90° by the normal
laws of precession, causes the aircraft to yaw to the right. As the basic laws of precession state that
the rate of precession is inversely proportional to the rate of rotation of the gyro, if the rate of roll
(torque) is high enough only a small amount of pull-up (rotation) will induce a yaw (precession). In
fact a badly designed aircraft
might diverge at this point,
especially if it lacks sufficient
directional (or weathercock)
stability, or if the situation is
aggravated by adverse aileron
drag. More often the aircraft will
continue rolling but it may have
quite a large sideslip angle as it
passes through the 90° bank
position since the yawing action
Fig 19-51: Yaw Divergence - Roll and Pull up
will dominate the directional
stability (Fig 19-51).
FIS Book 1: Aerodynamics 360

113. As the aircraft reaches the inverted


position (Fig 19-52) the longitudinal stability of the
tailplane becomes dominant and produces a
pitching moment.

114. Having rolled through 270°, the aircraft


again reaches the 90° bank position (Fig 19-53).
The pitching moment is translated into a yawing
Fig 19-52: Yaw Divergence – Pitching
moment, the gyroscopic couple adds to this and
the only stabilizing force is the weathercock
stability due to the fin. If the aircraft lacks
directional stability at this point yaw divergence
and structural failure, or a spin, may follow.

115. The most common cause of yaw


divergence is a rolling pull-out at high Mach
numbers. The pull-out establishes the large angle Fig 19-53: Yaw Divergence –
of attack which translates quickly into sideslip as Pitch Change to Yaw
the aircraft rolls. The high Mach number
reduces the effectiveness of the fin. Fig 19-54
shows four examples of fin stability. The
aerodynamic derivative of yaw due to sideslip
(i.e. weathercock stability) is nv. Increasing
values of nv signify increasing fin stability.
These curves are for a typical modern aircraft.
The degradation of nv at both ends of the
speed scale shows that divergence in yaw is
possible at high supersonic speeds, and also
at low speeds. In describing the sequence of
Fig 19-54: Yawing Moment Due to Sideslip at
events a full roll was used as an example, 40,000 ft
however a complete roll is not always necessary by any means. Aircraft are being developed with
features designed to reduce the hazard by improving directional stability. The MiG 27 was given a
larger fin, the Phantom had its wing tips cranked upwards, the twin fins of the MiG 29 and Su 30
indicate another possible solution. Restricted aileron movement at high speed is another solution.

Pitch Divergence

116. The cycle of events which leads to a pitch divergence is exactly the same as yaw divergence
except that the aircraft will have good directional stability and poor longitudinal static stability. It is not
very common and is normally only encountered at very high altitudes.
361 Transonic and Supersonic AD

Autorotation

Fig 19-55: Roll through 90 Degrees (Nose-up Fig 19-56: Autorotation (Nose-down Principal
Principal Axis) Axis)

117. The third type of inertial cross-coupling is autorotation. If a normal aircraft flying at moderate
air speed rolls through 90° it will tend to roll about its inertial axis, or principal axes, and sideslip along
its flight path, as shown in Fig 19-55. The difference in lift between the upper and lower wings will
tend to oppose the roll, i.e. it will be stabilizing in influence. Only a badly designed aircraft would
diverge in this situation. On the other hand, if an aircraft with a nose-down principal axis is flying at
high speed, or if an aircraft is flying in a negative g condition, its inertial axis may well point below its
flight path, in which case if it rolls through 90°, the differential lift from the wings as it passes the 90°
position will tend to increase the rate of roll and autorotation may follow, as shown in Fig 19-56.
Autorotation is not often encountered in this way
at high speed but it is quite common when
rolling from the inverted position at the top of a
loop.

118. This problem is only apparent where the


manoeuvre has been incorrectly completed by
rolling off too early, and using large aileron
deflection. If an aircraft with a large B/A ratio is
rolled off the top of a loop with a negative angle
of attack, it will tend to roll about its inertial axis
(Fig 19-57). The enhanced lateral stability adds
to the roll due to aileron and autorotation may
result. In this condition the aircraft has very little Fig 19-57: Autorotation from Loop
FIS Book 1: Aerodynamics 362

forward speed and a very high rate of roll. Weathercock stability is often poor at low air speed, and
the high rates of roll that result from this situation may lead to a yaw divergence and spin. This form
of cross-coupling is common to many aircraft.
363

CHAPTER 20

DESIGN FOR HIGH SPEED FLIGHT

Introduction

1. The aim of this chapter is to give a general insight into the problems which confront the
designers of high speed aircraft and the various methods available for overcoming or reducing them,
the great variety of planforms and profiles precludes a more detailed examination. References to
many of the design features considered here can be found in the earlier chapters on the basic
principles of flight, but some of the relevant sections and diagrams are repeated here for the
convenience of the reader.

Design Requirements for High Speed Flight

2. For flight at transonic and supersonic speeds, aircraft must incorporate certain design
features to minimize the effects of compressibility. Unless the extra drag due to compressibility is kept
to a minimum, the maximum speed would be seriously affected. Furthermore, the design features
must overcome unacceptable changes in stability or control, which may be encountered in transonic
and supersonic flight. As you know friction increases temperature and also an increase in pressure
raises the temperature, as in a pump. So when an aeroplane moves through the air it gets hot, some
parts more than others, some owing to the temperature increase created by skin friction, some owing
to that created by pressure. It has been said that aeroplanes made of wax melt at 550 to 750 kmph,
those made of Aluminium at 3000 to 3300 kmph, those of stainless steel at 4300 too 4400 kmph. A
very simple formula (V / 100) ², where V is the speed in knots, gives a fair approximation to the
temperature rise in degrees Celsius. So what is merely a rise of 1ºC at 100 knots, or 4ºC at 200 knots,
becomes 36ºC at 600, 100ºC at 1000, and 400ºC at 2000 knots.

3. The design features selected for a particular aircraft depend on its intended role. The
following list indicates some of the possible important requirements for current service aircraft:

(a) High critical drag rise Mach number, (MCDR).


(b) Low drag coefficient at supersonic speeds.
(c) High maximum and low minimum speed.
(d) Adequate stability and control throughout the speed range.
(e) Good manoeuvrability at high speeds and altitudes.
(f) Stable platform for weapon release.

4. The success of an aircraft generally rests on the aerodynamic efficiency of the wings, hence it
is with wing design that this chapter is primarily concerned. Since the wing cannot be divorced from
FIS Book 1: Aerodynamics 364

the remainder of the airframe, a brief reference is made to the more general design features of the
body/wing combinations.

Wing Sections

5. The ultimate choice of wing section rests on both supersonic and subsonic considerations.
As these are not fully compatible, the result, in practice, is a compromise. The following consideration
of wing sections is developed from the theoretical optimum wing section to the practical design.

6. The Flat Plate. A low thickness/chord (t/c) ratio wing is required for high speed flight, hence
the optimum supersonic aerofoil section employs the lowest t/c ratio possible, the flat plate satisfies
this requirement. If a flat plate is placed in
supersonic flow at a positive angle of attack, the
flow pattern shown in Fig 20-1 is produced. The
flat plate in a supersonic flow has two particular
characteristics:

(a) It has a uniform decrease in


pressure over the top surface and a
uniform increase in pressure on the under Fig 20-1: The Flat Plate

surface, thus producing a uniform pressure


differential (Fig 20-2).

(b) The centre of pressure on the


plate is in the mid-chord position at all
angles of attack (Fig 20-2).

The flat plate is the most efficient aerofoil section


for supersonic flight and is thus suitable for guided
missiles. However, the plate is unsuitable for Fig 20-2: Pressure Coefficients on a Flat
Plate
piloted aircraft since there is normally a
requirement for some thickness to the wing to
accommodate control systems. Wing strength is
also necessary to avoid aero-elastic problems, and
the physical construction of the wing requires a
certain thickness. Therefore some other section
Fig 20-3: Double Wedge Aerofoil
must be employed for piloted aircraft.

7. The Symmetrical Double Wedge. If a wing must have thickness, then the next optimum
shape for supersonic flight is the diamond or symmetrical double wedge, with as low a t/c ratio as
possible, as in Fig 20-3. The double wedge is advantageous for supersonic flight because some of
365 Design for High Speed Flight

the shockwaves may be avoided. It has an optimum angle of attack for the best lift/drag ratio, viz. half
the wedge angle. The flow pattern resulting when the double wedge is placed in supersonic flow at
the optimum angle of attack, is shown in Fig 20-4. At this optimum angle of attack, only two
shockwaves exist. There is an expansion over the top rear surface and compression under the bottom
front surface. The centre of pressure is again in the mid-chord position. If the angle of attack is
varied, either two further shockwaves appear, or the existing ones intensify with the drag increasing at
a greater rate than the lift. Although the
double wedge wing is efficient in supersonic
flight, the sharp corners present difficulty in
subsonic flight. Piloted aircraft must be able
to fly subsonically, if only during take-off and
landing and therefore some other section
must be employed.
Fig 20-4: Double Wedge Aerofoil at Optimum
8. The Symmetrical Bi-Convex Wing. Angle of Attack

The Symmetrical bi-convex aerofoil section


in Fig 20-5 is a compromise between the
optimum supersonic and subsonic
requirements. It has curves for low speed
work, and it has a low t/c ratio for high speed
Fig 20-5: Bi-Convex Aerofoil
flight. Large movements of the centre of
pressure are avoided and the wing can be
made of sufficient strength. As the surfaces
are curved, the pressure changes produced
are not sudden as in the optimum
supersonic sections. In supersonic flight,
when the angle of attack is increased to the
surface angle, the shockwaves are reduced
to a minimum of two as in Fig 20-6.
Fig 20-6: Angle of Attack Equal to Surface
Angle
Thickness/Chord Ratio

9. The t/c ratios of modern aircraft wing sections fall into three main categories depending upon
the design speed range, i.e.:

(a) Subsonic. Subsonic sections have t/c ratios of 15-20%, with a point of thickness at
approximately 30% chord.

(b) Transonic. Wings designed for operation within the transonic region, usually have
a t/c ratio of 10% or less e.g. Hawk aircraft (t/c ratio of 10.9% at the wing root to 9% at the
wing tip). However, the supercritical wing, (see para 10) has t/c ratios as high as 17%.
FIS Book 1: Aerodynamics 366

(c) Supersonic. Supersonic wings have t/c ratios of 5% or less e.g. a t/c ratio of 5.5%
of the Su-30 aircraft. Additionally, the section should have a sharp leading edge to reduce
MDET and the bow shockwave intensity. It is of interest to note that the trailing edge of a high
speed wing can be thick and blunt. This effectively reduces the t/c ratio, as the wing behaves
as a thinner section with the rear part chopped off.

The Supercritical Wing

10. The supercritical wing differs from


the conventional aerofoil section in having a
relatively flat upper surface, and a more
pronounced curvature on the lower surface.
Because the airflow does not achieve the
same increase of speed over the flattened
upper surface as with a conventional wing,
Fig 20-7: Supercritical Wing
the formation of shockwaves is delayed until
a higher MFS with a consequent increase in MCDR. There is also a significant reduction in the intensity
of the shockwaves when they do form. Some of the lift lost by the flattened upper surface of the
supercritical wing is recovered by a pronounced reflex camber at the trailing edge (Fig 20-7).

11. Other obvious features of the 17% t/c ratio wing are the large section depth and the relatively
blunt, rounded leading edge. The section depth has advantages of increased storage space for fuel,
undercarriage, etc. and also offers the possibility of reducing the number of externally carried stores
with the benefit of drag reductions. Another advantage of the deep section is that a high wing
strength can be combined with a light structure. The rounded leading edge should prevent leading
edge separation and so provide reasonable low speed handling characteristics.

12. This type of wing could be used to increase performance in one of two ways as described
below:

(a) Increased Payload. By using present day cruising speeds, the fuel consumption
would be reduced, thus allowing an increase in payload with little or no drag increase over a
conventional wing at the same speed.

(b) Increased Cruising Speed. By retaining present day payloads and adopting a
thinner section supercritical wing, the cruise Mach number could be increased with little or no
increase in drag.

To make full advantage of the benefits by the supercritical wing, it must be embodied in a supercritical
design, i.e. the wing/body combination must also be supercritical. This implies accurate area-ruling,
and the elimination of any high speed points on the airframe, e.g. cockpit canopies, etc.
367 Design for High Speed Flight

Low Thickness / Chord Ratio Wings

13. On a low t/c wing, the flow acceleration is reduced, thus raising the value of MCRIT for the
wing, e.g. if MCRIT for a 15% t/c wing is M 0.75, then MCRIT for a 5% t/c wing will be approximately M
0.85. Not only is the shockwave formation delayed by the thinner wing, but also the intensity of the
waves when they do form is correspondingly less. This reduces the drag caused by energy losses
through the shockwaves, and also reduces the boundary layer separation drag because the weaker
adverse pressure gradient across the wave may be insufficient to cause separation. Any separation
that does occur will cause a less severe drag rise, as the shockwaves move rapidly to the trailing
edge. To prevent the wing shockwave from forming simultaneously along the whole length of the
wing, the t/c ratio must reduce from root to tip. The aim is to achieve the first shockwave formation at
the wing root.

14. The graph at Fig 20-8 compares the


CD of a 7% t/c ratio straight wing with that for
a 12% t/c ratio straight wing at increasing
Mach number. It can be seen that the CD
peak of the thinner wing is only one third the
value of that for the thicker wing, and also
the MCDR of the former is increased. As the
shockwaves are less intense and move
faster to the wing trailing edge, the shock
Fig 20-8: Effect of Thickness/Chord Ratio on CD
stall is avoided.

15. A similar
comparison of CL
for the two t/c ratios
is given in Fig 20-9,
and this shows that
the fluctuations in
CL are smoother
and of decreased
amplitude on the
thinner wing. This
has the effect of Fig 20-9: Effect of Thickness/Chord Ratio on Coefficient of Lift

giving the CP a more predictable rearward movement which, in turn, improves the stability of the
aircraft over the transonic speed band. Another important advantage can be seen from Fig 20-9, i.e.
the transonic speed band is reduced for the thinner wing, M CRIT is increased, and M DET reduced.

16. The use of a low t/c ratio wing section incurs some disadvantages:
FIS Book 1: Aerodynamics 368

(a) The values of CL at low subsonic speeds are reduced (Fig 20-9), thus causing
problems at approach speeds. For a given wing loading, the stalling speed increases as the
thickness/chord ratio is reduced.

(b) On thin wings, particularly when swept, it is difficult to avoid aero-elastic problems
due to wing flexing and twisting.

(c) Limited stowage space is available in thin wings for high lift devices, fuel,
undercarriage, etc.

(d) Leading edge separation is present on thin wings even at low angles of attack,
particularly when the leading edge is sharp. In some configurations, however, this separation
is controlled and can be made to increase the lift by the subsequent vortex formation.

Leading Edge Separation

17. The adoption of the thinner wing and smaller leading edge radius has resulted in a type of
stall which has assumed great importance in recent years. This is the stall which follows flow
separation from the leading edge. This may happen with a 12% t/c ratio wing, but will almost certainly
occur on an 8% t/c ratio wing.

18. On the subsonic type of aerofoil, the


classic stall starts with separation of the
turbulent boundary layer near the trailing
edge (Fig 20-10). The stall is a
comparatively gentle process with buffeting
providing plenty of pre-stall warning. As the
stall progresses, the point of separation Fig 20-10: The Low Speed Stall
moves slowly forwards towards the
transition point. On the subsonic type of
aerofoil there is no possibility of the laminar
flow separating.

19. As the angle of attack of a thin wing


is increased, the stagnation point moves to
the undersurface. The subsonic flow has to
turn an extremely sharp corner and this
Fig 20-11: Formation of a Separation Bubble
results in separation of the boundary layer
laminar flow at the leading edge with turbulent re-attachment behind a "bubble". Fig 20-11 shows an
aerofoil at 5° angle of attack.
369 Design for High Speed Flight

20. As the angle of attack increases further,


the separation bubble spreads chordwise until it
bursts at the stalling angle. Laminar flow
separation from the leading edge, without
reattachment, results. Fig 20-12 shows this
separation occurring at 10° angle of attack.

21. The leading edge stall therefore


Fig 20-12: Laminar Flow Separation
presents a serious problem since no pre-stall
warning is available. But, unlike a conventional
stall, the flow can be reattached with an increase
in speed while maintaining the same angle of
attack. After re-attachment has occurred, the
critical angle of attack increases with speed.
This is shown at Fig 20-13.

22. Points to Maximum Thickness. The


Fig 20-13: Variation of Critical Angle of
position of the maximum thickness point of a Attack with Free Stream M No.
wing also has an influence on the drag
produced. Thin wings are usually associated with a maximum thickness point at mid-chord. This has
the effect of reducing the leading edge angles to a minimum, with a corresponding reduction in
shockwave intensity. Additionally, the transition point of the boundary layer is well aft and the initial
shockwave formation is well back towards the trailing edge. Less of the wing surface is affected by
conditions behind the wave and the shockwave travels over a smaller chord distance in the transonic
region. The optimum maximum thickness position for
minimum drag is 50% chord. However, when the point is
between 40%-60% chord, the increase in CD above the
optimum is negligible.

SWEEPBACK

Sweepback and its Advantages

23. In Fig 20-14, the flow (V) across a swept wing is


divided into two flows, V1 is the component at right angles
to the leading edge and V2 is parallel to the leading edge.
As the flow componenent V2 has no effect on the flow
across the wing, the component V1 will produce the entire
pressure pattern over the wing. So it is the component of
Fig 20-14: Effect of Sweepback on
the flow normal to the leading edge that effects the value M CRIT.
FIS Book 1: Aerodynamics 370

of MCRIT and MCDR, and a higher MFS can be reached before the ML normal to the leading edge
reaches 1.0. The theoretical increase in MCDR is a function of the cosine of the angle of sweep (λ)
such that:

MCDR (Swept) = MCDR (Unswept)


Cos λ

24. To see why the component V1 normal to the leading edge determines whether the flow V will
behave in a subsonic fashion, and undergo smooth, gradual accelerations despite the flow speed
being supersonic, consider Fig 20-15. Two streamlines at some supersonic speed are shown
approaching the leading edge of a swept wing. Points O, P, Q and R are stagnation points on the
leading edge. Considering streamline A in isolation, it will be brought to rest and compressed at P
and will give rise to a Mach wave which will lie at the
appropriate Mach angle. Note that it lies ahead of the
leading edge and streamline B passes through it before
arriving at point Q. Since there is a small pressure rise
across the Mach line, streamline B will experience a
small deceleration. Between points P and Q there are a
large number of points, all sources of Mach waves,
through each of which streamline B will have to pass
and, in so doing, will be progressively decelerated before
finally coming to rest at Q. Streamlines above and below
the streamline Q will be deflected as they pass through
the part of the Mach waves oblique to the streamlines,
this divides the supersonic flow to allow the wing to pass.
Fig 20-15: Mach Waves on a Swept
There is now supersonic flow behaving in a subsonic Wing
manner and a subsonic type of wing section could be
used.

25. The Mach waves affecting streamline B will not be restricted to those originating between P
and Q. Mach waves originating at an infinite number of points on the leading edge of the wing
inboard point Q will have their effect. By the same argument, all streamlines outboard of the wing root
at O will be modified to some extent before they reach the leading edge. Consider the flow at or close
to O. Since this point is the wing/fuselage junction, there is no wing up-stream and the flow will not be
affected until it reaches the leading edge, where it will be suddenly compressed giving rise to a
shockwave. The wing in this region, more accurately described as close to the plane of symmetry,
behaves as if it were unswept and this is one of the reasons for the reduction of the theoretical
benefits of sweepback.

26. When the Mach lines lie along the leading edge, no pre-warning is possible and, since the
flow will behave in a supersonic manner, a supersonic leading edge is essential. Exactly the same
371 Design for High Speed Flight

argument can be used for points behind the leading edge. When considering the rest of the wing, the
sweep of the line of maximum thickness and the sweep of the trailing edge are important. It follows
that, certainly with a delta planform, their sweep must be less than that of the leading edge.

Effect of Sweepback on Lift and Drag

27. Fig 20-16 shows the effect on


the drag coefficient as the angle of
sweepback is increased at constant
angle of attack. It can be seen that at
the higher Mach numbers a straight wing
has the smaller drag coefficient.

28. If the total aircraft drag is


considered, the variation of the drag
coefficient will be as shown in Fig 20-17. Fig 20-16: Effect of Sweepback on CD
By delaying the wing peak CD rise until
after the fuselage peak CD at M = 1.0,
the aircraft's total CD increase is spread
over a large range of Mach numbers and
the maximum value is less than if both
fuselage and wing reached their peaks
at the same speed.

29. Lift can be treated in the same Fig 20-17: Total Aircraft Drag and M No.
way as in para 23, i.e. only V cos λ, will
produce lift. This results in the lower
slope shown in Fig 20-18.

Reduced Advantages of Sweepback

30. In paras 23 to 26 the theoretical


advantages of sweepback were
explained. In practice, the advantage is
Fig 20-18: Effect of Sweepback on CL
only about half the theoretical value.
The reason for this is the result of two effects:

(a) The change in the free-stream flow at the wing tip.


(b) Compression at the wing/fuselage junction.
FIS Book 1: Aerodynamics 372

31. Change in Free Stream Flow at the Tip.


In the wing tip area, there is a change in direction
of the free stream flow due to the influence of a
pressure gradient from the free stream pressure to
the reduced pressure on top of the wing. The
modified flow, shown in Fig 20-19, effectively
reduces the sweepback angle causing the onset of
compressibility problems at a lower MFS.

32. Compression at the Wing/Fuselage


Junction. To understand the formation of the
Fig 20-19: Change in Free Stream Flow
rear shockwave, it is necessary to examine the
flow at the wing root. As the air passes over the wing the
component of flow normal to the leading edge is accelerated by the
shape of the wing, but the spanwise component remains unaltered.
This results in the streamlines being deflected towards the wing
root as shown in Fig 20-20. As the velocity of the normal
component of the flow over the wing changes, the streamlines
follow a curved path. The effect is maximum where the
acceleration is greatest, and therefore occurs most noticeably near
the wing root where the t/c ratio is a maximum. The streamline at
the root, however, is constrained by the fuselage and will itself
affect the curvature of adjacent outboard streamlines. When the
flow passes the point of maximum thickness, the velocity
decreases and recompression takes place. At the wing root, the Fig 20-20: Inboard Flow
recompression is diffused, but the compression waves may Over a Swept Wing

coalesce along the wing and form the rear shock as shown in Fig
20-21. These two
effects cause
shockwaves to form
at a lower MFS than
the theoretical MFS
for a given angle of
sweepback. The
effectively reduced
angle of sweepback
caused by the
inboard deflection of
the streamlines Fig 20-21: Compression at the Fig 20-22: Position of Forward
results in the Wing Root and Initial Tip Shockwaves
373 Design for High Speed Flight

formation of the initial tip


shockwave and the
forward shockwave as in
Fig 20-22. The
combined effects shown
in Figs 20-19 to 22
produce a shockwave
pattern over a swept
wing as shown in Fig
20-23. With increasing
speed, the tip shockwave Fig 20-24: Shockwave Pattern
Fig 20-23: Shockwave Pattern
merges with the rear on a Swept Wing with Increased Speed And α
shockwave and the
forward shockwave joins the rear shockwave to form the outboard shockwave giving the pattern
shown in Fig 20-24. As the shockwaves combine, the total intensity increases as the sum of
individual intensities. If no flow separation has occurred with the original individual shockwaves, it
almost certainly will when the outboard shockwave forms. An increase in angle of attack will cause
the outboard shockwave to move forward and inboard as shown in Fig 20-24. The separated flow
behind this shockwave usually causes loss of lift at the wing tip which, in turn may cause pitch-up
during manoeuvre at transonic and low supersonic speeds.

High α Problems of Swept Wing Aircraft

33. The use of sweepback to obtain the advantage described above creates several undesirable
handling characteristics. The main ones are listed below and described, with their remedies, in the
subsequent paragraphs:

(a) Low slope of CL / α curve.


(b) Very high drag at large angles of attack.
(c) Tip stall.
(d) Pitch-up.
(e) Wing flexure.
(f) Dutch roll.

34. Low Slope of CL /α Curve. The strong downwash over the centre sections of a swept wing
effectively reduces the angle of attack in that region. It also delays flow separation until a very high
angle of attack is reached (Fig 20-25). For a given value of CL the higher angle of attack for the
swept wing incurs problems of visibility at low speeds, particularly on the approach to a landing. The
simplest method of overcoming this problem is to use a lower angle of attack corresponding to a lower
value of CL, and accept the resulting higher approach speeds. The problem then becomes one of
FIS Book 1: Aerodynamics 374

stopping the aircraft safely, by the use of


expensive braking devices, or using longer
runways. An alternative is to use the higher angle
of attack and lower speeds and have long
undercarriage legs to give the necessary ground
clearance in the landing attitude. This may also
require a droopable nose section to improve the
visibility from the cockpit. A further solution is the
use of variable incidence wings, the higher
incidence being used for landing, giving a high
Fig 20-25: Comparison of Straight and
angle of attack for the wing, while leaving the Swept Wing CL Curves
fuselage in a more acceptable attitude for landing
and low speed handling.

35. High Drag at High Angles of Attack. Swept wing aircraft suffer a large increase in lift
dependent drag at high angles of attack, mainly because of the development of the ram's horn vortex.
The combination of effects which causes this vortex formation is described in the chapter on Wing
Planforms. As the ram's horn vortex accounts for a major proportion of the drag at high angles of
attack, any measure which delays its onset or limits its subsequent growth is beneficial. The design
methods used to achieve these objects usually involve one or more of the following viz. checking the
boundary layer outflow, reducing leading edge separation and limiting the inboard development of the
vortex. The boundary layer outflow is reduced by:

(a) Wing fences, which not only provide a physical barrier to the outflow, but also induce
a small fore and aft vortex which carries the boundary layer rearwards with it. Inboard of the
fences, the boundary layer is thickened, thus reducing the slope of the local lift coefficient
curve (CL/α) and increasing the wing root stalling tendency. The overall effect is, however,
beneficial in that the stalling tendency is reduced at the more critical tip regions.

(b) Leading edge slats or slots, which increase the velocity of airflow over the wing
surface and re-energize the boundary layer.

(c) Boundary layer control by sucking or blowing.

(d) Vortex generators, which reenergize the boundary layer by mixing in the fast moving
air above the boundary layer.

Reduction in leading edge separation is achieved by drooping the leading edge. Examples of both
fixed and movable leading edges are illustrated below:
375 Design for High Speed Flight

(e) The droopable leading


edge operates through some
24° (Fig 20-26) this is used in
conjunction with the trailing edge
flap to give the optimum wing
camber for the phase of flight.

(f) The Kruger flap (Fig 20-


27) is used on many swept wing
transport aircraft.
Fig 20-26: Droopable Leading Edge

(g) The Concorde uses a


permanently drooped, or
cambered, leading edge (Fig Fig 20-27: Kruger Flap
20-28).

The inboard development of the ram's


horn vortex is limited by:
Fig 20-28: Cambered Leading Edge

(h) A saw-tooth leading


edge (Fig 20-29), which
provides a discontinuity in the
leading edge. This tends to
anchor the vortex at that point
and prevents further inboard
encroachment. By acting as a
vortex generator, the saw-tooth
also reduces boundary layer
outflow. The reduced t/c ratio of
the tip area has a beneficial
effect on tip stalling tendencies
(see para 37).

Fig 20-29: Saw-Tooth Leading Edge


(j) A notched leading edge
(Fig 20-30) performs the same
function as the saw-tooth,
except that it has no effect on
the t/c ratio. However, this
device is sometimes combined

Fig 20-30: Notched Leading Edge


FIS Book 1: Aerodynamics 376

with an increased t/c ratio wing section outboard of the notch.


36. Modified Drag Curve for Approach Speed Stability. Fig 20-31 compares the drag curves
for a straight wing and a swept wing aircraft, both in a clean configuration. Assuming a common
approach speed (VAPP), it can be seen that, for the same value of speed loss on the approach, the
swept wing aircraft suffers a greater drag penalty than its straight wing equivalent. If the zero lift drag
of the swept wing is increased (by lowering large flaps, extending airbrakes or deploying a drag
parachute), the total drag curve can be modified to approximately the same profile as that for the
straight wing aircraft,
although at the penalty of
higher drag values. The
effect is to reduce the drag
rise following a loss of speed
on the approach, thus making
the approach speed less
unstable. A further benefit
from the higher total drag is
that higher power settings are
needed at lower speeds and
the turbojet engine response
is much quicker from these Fig 20-31: Modification to Drag Curve for Swept Wing Low
settings than from near idling. Speed Stability

37. Tip Stall. The development of tip stalling is


covered in the chapter on wing planforms. The
alleviation of tip stalling is achieved by reducing the
boundary layer outflow. It follows, therefore, that all of
the methods described above for reducing the ram's
horn vortex effect will also be successful in varying
degrees in preventing tip stalling. In Fig 20-32, the
shaded portion of the wing represents the area of
minimum pressure over the wing top surface, with
the pressure increasing both ahead of and behind that
area. Between the two points A and B, a pressure
gradient exists. It is this gradient which gives the outflow Fig 20-32: Pressure Gradient on a
component to the wing. If the pressure gradient is Swept Wing

reduced by reducing the CL in the area of B, the outflow component is decreased. The reduction in lift
at B is achieved by washout from wing root to tip, and its effect is shown in Fig 20-33.

38. Pitch-up. Pitch-up is a form of longitudinal instability. In this case, the behaviour of the
aircraft following an increase in angle of attack is for the aircraft to continue to diverge in pitch, thus
377 Design for High Speed Flight

further increasing the angle of attack. The sequence of events following a change in angle of attack
form the self-perpetuating cycle shown in Fig
20-34. MW is the pitching moment of the wing
(normally nose-down), and MT the pitching
moment of the tail (normally nose-up). The
effectiveness of any attempt to interrupt the
cycle will be influenced by the rate of change of
pitch which has developed before the control
column is moved forward. Aircraft having large
inertia in pitch may, in extreme cases, overshoot
the demanded increase in angle of attack by a Fig 20-33: Effect of Washout on Spanwise
Loading
large margin. Therefore anticipation is required
by the pilot, especially at high angles of attack,
otherwise the aircraft may pitch-up beyond the
stalling angle of attack. On some T-tail aircraft,
pitch-up can lead to a deep stall attitude (see
para 41).

39. Wing Flexure. A thin, swept wing can


flex under load if it has insufficient stiffness.
Upward flexing of the wing causes the angle of
attack to be reduced along the wing, with the
Fig 20-34: Pitch-Up Cycle
maximum effect occurring at the wing tips. The
reduced lift on the outboard wing section causes the CP to move forward and this, in turn, may induce
pitch-up. The geometry of this washout due to flexure is described in the chapter on Wing Planforms.
To minimize the effect, a swept wing must be made stronger, and therefore heavier, than an
equivalent straight wing. A delta planform almost eliminates this problem while still retaining the
advantages of sweepback.

40. Dutch Roll. All types of aircraft can suffer from Dutch rolling. This oscillatory yawing and
rolling motion, which varies from predominantly roll (wing rock) to predominantly yaw (snaking), may
occur when lateral stability overcomes directional stability. Methods of reducing Dutch roll include the
use of autostabilizers and the reduction of the excessive lateral restoring moment inherent in a swept
wing by setting the wings at a small anhedral angle.

41. Deep Stall. Some of the characteristics of swept wings (i.e. tip stall and pitch-up) increase
the possibility of entering the deep stall region. Recovery from a stall is dependent on the pitching
moment from the wing and the pitching moment from the tailplane. In the case of a straight wing
aircraft at the stall, the CP tends to move rearwards, giving a nose-down pitching moment. On a
swept wing, however, the tendency is for the CP to remain forward, possibly giving a nose-up pitching
FIS Book 1: Aerodynamics 378

moment. All now depends on the


tailplane, which, if set low relative to
the mainplane, will be working clear of
the separated flow from the mainplane
and will therefore provide a nose-down
pitching moment and give a
conventional reaction to the stall. If,
however, the tailplane is set high (i.e.
T-tail), then increasing the angle of
attack reduces the vertical separation
between mainplane and tailplane and,
at the stall, the tailplane is immersed in
the separated mainplane wake and
therefore suffers a reduction in
effectiveness (Fig 20-35). The nose- Fig 20-35: Tailplane Position at the Stall

up pitching moment of the mainplane may be sufficient to hold the aircraft in its stalled condition. The
rate of descent which develops aggravates the condition by increasing the angle of attack even further
and may make subsequent recovery impossible. While being unsuitable from the deep stall and
pitch-up aspects, the T-tail configuration is dictated by the mounting of engines on the rear fuselage.
When mounted on top of a swept fin, the tailplane has the advantage of operating over a longer
moment arm. The tailplane also acts as an endplate to the fin and increases the effective aspect ratio
of the fin. Where the T-tail is used, an artificial stall reaction is usually designed into the control
system in the form of a stick shaker, stick pusher or both. Both are sensitive to pitching rates and
angle of attack. The stick shaker provides the stall warning (sometimes accompanied by an audible
warning), and then, if this is ignored by the pilot, at a slightly higher angle of attack the stick pusher
simulates the conventional nose-down pitching at the stall by pushing the stick forward. Some
aircraft, particularly those involved in the investigation of stall characteristics, are equipped with a drag
parachute or rocket in the tail. When used in a deep stall, the chute or rocket provides the essential
nose-down pitch moment to enable recovery to be made.

Modification to the Swept Planform

42. Para 32 explained the formation of the shockwave pattern on a swept wing. It is possible to
design a wing which is completely free of shockwaves at a supersonic MFS. The shockwaves are
eliminated by:

(a) Increasing the sweep of the wing root.


(b) Increasing the sweep of the wing tip.
(c) Fuselage waisting (area rule).
(d) Suitable wing twist and camber.
379 Design for High Speed Flight

The combined effects are illustrated in Fig 20-36 and described below.
43. Wing Root Sweep. It
has already been seen (para
32) that the wing close to the
junction with the fuselage
behaves as if it were unswept
and gives rise to a shockwave
at the wing root. This effect is
countered by increasing the
sweepback of the wing root as
shown in Fig 20-36.
Fig 20-36: Ideal Supersonic Wing

44. Wing Tip Sweep. The free stream flow at the wing tip is influenced by the disposition of the
areas of low and high pressure (Fig 20-19) and results in an effective reduction in the wing
sweepback and the formation of the tip shockwave. This effect is alleviated by rounding off the wing
tips so that the deflected flow meets the leading edge at a constant angle. Such a shape is called a
streamwise or Kuchemann tip.

45. Fuselage Waisting. Fuselage waisting prevents the re-compression waves produced at the
root (Fig 20-21) from coalescing along the span, and this prevents the formation of the rear
shockwave. Fuselage waisting is also a feature of area ruling which is discussed in para 62 et seq.

46. Wing Twist and Camber. Wing twist and camber control the spanwise and chordwise lift
distribution and thus eliminate local high lift spots which could stall or develop shockwaves before the
rest of the wing.

OTHER PLANFORMS

Delta Wings

47. There are many types of delta wings, the three main parameters being sweep of the leading
edge, sweep of the line of maximum thickness and sweep of the trailing edge. The most important
feature is to sweep the leading edge inside the Mach cone.

48. The delta wing reduces wave drag by the methods discussed under sweepback. There is also
a reduction due to the very low aspect ratio which gives a low thickness chord ratio.

49. For a brief consideration of the delta, a straight unswept trailing edge is assumed. It is, in
effect, a swept wing with the rear portion between the tips and fuselage filled in. Assuming that the
delta wing has the same advantages as the swept wing, the additional comments may be made:
FIS Book 1: Aerodynamics 380

(a) The long chord allows a good wing thickness for engine and fuel installations yet
maintains a low thickness/chord ratio.

(b) The torsional stress of a swept wing is relieved. More structural rigidity along the
flexural axis can be achieved and hence flexing under load is reduced. Stability in
manoeuvre is improved.

(c) The tip stalling tendency associated with swept wings is alleviated. The low aspect
ratio reduces the boundary layer outflow and any increased loading at the tips can be
minimized by decreasing the thickness/chord ratio in that area.

(d) An increased wing area reduces the wing loading thus permitting high altitude flight
and lower stalling speeds. It also creates more skin friction. A very high stalling angle is
produced.

(e) At the higher angles of attack, the vortices above the wing are far more pronounced.
This vortex is a combination of chord-wise growth of the separation bubble and inflow due to
the low aspect ratio.

(f) A large chord leads to an increase in skin friction drag and also requires long intakes
and/or jet pipes, etc.

Slender Deltas

50. The salient feature of the Concorde was its slender delta planform which provided low drag at
supersonic speeds because of its low aspect ratio (slightly less than 1.0), great structural strength for
durability, and plenty of stowage for fuel, etc.

51. The Concorde was deigned to cruise in the region of M 2.2 because this is the highest speed
at which cheap and well known aluminium alloys may be used for the main structure, whilst
withstanding the kinetic heating (limited to 400K). A L/D ratio of about 7.5:1 (essential for range) was
obtained at M 2.2, whilst the subsonic L/D ratio was about 12.5:1.

52. The low aspect ratio and sharp leading edge produce leading edge separation at a low angle
of attack. Previously, the vortex so produced had been an embarrassment because it was unstable,
varied greatly with angle of attack, caused buffet, increased drag and decreased the CL max. However,
with careful design the vortex can be controlled and used to advantage. By having a sharp leading
edge and drooping it by varying amounts from root to tip, such that it presents the same angle to the
local flow, attached flow is assured up to a particular small angle of attack above which the flow
381 Design for High Speed Flight

separates from root to tip with a resulting vortex all along the leading edge. Subsequent increases in
angle of attack do not significantly change the pattern of flow until an angle of about 40° is reached.

Crescent Wings

53. The crescent planform employs two methods of


raising the M CRIT and reducing the CD rise, i.e. a
combination of sweepback and thickness/chord ratio. At
the root section where the wing is thickest, the angle of
sweep is greatest; as the t/c ratio is reduced spanwise, so
is the angle of sweep. The crescent wing planform shown
in Fig 20-37 has three advantages:

(a) The critical drag rise Mach number is


raised.
(b) The peak drag is reduced.
(c) Tip stalling is prevented since the
outboard sections of the wing are virtually Fig 20-37: The Crescent Wing
unswept; hence there is less outflow of boundary
layer which gives a more uniform spanwise loading.

Polymorph Wings

54. Polymorphism is a change in wing


planform. The most common current methods of
achieving this are described below:

(a) The Fowler flap (Fig 20-38 a) is a


very efficient method of altering the wing
geometry. Extension of the flap alters
both the wing chord and camber and thus
improves the low speed handling
characteristics of high speed wing
sections. It is used extensively on modern
aircraft.

(b) Tip droop (Fig 20-38 b) performs


three functions:

(i) It improves stability and Fig 20-38: Examples of Polymorphic


Variable Geometry
manoeuvrability in supersonic
FIS Book 1: Aerodynamics 382

flight by increasing the effective fin area.


(ii) It controls the movement of the CP. The CP is moved forwards at supersonic
speeds, thus reducing the nose-down pitching moment and, consequently, reducing
the trim drag.
(iii) At supersonic speeds, the drooped tips increase the effect of compression lift
by partially enclosing the high under body pressure (Fig 19-33).

(c) Variable sweep or swing wing (Fig 20-38c) has been adopted for a number of high
speed aircraft. The advantages
gained from variable sweep are
examined below.

55. In Fig 20-38c the two extremes of


wing position can be seen. Fully swept, the
aircraft has all the advantages of a swept
wing for high speed flight, fully forward the
aircraft is an efficient subsonic aircraft with
all the advantages of a straight winged
aircraft. Fig 20-39 illustrates the difference
in drag between a straight and a swept
winged aircraft. Taking two extremes, a Fig 20-39: Drag Comparison Between Straight
straight winged aircraft would require thrust and Swept Wings

to equal "A" amount of drag for level flight at M = 1.2. However, if the wing were swept, it would
require thrust to overcome "B" amount of drag to fly at the same speed. Therefore the engine in a
variable-geometry (VG) aircraft can be smaller than for a fixed wing aircraft.

56. As the aircraft with "B" amount of drag


accelerates, the drag can be kept constant as the
wing position is progressively changed from the
straight to the swept position. In practice, the
speed at which the wings would be swept back
would be about MCDR.

57. Fig 20-40 shows a graph of the L/D ratios


for a straight and swept wing aircraft. When fixed
wing aircraft have to operate "off-design", they are
inefficient and so the fuel consumption increases.
A VG aircraft can operate over a wide speed
Fig 20-40: Lift/Drag Ratios for Straight and
range and still remain efficient. Swept Wings
383 Design for High Speed Flight

58. If the wings are arranged so that they


move within a swept root stub or "glove", i.e. only
part of the wing moves, then the aft movement of
the CP is reduced as the aircraft becomes
supersonic (Fig 20-41).

Canard Designs

59. An example of a Canard design is given in


Fig 20-42. This design can be of two kinds:

(a) The foreplane acting as a pure


trimmer; the manoeuvre in pitch is then
Fig 20-41: Movement of the CP in a VG
controlled by elevons on the mainplane. Aircraft

(b) The foreplane acting as a surface which has the same functions as a conventional
tailplane.

Fig 20-42: Canard Configuration

60. The main advantages of a Canard design are:

(a) The up-force on the foreplane generally contributes to the lift and therefore:

(i) For a given speed, a lower angle of attack is needed on the mainplane, thus
results in less drag.
(ii) There is a lower overall stalling speed which gives a better short field
performance.
FIS Book 1: Aerodynamics 384

(b) It may be possible to give the foreplane a long moment arm, thus requiring a smaller
control deflection.

61. The two main disadvantages are:

(a) For a stable configuration, the canard must always be rigged so that the foreplanes
stall before the main wings.

(b) There may be a problem with directional stability because the vortices shed from the
tips of the foreplanes react on the fin.

BODY / WING COMBINATION

Area Rule

62. In the explanation of wave drag in the previous chapter it was mentioned that, if the intensity
of the shockwaves could be reduced, the drag would be reduced. This was achieved by reducing the
thickness of the wing. The same reasoning follows for a fuselage, except that there must be minimum
cross-section for the accommodation of a crew, equipment, weapons, etc.

63. The design feature used for mating the various parts of the aircraft together to produce
minimum drag is called the Area Rule.

64. On theoretical
grounds it can be
shown that there is a
certain ideal shape for
a body of specified
volume or frontal area
which has minimum
drag at a specified
transonic or low
supersonic speed. A Fig 20-43: Principle of Area Rule
graph distribution of
its cross-sectional area along its longitudinal axis is a smooth curve. The area rule states that, if an
irregular body such as an aircraft has the same cross-sectional area distribution as this ideal body, it
will also have the minimum possible drag. Fig 20-43 illustrates this principle.

65. To see why it is possible to equate the drag of an aeroplane with that of a body of revolution,
consider the stream tube of a diameter greater than the span of the aircraft in Fig 20-44. The flow
within the tube will be distorted because of the presence of the aircraft but the distortion will be
385 Design for High Speed Flight

smoothed by pressure changes along the


circumference of the stream tube and by
the tendency of all streamlines to resist
displacement. If the diameter of the
stream tube is to remain constant, then
the acceleration within the tube will be a
function of the rate of change of cross-
sectional area of the aircraft. A body of
revolution having the same variation of
cross-sectional area will produce the same Fig 20-44: Area Rule
flow pattern as the aircraft and a typical
body for a non-area ruled aircraft is shown in Fig 20-44. The presence of the "lump", corresponding
to the aircraft's wings will obviously mean high velocities round it and a very strong shockwave behind
it. If the area is increased as shown by the dotted line, while keeping the same maximum thickness,
then the acceleration will be decreased with a corresponding decrease in drag.

66. Area ruling may be achieved in several ways:

(a) By "waisting" the fuselage.


(b) By "thickening" the fuselage ahead of and behind the wings (Fig 20-43).
(c) By combining a flat-sided fuselage with highly swept wings.
(d) By fitting pods on the trailing edge of the wings (this also "corners" the boundary
layer).

The effect of good cross-sectional distribution can be seriously reduced if care is not taken to reduce
interference drag.

Interference Drag

67. Interference drag is caused when two flows around an aircraft meet. If the flow at the
body/wing junction is considered, the speed of the flow over the wing is much greater than that along
the fuselage. When the flows mix, the speeds of the two flows change so that the flow over the wing
is reduced and the pressure pattern is changed in such a way that the wing root produces
comparatively less lift than the rest of the wing.

68. On a swept wing this change in pressure pattern has the effect of reducing the sweep of the
wing at the root. This effect, explained in para 36, has an adverse effect on transonic flight.

69. At the root trailing edge the retarded flow is entering an area of divergence between the
fuselage and wing trailing edges. This slows up the flow even more and causes an earlier boundary
layer separation, thus increasing the drag.
FIS Book 1: Aerodynamics 386

70. To reduce the interference drag at high speeds, an


effort is made to make all component parts meet at right
angles. This produces such wing/body junctions as those
shown in Fig 20-45.

Trim Drag

71. In the chapter on high speed aerodynamics, the origin


of trim drag was explained. There are two basic methods of
minimizing the effect:

(a) Reduce the movement of the CP.


(b) Rebalance of forces acting on an aeroplane in
supersonic flight so that the supersonic balance is the
same as the subsonic balance.

72. Reducing the CP Movement. There are four


Fig 20-45: Reducing
methods of reducing the movement of CP: Interference Drag

(a) Reflex Camber. Subsonically, the pressure distribution across the chord of a reflex
cambered wing is more uniform than for a section without reflex camber. This means that the
CP is further back from the leading edge than for a positive cambered wing; therefore the
rearward movement of the CP in supersonic flight is reduced.

(b) Double Delta. The double delta design (Fig 20-46) is used as a method of reducing
the movement of the CP between
subsonic and supersonic flight. At
subsonic speeds, the highly swept forward
part of the double delta is inefficient and
the CP is controlled by the more efficient
reduced sweep rear part of the wing. At
supersonic speeds, the whole wing
becomes efficient and the CP is controlled
by the whole wing, thus countering the
Fig 20-46: The Double Delta Configuration
rearward movement of the CP.

(c) Vortex Core Effect. As the vortex on top of a slender delta develops across the
chord of the wing the pressure at the core reduces from the origin to the trailing edge, and the
pressure difference from top to bottom of its wing is greater towards the trailing edge. This
again has the effect of positioning the subsonic CP in a rearwards position, reducing the
supersonic rearward movement.
387 Design for High Speed Flight

(d) Polymorph. Provided that there is a fixed section of the wing ahead of the moving
section, usually a highly swept root section, then there is a reduction in the movement of the CP
from the subsonic straight to the supersonic swept position compared to an all-swept aircraft.

73. CG Movement. Having accepted a large movement of the CP in supersonic flight, there is
currently only one method of rebalancing the forces acting on an aircraft, this is to move the CG. With
the rearward movement of CP in supersonic flight, the wing pitching moment (MW) increases and so
the correcting moment from the tailplane (MT) must also increase, giving two sources of increased trim
drag. If MW could be reduced by moving the CG closer to the CP then the aerodynamic forces
required to overcome the increased MW could be reduced; thus trim drag would be reduced. Many
aircraft move fuel to adjust their CG to reduce trim drag, e.g. Concorde, SR-71 etc.

74. Balance in Pitch. In a conventional


aircraft (Fig 20-47a) the pitching moment of the
wing (MW) is balanced by a download on the
tailplane. The result is that the wing now has to
produce more lift (LW) to balance the sum of
the weight and the elevator download (LT).
The greater lift is produced at the expense of a
higher angle of attack and thus more drag. In
the canard configuration (Fig 20-47b) the wing
pitching moment (MW) is still nose-down but,
since the foreplane is ahead of the CG, the
trimming force (LF) acts upwards and so
contributes to the total lift. The result is that
the wing lift (LW) is less than the weight and
therefore requires a smaller angle of attack
with a corresponding reduction in drag. The
final trim drag is dependent only on the
efficiency of the foreplane. Fig 20-47: Comparison Between Conventional
and Canard Trim Methods

Control

75. The traditional simple flap-type control has several disadvantages on high speed aircraft.
These are:

(a) Once shockwaves are established on a wing (or tailplane), a control deflection can no
longer modify the total flow over the wing. Its influence is limited to aft of the shockwave
where the shockwave is located ahead of the control surface or, with the shockwave at the
trailing edge, the flow in the immediate vicinity of the control. This results in a loss of control
effectiveness at higher speeds.
FIS Book 1: Aerodynamics 388

(b) Control effectiveness can be reduced even further if the control is working in a region
of shock-induced separation of the boundary layer.

(c) When shockwaves are located on or close to the control surface, a deflection of the
control will cause them to move. Their movement can significantly change the magnitude of
the forces and position of the centre of pressure on the control surfaces. The hinge moments
will therefore fluctuate rapidly, and a high frequency "buzz" will be felt through the controls by
the pilot.

76. Control Developments. The significant developments in the design of control systems are:

(a) Powered controls.


(b) All-flying, or slab surfaces.
(c) Spoilers, or inboard ailerons.

77. Powered Controls. The earliest type of powered system was the power-assisted control.
The efforts of the pilot were partially relieved, but a proportion of the force required to move the
control was still felt on the stick. The present-day philosophy is to install power-operated controls
which make the manual stick forces independent of the actual hinge moment. Such a system, by its
irreversible nature, will also prevent the erratic control movements previously mentioned. No external
force will move the control surface; moreover, mass balances are unnecessary provided that the
controls are never operated manually. Power-operated controls require artificial feel otherwise no
resistance to control deflections would be felt by the pilot, with the possible consequence that the
aircraft's structural limitations could be exceeded. The most common artificial feel systems, which
can be used singly or in combination, are covered in the Airframe précis and are described briefly
below:

(a) Spring Feel. Spring feel is the simplest system. Movement of the control
compresses the spring, and the pilot experiences a force which is proportional to the strength
of the spring and the control deflection, but independent of aircraft speed.

(b) 'q' Feel. In the 'q' feel system, resistance to control movement is provided by either
actual or simulated dynamic pressure (i.e. ρVe2 or q) and is thus proportional to Ve2.

(c) V and V3 Feel. The 'q' feel system can be modified by suitable springs and linkages
to give control feel approximately proportional to Ve and Ve3. The Ve3 feel system is used on
the rudder controls of a few aircraft to prevent the application of large control deflections and
corresponding large side-slip angles at high speeds.
389 Design for High Speed Flight

(d) g Feel. The longitudinal control forces can be modified to react to the application of
g, by the inclusion of a bob-weight into the control system. This increases the ratio of stick
force/g, and prevents excessive g application during manoeuvres.

78 Slab Surface. In the slab surface type of control, the hinged flying controls have been
dispensed with and the whole of the flying surface moves. Any change in incidence modifies the flow
over the whole surface, giving a much greater aerodynamic response. Although this type of control is
most commonly found at the tailplane, slab fins have been used and all-flying wing tips fitted to
experimental aircraft.

79. Inboard Ailerons and Spoilers. In addition to the disadvantages mentioned above, flap-
type ailerons have other undesirable features:

(a) At high IAS, large ailerons on a thin swept, high speed section could cause the wing
to twist and possibly produce aileron reversal effects.

(b) Adverse yaw can be significant at either high speed or low speed.

Two methods are used, either separately, or in conjunction, in an attempt to reduce these effects.
Inboard ailerons reduce the risk of aileron reversal because of the decreased moment arm and the
more rigid wing root section. These ailerons are sometimes supplemented by outboard ailerons for
improved low speed handling. Spoilers function by extending into the airflow and causing a
breakdown in flow and loss of lift on the required side. The main advantages of this system of roll
control are summarized below:

(c) The wing trailing edge is left clear for full span
flaps e.g. Jaguar (Fig 20-48).

(d) Because of the lower hinge moments involved,


lower control forces are required.

Ideally, a spoiler should also produce yaw in the direction of


the down-going wing (proverse yaw) and thus eliminate the
adverse yaw incurred by aileron systems. However, the use of
spoilers can incur certain disadvantages which may offset the
proverse yaw benefit. These disadvantages are summarized
Fig 20-48: Spoilers and Full-
as follows:
Span Flaps

(e) The effect of the spoiler is non-linear, with a marked reduction in performance at both
the high and low ends of the speed scale. At low speeds, ailerons may be required to
achieve a safe degree of roll control. Alternatively, a differential tailplane can be used.
FIS Book 1: Aerodynamics 390

(f) Where spoilers are used on a swept wing, loss of lift on one wing moves the CP
forward and thus introduces a pitch-up moment. This can be removed by suitable
interconnections with the elevator, but involves a more complex and heavier control system.

(g) At small deflections, the spoiler can cause control reversal. This is because the
spoiler acts as a vortex generator and actually increases lift on the wing instead of reducing it.
This can be avoided by having a minimum spoiler deflection greater than that which increases
the lift, but this can lead to either coarse control for small stick deflections, or a dead range of
movement about the stick central position.
391

CHAPTER 21

PROPELLER THEORY

Introduction

1. The turbojet produces thrust


solely from the jet efflux by accelerating a
small mass of air to a high velocity. Given
that thrust equals mass X velocity, the
same amount of thrust can be obtained by
accelerating a larger mass of air through a
lower velocity. This makes the propeller
the potentially most efficient propulsion
system. Pure jets are inherently less Fig 21-1: Thrust from a propeller
efficient but since there is virtually no limit
to the speed at which it can be operated and it works well at high altitudes, it is used where speed and
altitude are more important than economy of effort.

2. The bypass or fan engine is the half-way house between the turboprop and the turbojet. The
propeller produces thrust, at subsonic speeds and medium altitudes, more efficiently than the pure jet
or the fan-jet because, inter alia, the efficiency of a propulsive system is related inversely to the
square of the velocity of the efflux or slipstream.

3. Propellers can be divided into three broad classifications:

(a) Fixed pitch.


(b) Variable pitch, perhaps with feathering and reverse thrust capability, driven by a
piston engine. Governed in flight to a rpm selected by the pilot.
(c) Variable pitch, driven by a gas turbine, perhaps with feathering and reverse thrust
options.

4. The aerodynamics is similar for all, but the pitch control is more complex for the turboprop.
Hydraulic and mechanical pitchlocks are provided for different operating regimes. In flight, the
propeller is governed to run at constant rpm.

5. Mechanical details of propellers and their control mechanisms are explained in the précis on
aero engines.
FIS Book 1: Aerodynamics 392

Basic Principles

6. Propeller Disc. The propeller disc is the circular


area traced out by the rotating propeller blades (Fig 21-2).

7. Blade Section / Element. Each blade of the


propeller is an aerofoil and its section is an aerofoil section,
usually highly cambered. The shape, size and attitude of the
blade section may vary along the span of the blade (Fig 21-3)
i.e. with distance from the hub to tips of the propeller. Fig 21-4
shows a typical blade section in which the axis of reference is
such that OX points in the direction of motion of the axis of the
Fig 21-2: Propeller Disc
propeller while OY is the instantaneous direction of motion of
the blade section relative to that axis.

8. Blade Chord. The line OA in Fig 21-4 represents the


chord of blade element. Since the blade section varies with
radius r, i.e. distance from the propeller axis, this chord also
varies.

9. Solidity. Solidity of the propeller at radius r is


denoted by ρ, and is defined as the ratio of the sum of the
blade chords to the circumference at that radius.

ρ = Nc Fig 21-3: Propeller


2πr
Where N = No. of blades
c = Chord of blade section at radius r

10. Rotational Velocity. When the aircraft is stationary, the motion of the element under
consideration is purely rotational. The linear velocity of an element, radius (r) from the centerline, will
therefore depend on the angular velocity (N), and the radius, i.e. rotational velocity = 2πrN.
Furthermore, at a given rpm the rotational velocity will increase as r moves towards the propeller tip.

11. Forward Velocity. When the propeller is stationary, the motion of the element is entirely
due to the forward speed of the aircraft, i.e. the TAS, but when the blade is rotating, and therefore
drawing air through the disc, there is an additional element, which is induced flow. For simplicity it will
be assumed that the propeller shaft is parallel to the direction of flight. Fig 21-4 illustrates the
resultant path of the blade element under steady cruise conditions.
393 Propeller Theory

12. Helix angle. The direction


of motion of propeller is a helix instead
of a straight line, and every section of
line blade travels on a different helix.
The angle (φ) between the resultant
air flow and the plane of rotation (Fig
21-4) is called the angle of advance of
helix angle and is a different angle for
each section of the blade since the
rotational velocity varies along the
length of the blade. The sections near
the tip move on a helix of much larger
diameter, higher velocity and smaller Fig 21-4: Blade Section
helix angle as compared to blade section closer to the hub.

13. Blade Angle. In order to produce the desired aerodynamic force on the element, it is
necessary to set the blade at a small positive angle of attack to the resultant relative air flow (RAF).
Helix angle +" is the blade angle or pitch of that element so blade angle can be defined as the angle

between the blade chord and the plane of rotation of the propeller represented by angle θ in Fig 21-4.
Fig 21-5 shows the propeller of HPT-32 aircraft with its axis vertical. (Notice the angle the chord of
the propeller makes with the horizontal plane at root & at tip).

14. Blade Twist. As the helix angle reduces from hub to tip for a propeller, if all sections of the
blade are to meet the resultant airflow at the same effective angle of attack, the blade will need to be
twisted i.e. the blade angle reduced towards the tip. This is the reason for the lateral twist on a blade,
as can be made out from the propeller of HPT 32 in Fig 21-5.

Fig 21-5 Propeller of HPT-32

15. Advance per Revolution. In a propeller, at normal operating conditions, the blade angle at
each section is greater than the helix angle. The distance moved forward by the propeller in one
revolution is called the advance per revolution. It is not a fixed quantity, as it depends entirely on the
forward speed of the aeroplane. For e.g. if an aeroplane is flying at 100 m/sec and the propeller is
making 1200 rpm, i.e. 20 rps, then the advance per revolution will be 100/20 = 5 m. But the same
FIS Book 1: Aerodynamics 394

aeroplane may fly at 80 m/sec with the same rpm of the propeller and now the advance per revolution
will be only 4 m and when the engine is running on ground while there is no forward motion, the
advance per revolution will be 0.

16. Geometric Mean Pitch. If the angle of a blade section (of a fixed pitch propeller) at a radius
of r is 2º, and if this particular blade section were to move parallel to its chord so that its angle of
attack was 0º making the Helix angle the same as the Blade angle i.e. 2º while at the same time it
made one complete revolution, then the distance traveled forward would be a definite quantity and
would correspond to the pitch of the ordinary screw and the relation p = 2 π r tan 2 being true. This
advance per revolution, when the blade is at 00 angle of attack is called geometric mean pitch. This is
best seen graphically by setting off the blade angle 2 from the distance 2πr drawn horizontally, p
being the vertical height (Fig 21-6). If
the same operation is carried out at
different distances from the axis of the
propeller, it will be found that the value
of p is practically the same for all
sections of the blade, since as the
radius r increases there is a
corresponding decrease in the blade
angle 2, and 2πrtan2 remains
constant. This quantity p is called the
Geometric Pitch of the Propeller, since
it depends only on its geometric
Fig 21-6: Geometric Pitch
dimensions and not on its actual
performance. Refer to Fig 21-6.

17. Experimental Mean Pitch. The designer of a propeller finds it convenient to consider the
pitch from a totally different viewpoint. When the advance per revolution reaches a certain value, the
thrust will become zero. The reason being that the angle of attack of each part of the blade has
become so small that the aerofoil section of the blade reaches the zero lift angle of attack and thus
provides no thrust.( Notice that this corresponds to the small negative angle at which a cambered
airfoil ceases to give lift). The Experimental Mean Pitch is defined as the distance the propeller will
move forward in one revolution when it is giving no thrust. Thus in effect, when propeller is advancing
in each revolution a distance equal to the Experimental pitch, the angle of attack of the blades will be
slightly negative. On the other hand when the distance advanced per revolution is equal to the
geometric pitch, the angle of attack will be 0º. Therefore, as a general rule, the Experimental pitch will
be slightly greater than the geometric pitch, although there can be no direct relationship between the
two values since the former depends entirely on the characteristics of the aerofoil sections of which
the propeller is composed.
395 Propeller Theory

18. Forces on a Blade Element. The


aerodynamic force produced by setting the blade
element at a small positive angle of attack (i.e. the
total reaction (TR)) may be resolved with respect
to the direction of motion of the aircraft (Fig 21-7).
The component thus obtained which is parallel to
the flight path is the thrust force, and that which
remains is the propeller torque force. Notice that
the propeller torque force is the resistance to
motion in the plane of rotation. An effort must be
provided to overcome the propeller torque in order
to obtain thrust in the same way that the drag must
Fig 21-7: Forces on a Blade
be overcome to obtain lift. This torque is provided
by the engine.

PROPELLER EFFICIENCY

19. The efficiency of any system can be measured from the ratio, Output / Input or, Power out /
Power in. The power extracted from this system is the product of Force X Velocity, or Thrust X TAS.
The power put into the system is necessary to overcome the rotational drag force and is therefore the
product of Propeller Torque (Force) X Rotational Velocity. The efficiency of the propeller can
therefore be calculated from the relationship:

Propeller Efficiency = ____Thrust X Velocity_____ = Thrust X TAS (21.1)


Torque X Rotational Velocity Torque X 2πN

20. Variation of Propeller Efficiency with Speed. Fig 21-8 illustrates a fixed-pitch propeller
traveling at different speeds at a constant rpm. If the blade angle is fixed, the angle of attack will
change with variations in forward speed. In
particular, as speed increases, the angle of attack
decreases and with it, the thrust. The effect on the
propeller efficiency is as follows:

(a) At some high forward speed, the


blade will be close to the zero lift angle of
attack (Practical Pitch = EMP) and thrust
will reduce to zero. From the equation
(21.1), it will be seen that the propeller
efficiency will be zero. Fig 21-8: Effect of Speed on Fixed
Pitch Propeller
(b) There will be only one speed at which the blade is operating at its most efficient angle
of attack and where propeller efficiency will be maximum.
FIS Book 1: Aerodynamics 396

(c) At low speeds, the thrust will increase as the angle of attack is increased. Provided
that the blade is not stalled, the thrust is very large, but the speed is low and the propeller
efficiency is low. Thus no useful work is being done when the aircraft is held against the
brakes.

(d) From equation (21.1) it can also be seen that efficiency depends not only on Thrust /
Torque but also on forward velocity and rotational speed.

21. It is also clear from the above explanation that angle of attack is a function of some proportion
of the flight velocity, V, and the velocity due to rotation which is πnD at the tip. The proportions of
these terms describe the propeller "advance ratio", J.

J = _v_ Where J = propeller advance ratio


nD v = flight velocity.
n = propeller rotative speed.
D = propeller diameter, ft.

22. Theoretical analysis shows that for high efficiency, the blades will need to be like a wing
producing a high ratio of lift /drag and that under those conditions, the best helix angle approaches
450. Since the helix angle increases towards the hub, it follows that only a part of the blade span can
be operating at the most efficient angle at any time. The outer portion of the blades produces the
majority of the thrust, so the blades are normally operating at a pitch angle which gives maximum
efficiency on their outer portion. If the outer portion is running at its optimum helix angle, then a large
portion of the inner section will be running at a high and inefficient angle. The central portion of the
propeller often does little more than increase the resistance torque. It is therefore normal to terminate
the inboard ends of the blades at a streamlined spinner dome.

23. Slip. Under condition of maximum efficiency, the advance per revolution is usually
considerably less than the EMP. The difference between the practical pitch and EMP, expressed as a
percentage of EMP is called slip. For instance, suppose the Experimental pitch of a certain propeller
is 10 ft and the actual advance per revolution under certain conditions is 7 ft then the slip is 3 ft or
30%. It is thus clear that when the slip is zero the advance per revolution is equal to the experimental
pitch and this is a condition of zero thrust and efficiency. It would be incorrect to say that when the
slip is 30 % the efficiency must be 70% because this 70% is merely is ratio of distances, where as
efficiency is a ratio of work done i.e. it is concerned with force as well as distance. The best efficiency
is obtained when the slip is of the order of 30%.

24. Variable Pitch. If the engine is run at high rpm and propeller is set to have a helix angle of
0
45 near the tips for efficient cruising, then at low flight speeds, the blade α will be high. The blade
L/D ratio will be poor and if the α is too large, the blade may even stall. It is advantageous to fit a
mechanism which allows the pitch angle to be altered as the speed is varied. Fine pitch is necessary
397 Propeller Theory

for low speed & initial acceleration and


coarse pitch is needed for high speed
flight. The Fig 21-9 compares the
efficiency of a propeller at different
pitch settings. To maintain the best
pitch at all positions along the blade, it
would really be necessary to be able
to alter the blade twist as well, but as
most of the thrust comes from the
Fig 21-9: Efficiency with Aircraft Speed
outer portion of the blade, the loss of
efficiency due to non-optimal twist is
small in practice.

Propulsive Efficiency

25. The action of the


propeller can be idealized
by the assumption that the
rotating propeller is simply
an actuating disc. As shown
in Fig 21-10 the inflow
approaching the propeller
disc indicates converging
streamlines with an
increase in velocity and
drop in pressure. The
converging streamlines Fig 21-10: Propeller Disc
leaving he propeller disc
indicate a drop in pressure and increase in velocity behind the propeller. The pressure change
through the disc results from the distribution of thrust over the area of the propeller disc. In this
idealized propeller disc, the pressure difference is uniformly distributed over the disc area but the
actual case is rather different from this. The final velocity of the propeller slipstream, v, is achieved
some distance behind the propeller. Because of the nature of the flow pattern produced by the
propeller, one half of the total velocity change is produced as the flow reaches the propeller disc. If the
complete velocity increase amounts to 2a, the flow velocity has increased by the amount 'a' at the
propeller disc. The propulsive efficiency of the ideal propeller could be expressed by the following
relationship:

Propulsive efficiency = output power


Input power
FIS Book 1: Aerodynamics 398

Let the air passing through the propeller disc be subject to a force T and has a velocity u + a.
Then the rate of work done on the air = T X (u + a)

The propeller disc itself, while also subject to a force of T, has a velocity equal to u, the rate of
output work done by it is = T X u.

∴ ηp = ___T X u___ = __u__


T X (u + a) u+a
Where T = thrust.
u = flight velocity.
a = velocity increment at the propeller disc.

26. Since the final velocity, v, Is the sum of total velocity change 2a and the initial velocity, u, the
propulsive efficiency expression can be rewritten as:

ηp = __u__ Multiply and divide by 2


u+a
= ___2u__ = ____2u___ = _2u_
2u + 2a u + u + 2a u+v

So the same relationship exists as with the turbojet engine in that high efficiency is developed by
producing thrust with the highest possible mass flow and smallest necessary velocity change. The
actual propeller must be evaluated in a more exact sense to appreciate the effect of non uniform disc
loading, propeller blade drag forces, interference flow between blades, etc. With these differences
from the ideal propeller, it is more appropriate to define propeller efficiency in the following manner:

Propulsive efficiency = Output propulsive power = (T) (V)


Input shaft horsepower BHP
Where T = propeller thrust
V = fight velocity, knots ;
BHP = brake horsepower applied to the propeller.

Propeller Control

27. Piston Engine. With fixed pitch, the rpm is dependent on the power setting and the
airspeed of the aircraft. Care must be taken to avoid over speeding the engine in a dive, by throttling
back if necessary.

28. Constant Speed Propellers. Subsequent development led to the so-called ‘constant
speed’ propeller for piston engines. Power settings for piston engines are defined by rpm and
manifold pressure. The constant speed propeller enables the pitch and therefore the rpm, to be
selected anywhere between the coarse and fine pitch stops. Once selected, the engine speed is
maintained by the pitch control unit, basically a governor, despite airspeed and power variations. The
399 Propeller Theory

constant speed operation can be obtained by employing a mechanism that automatically adjusts the
blade pitch angle to alter the aerodynamic resistance torque. If an increase in speed (rpm) is sensed,
the blade is made more coarse to increase the resistance torque. The rpm setting can be altered by
the pilot by means of a selector lever. The dotted line in Fig 21-9 shows the overall improvement in
efficiency given by a variable pitch propeller, which can match the maximum efficiency envelope.

29. Turboprop. The three types of gas turbine need to be considered (Fig 21-11):

(a) Fixed Shaft. Also called the


direct drive turbine engine, the shaft
carrying the compressor and turbine
assembly drives the propeller through a
suitable reduction gear assembly. The
control of fuel flow and propeller rpm must
be matched to avoid the risk of surging,
over-fuelling or flame-out. (a) Direct Drive Turbine

(b) Free Turbine. Here, the gas


generator, comprising the compressor
and turbine driving it, is not mechanically
connected to the propeller. A separate
turbine takes power from the gas
generator, to drive the propeller. This
gives greater freedom in the control of (b) Free Powered Turbine
propeller rpm. The power available is
controlled by the throttle or power lever
and propeller rpm by a choice of governor
settings.

(c) Compound Engine. It may be


regarded as a fixed shaft engine, with
(c) Compounded Engine
similar control of the propeller. This type
of arrangement has an additional, Fig 21-11: Turbo Prop Arrangements

independent turbine driving the HP compressor.

Other Modes

30. Windmilling. If the propeller suffers an engine failure, i.e. a loss of positive torque, the
aerodynamic torque will tend to reduce the rpm. This will be sensed and corrected by the CSU and
the pitch will be fined off to reduce the angle of attack which will reduce the TR and hence the torque.
FIS Book 1: Aerodynamics 400

In this condition the RAF will impinge on the


front surface of the blade section and cause
drag and negative torque which will drive the
engine. At these small blade angles, the
propeller windmilling at high rpm can create
such a tremendous amount of drag that the
aeroplane may be uncontrollable. The
propeller windmilling at high speed in the
low range of blade angles can produce an
increase in zero lift drag which may be as
great as the zero lift drag of the basic
airplane. An indication of this powerful drag Fig 21-12: Forces on a Windmilling Propeller
is seen by the helicopter in autorotation. The
windmilling rotor is capable of producing autorotation rates of descent which approach that of a
parachute canopy with the identical disc area loading. Thus, small blade angle can produce an
effective drag coefficient of the disc area which compares with that of a parachute canopy (Fig 21-12).
A very high drag force can be generated if the propeller is directly coupled to the compressor of a gas
turbine.

31. Feathering. Following engine failure, the windmilling propeller would cause drag which
would limit range and degrade performance, perhaps leading to loss of control. Also, by continuing to
turn a possibly damaged engine, eventual seizure or fire may result. By turning the blades so that the
aggregate effect of the blade section produces zero torque, the propeller is stopped and drag reduced
to a minimum, see Fig 21-13(a) & (b). The feathered angle of most of the turboprops e.g. AN-32,
Avro, Dornier & TU-142 is close to 85º. The flight Idle angle is generally close to about 15º & during
flight the blade angle varies from 15 ºto 46 º.

Fig 21-13 a: Feathered Prop Fig 21-13 b: TU 142 (Indian Navy)

32. Reverse Thrust. If the propeller is turned through the fine pitch stop to around minus 20°
and power applied, reverse thrust is obtained. The blade section is working inefficiently, ‘up-side-
401 Propeller Theory

down’, analogous with the lift produced from


the mainplane in inverted flight. Mechanical
devices are used to prevent application of
power during the potential overspeed
situation as the blade passes beyond the
fine pitch stop, until safely in the braking
range (Fig 21-14).

33. Power Absorption. The propeller


must be able to absorb the power given to it
by the engine, i.e. it must have a resisting
torque enough to balance the engine torque,
otherwise the propeller rpm will increase and Fig 21-14: Forces on a Reverse Thrust Propeller
both the propeller and engine will become
inefficient. To increase the power absorption of the propeller there are several methods but these
reduce efficiency of the propeller in high speed flight. Hence the propeller is also a compromise, like
so many of other aspects of aerodynamics . The critical factor in matching a propeller to the power of
the engine is the tip velocity. Compressibility effects will decrease the thrust and increase the
rotational drag, considerably reducing the efficiency of the propeller blade. This consideration imposes
a limit on the propeller diameter and rpm, and the TAS at which it can be used. The factors which
increase the ability of propeller to absorb power with the limitations of each one of them are:

(a) Increasing the Blade Angle. This increases the angle of attack of the blades. The
blade angle however must be such that the angle of attack is that giving maximum efficiency.
Therefore there is little point in trying to absorb more power if in so doing we lose efficiency.

(b) Increasing the Length of the Blades. The diameter of the blade disc is increased.
This increases the aspect ratio of the blade, but apart from the limiting the speeds, the aircraft
design will limit the overall diameter of blade.

(c) Increasing the RPM. This would mean high tip speed and the consequent loss of
efficiency.

(d) Increasing the Camber. Increasing the camber of the aerofoil section of which the
blade is made, as with the aerofoil, would simply mean a less efficient section. In fact, we
need to use thinner sections to avoid loss of efficiency at high speeds.

(e) Increasing the Solidity of Propeller. Solidity of a propeller, at the radius r is given
by the expression Solidity = No of blades X chord a radius r
Circumference at radius r
FIS Book 1: Aerodynamics 402

The grater the solidity the greater the


power can be absorbed. The solidity can
either be increased by increasing the
chord or increasing the numbers of
blades. Of the two, the former is easier,
the latter more efficient.

(i) Increasing the chord.


E.g. Paddle blades, has a
limitation due to poor aspect ratio, Fig 21-15: Solidity of a propeller
which makes it less efficient.

(ii) Increasing the number of blades. A smaller number of blades, although


reduce the mutual interference effect between blades, but maintaining sufficient
solidity to transmit the required power through a given diameter may necessitated a
compromise between aspect ratio and number of blades. Therefore increasing the
number of blades is a very attractive proposition. Increasing the number of blades
reduces the amount of thrust that has to be produced by each blade, which is an
advantage in high speed operation, as it lowers the critical local mach number on the
blade. Fitting more than 4 to 5 blades onto one hub becomes inconvenient and it
becomes necessary to have two propellers for each engine. If there are two
propellers, it is beneficial to have them contra rotating which gives additional
advantage.

Contra-Rotation

34. In a simple propeller, a considerable


amount of energy is lost in the swirling motion of
the air in the slipstream. Some of this energy can
be recovered if a second propeller, rotating in the
opposite direction, is placed just downstream, as
shown in Fig 21-16. The second propeller tries to
swirl the air in the opposite direction, thereby
tending to cancel the initial swirl.

35. Contra-rotation also provides a convenient


method of increasing the power throughout for a
given propeller diameter. High-powered piston
engines produce a considerable torque reaction Fig 21-16: Contra Rotating Propeller
which tries to roll the aircraft in the opposite Accommodating a Large No. of Blades
403 Propeller Theory

direction to the propeller rotational direction. On ground, the roll is resisted by the runway, but
immediately after take-off the resistance is suddenly lost, and the aircraft is liable to start heading
rapidly probably towards an obstruction. Contra Rotating propellers overcome this problem, as they
produce no net torque reaction. Gyroscopic precession effects are also cancelled, and the lack of
swirl in the slipstream makes the flow around the aircraft less asymmetric, which further improves the
handling qualities. For small aircraft, however, the extra cost and complication of contra-rotating
propellers outweigh the advantages. On twin-engined aircraft a similar effect could be obtained by
having the two propellers (and hence engines) rotating in opposite directions. For practical reasons
this has rarely been adopted.

36. The disadvantages of contra-rotation are the extra complexity, the weight of the necessary
gearing, and the noise caused by the highly alternating flow as the second propeller chops through
the vortex system of the first.

37. Propeller Twisting Moments. Considerable stresses are placed on the propeller and the
pitch-changing mechanism in flight. The most important of these are:

(a) Centrifugal twisting moment


(b) Aerodynamic twisting moment

38. Centrifugal Twisting Moment (CTM). Fig 21-17 illustrates how the centrifugal force on the
blade can be resolved into two components. Consider points A and B on the leading and trailing
edges respectively. The centrifugal force has one component, X, which is parallel to the pitch change
axis and component Y which is at right angles to the pitch change axis. Component X produces
tensile stress along the blade but, because points A and B are not in the same plane, component Y
produces a moment around the pitch
change axis which tends to ‘fine’ the blade
off. This is the centrifugal twisting
moment. The effect of the CTM is to
make it more difficult for the pitch change
mechanism to increase the pitch of the
propeller. It can be seen from Fig 21-17
that a wider blade would increase the
moment arm of A and B about the pitch
change axis and therefore increase the
CTM. If it should be necessary to
increase propeller solidity it is therefore
preferable to increase the number of
Fig 21-17: Centrifugal Twisting Moment
blades rather than to increase the blade
width.
FIS Book 1: Aerodynamics 404

39. Aerodynamic Twisting Moment


(ATM). Fig 21-18 illustrates how the
aerodynamic twisting moment is produced.
If the pitch-change axis is behind the centre
of pressure of the blade, the torque on the
blade will be increased by an increase in
angle of attack and vice versa. It can be
seen that the ATM tends to coarsen the
pitch and therefore partially offsets the
CTM under normal operating conditions.

Fig 21-18: Aerodynamic Twisting Moment


40. Windmilling Effect on ATM.
When the propeller is windmilling in a
steep dive however, the ATM is reversed
and acts in the same direction as the CTM
(Fig 21-19). The cumulative effect may
prove a critical factor in the successful
operation of the pitch-change mechanism.

41. Swing on Take-Off. There is


often a tendency for a propeller-driven Fig 21-19: ATM & CTM on a Windmilling Propeller
aircraft with a tail wheel-type
undercarriage to ‘swing’ to one side on take-off. The causes of swing on take-off are:

(a) Asymmetric blade effect.


(b) Gyroscopic effect.
(c) Slip-stream effect.
(d) Torque reaction.
(e) Cross-wind (weathercock ) effect.

42. Asymmetric Blade Effect. Asymmetric blade effect arises from the axis of rotation of the
propeller being inclined with respect to the horizontal path of the aircraft when the tail wheel is down.
Fig 21-20 a and b illustrates the difference in propeller blade angle of attack due to the vector addition
of forward speed (horizontal) on rotational velocity (in the plane of the propeller disc).

43. Fig 21-20(a) shows the condition where the propeller axis is in the line of flight. In this case
the angle of attack and relative airflows on both propeller sections are equal. Also, the distances
travelled in unit time by both the down-going and up-going blades are equal and therefore, the speeds
of the blades are equal. Fig 21-20(b) shows how the angles of attack and resultant velocities are
changed when the axis of rotation is inclined upward.
405 Propeller Theory

Fig 21-20(a): No Asymmetric Blade Effect with Propeller Axis Parallel to Flight Path

Fig 21-20(b): Asymmetric Blade Effect with Propeller Axis tilted to the Flight Path

44. The down-going blades has a higher angle of attack, and is therefore producing more thrust
than the up-going blade. Also, the distance traveled in unit time by the downgoing blade is greater
than that for the up-going blade. This means that the down-going blade has a higher speed relative to
the airflow than the up-going blade and will, for a given angle of attack, produce more thrust. When
the propeller is rotating anticlockwise (as seen from behind), the left, hand half of the propeller disc
produces greater thrust and will yaw the aircraft to the right (HT -2 was known for this effect). If the
propeller is rotating clockwise then the right hand half of the propeller disc produces greater thrust and
will yaw the aircraft to the left. Similar lift variations are produced if the propeller is yawed instead of
pitched. If the propeller is yawed to the right, then the lower blade has a higher angle of attack and the
aircraft will pitch nose-up, similarly, the aircraft will pitch nose-down with a yaw to the left. If the
direction of rotation of the propeller is changed, then the resultant yawing and pitching directions are
changed.
FIS Book 1: Aerodynamics 406

45. Gyroscopic Effect. As the tail


wheel comes off the ground, a torque is
applied to the rotating propeller in a nose-
down sense. If the propeller is rotating anti-
clockwise (as seen from behind), the effect
of this torque on the angular momentum of
the propeller disc is to produce a ‘gyroscopic’
yawing moment to the right. Like a
processing gyroscope, the torque appears
90º removed from the point of application Fig 21-21: Gyroscopic Effect on Take Off
and in the direction of rotation. This again,
produces a yaw to the right (Fig 21-21).

46. Slip-Stream Effect. A propeller which is rotating


in clockwise direction will impart a rotation to the slip-stream
in the same sense. This rotation produces an asymmetric
flow over the fin and rudder such as to induce an
aerodynamic force to the right. This, in turn, will cause the
aircraft to yaw to the left. A propeller rotating anti-clockwise,
will produce a yaw to right (Fig 21-22).
Fig 21-22: Slipstream Effect
47. Torque Reaction. If the propeller rotates
clockwise viewed from behind, the torque reaction will tend
to rotate the aircraft in the opposite sense, i.e., roll to the left
(Fig 21-23). The rolling motion is prevented by the wheels
being in contact with the ground and this results in more
weight being supported by the left wheel than the right
wheel. This will increase the rolling resistance of the left
wheel. Consequently the aircraft will tend to swing to the left
Fig 21-23: Torque Reaction
until the wings take the weight off the main wheels.

48. Cross Wind Effect. An aircraft tends to weathercock into wind (while on ground i.e. during
rolling for take off & after landing) and this tendency must be anticipated. Taking a case of a tail
wheeled aircraft with the direction of rotation anti clockwise; all the above mentioned effects would
cause the aircraft to swing to the right. Now if the winds during take off are from the right this would
aggravate the swing. On the other hand if the winds are from the left it would reduce the swing.
[Because the cross-wind (weathercock) effect is not a purely propeller generated cause of swing on
take-off, it will not be discussed further in this chapter ; reference to it will be found in AP 3456 vol 5,
Part 2, Sec.1, Chap 2]
407 Propeller Theory

49. Sum of Forces producing Yaw on Take off. It will be noted that the effects causing a
swing on take-off, all produce yaw in the opposite direction to the direction of rotation of the propeller.
However, all may be countered by correct
application of rudder and become more
manageable as speed is increased. The
totality of effects will be more noticeable in
tailwheel aircraft (Fig 21-24). In propeller-
driven aircraft with a tail wheel
undercarriage, with the propeller turning
anti-clockwise when viewed from the
cockpit, and a cross-wind from the right,
the causes listed at sub-para (a) to (e) all
act in the same direction and will tend to
swing the aircraft to the right on take-off.
However, some aircraft configuration
compensate for some of these factors,
e.g., nose-wheel types of undercarriage,
contra-rotation propellers, biased
directional trim, etc. Fig 21-24: Swing on Take Off

50. Nose Wheel Undercarriage. Aircraft with this type of landing gear have inherent directional
stability when moving on ground. This is because the CG is ahead of the main wheels, unlike the tail
wheel configuration, which exhibits divergent tendencies. Being virtually in the flying attitude during
the whole of the take-off run, asymmetric blade effect and gyroscopic effect can be ignored. Although
still subject to torque reaction, slipstream effect and cross-winds, the swing is more easily controlled.

Future Developments.

51. As mentioned in the introduction to this chapter, propeller propulsion offers greater efficiency
is subsonic, medium altitude operation than a turbojet. This efficiency can be used to produce more
thrust from a given engine size or fuel economy by using a small engine with lower fuel consumption.

52. To extend the use of the gas turbine driven propeller into realms hitherto dominated by the
turbojet, Mach 0.8 & cruise above 30,000 ft, the advanced propeller or propfan has been developed.
Because the propeller blade is an aerofoil, by applying existing high-speed aerodynamic techniques, it
is possible to delay the limiting Mach effect at the out-board part of the blade to permit higher
operating speeds. These include reduced blade thickness, using improved aerofoil section and tip
sweep to reduce compressibility losses. The ATR- 72 which is in service with Jet Airways uses
propfan for propulsion. The Antonov Bureau has developed a medium transport aircraft AN-70 (Fig
21-25(a) & (b)), which uses propfan to produce 14000 e.h.p. To absorb this envisaged 14000 hp and
FIS Book 1: Aerodynamics 408

keep the propeller diameter within


reasonable limits, eight contra rotating
blades have been used. The aircraft can
carry max payload of 47tonnes (more
than Il-76) & can get airborne within
1800 mtrs of runway length.

High Speed Propellers

53. The primary approach to solving


the problem of supersonic tips, is the
same as that used for transonic wing
design, as described in the Chapter on
high speed aerodynamics. Essentially, it
is necessary to keep the maximum Fig 21-25: Prop Fan AN-70
relative airflow velocity on the blade
surface as small as possible. Thin 'transonic' blade sections may be used, and the blades may be
swept back producing the characteristic similar shape. Accurate control of pitch enables high lift/drag
ratio sections to be employed, and this in turn allows the use of large helix angles, so that the
resultant relative flow speed past the blade is minimised. A large number of blades are used, in order
to reduce the thrust force per blade. As with wing lift, blade thrust is related to the circulation strength.
By reducing the thrust per blade, the circulation is reduced, and hence the maximum relative speed
on the upper surface is lowered.

54. Fig 21-26 shows a typical design


for a contra-rotating configuration. Such
propellers are variously described as prop-
fans or unducted fans, and though they
may look very different from older designs,
they are, in principle, still propellers.
Although not providing such efficient
propulsion as low speed propeller
designs, unducted fan propulsion is more
efficient than the turbo-fan system that it is Fig 21-26: AN-70
intended to replace (High speed regimes). The relative airflow speed at the tips of such propellers is
designed to be supersonic in high speed flight, and they are therefore very noisy. This is the major
obstacle to their use in civil transport aircraft. It should be noted that reducing the propeller diameter
will not reduce the tip Mach number because in order to produce the same amount of thrust, the
smaller diameter propeller will have to rotate at a higher speed (rpm).
409 Propeller Theory

55. Fan Propulsion. A fan is essentially a propeller with a large number of blades, and
therefore, provides a means of producing a large amount of thrust for a given disc area. When there
are many blades, so that they are close together, each blade strongly affects the flow around its
adjacent neighbors. This interference can have a beneficial effect if the relative flow speed is
supersonic. The flow can be compressed gradually through a series of reflected shock waves,
creating a smaller loss of energy than when the flow is compressed through a single shock.

Ducted Fan

56. By placing a fan or propeller in a duct


or shroud, as in Fig 21-27, flow patterns can
be obtained, that are significantly different
from those produced by an unducted propeller
or fan. The flow patterns depend on the
relationship between the flight speed and the
engine thrust. Fig 21-27 shows two sets of
patterns, one corresponding to the low speed
high-power take off condition, and the other to
the high-speed cruise case. In this figure, the Fig 21-27: Alternative Flow Patterns for a
curved lines represent streamlines which Ducted Fan
divide the flow that goes round the outside of the duct, from that which flows through it. Again, in
three dimensions, these would correspond to stream-tubes, which may be termed dividing stream-
tubes.

57. To explain how the duct or


shroud works, we shall look first, at a
subsonic flow of air through a
converging streamlined duct with no
fan to assist it, as illustrated in Fig
21-28. Since no energy is being
added, the device cannot produce a
thrust, so the jet of air at C (where the
pressure is at the free-stream value)
cannot be moving faster than the free
stream at A. The dividing stream tube
Fig 21-28: Flow through a Streamlined Duct
diameter at A must, therefore, be no
larger than at C, since the same amount of air is passing at about the same speed at both A and C.
As the flow enters the duct at B, however, the area is larger, so the speed must be lower there. If the
speed reduces, then the Bernoulli relationship tells us that the pressure will be higher. A duct can,
therefore, provide a means of reducing the air speed and increasing its pressure locally. If we place a
FIS Book 1: Aerodynamics 410

fan in the duct, then the addition of energy to the flow can create a jet, and the streamline pattern can
be as shown in Fig 21-29. This is similar to the cruise case shown in Fig 21-27.

58. As shown, there is still a reduction


in speed, and an accompanying increase
in pressure as the flow enters the duct.
This is a very useful feature if the aircraft
is flying at a high subsonic Mach number,
because the air now enters the fan at a
lower Mach number. The Mach number is
lowered further, by the fact that the rise in
pressure is accompanied by a rise in
temperature, so that the local speed of
sound is also increased. In Fig 21-29 the
surrounding stream-tube for an unducted
Fig 21-29: Ducted Fan at High Speed
propeller has been shown, having the
same diameter as the ducted fan, and producing the same amount of thrust. A fan operated in this
way is less efficient than a free propeller of the same diameter, since the fan draws its air from a
smaller area of the free stream. The mass of air used per second, and the resulting efficiency are,
therefore, both smaller for the fan. The price is, however, worth paying, as the fan may be used at
flight Mach numbers where conventional propellers suffer excessive losses due to compressibility
effects.

59. It should be noted, that for high speed turbo-fan (fan-jet) propulsion, it is normal for the outer
part of the blade to run with supersonic relative flow between the blade and the air, but with subsonic
relative flow for the inner part

The Low Speed_ Ducted Fan or Propulsor

60. The alternative flow pattern shown in Fig 21-27 and 21-30 is obtained when the thrust is high
in relation to the free-stream speed. It therefore occurs when any turbofan aircraft is taking off. As the
aircraft speed increases, the flow pattern gradually changes to that shown in Fig 21-29. In Fig 21-30,
we have superimposed the surrounding stream tube shape for a propeller producing the same thrust.
It will be seen that when operated at low speed, the ducted fan is equivalent to a propeller of larger
diameter. The propulsive efficiency of the ducted fan should thus be higher than for an unducted
propeller of the same diameter.

61. In the situation illustrated in Fig 21-30, the flow speeds up as it approaches the fan, and the
pressure at inlet is thus lower than the free stream value. This is not a disadvantage in low-speed
flight, where there are no problems due to compressibility effects. Apart from increasing the effective
411 Propeller Theory

diameter, the ducted fan can reduce noise,


and also provide a means of containment, if
one of the blades should come off, an
important feature for the propulsion of
aircrafts. Because the fan diameter is smaller
than the equivalent propeller, it can be run at
a higher rotational speed, which is an
advantage when the drive is taken directly
from the engine shaft.

62. To prevent flow separation, the


propulsor duct intake needs to be shaped
quite differently from the high speed type, as
Fig 21-30: Ducted Fan at Low Speed
may be seen from comparison of Figs 21-29
and 21-30. The propulsor duct is rather like an annular aerofoil, and sustains a circulation. Leading
edge suction provides part of the overall thrust. The power required to produce that thrust still
ultimately comes from the engine, of course. The disadvantage of the propulsor is that it adds to the
weight, cost and the complexity of the aircraft. The duct also produces some extra surface friction
drag, and the overall increase in efficiency may be small.

63. Noise Suppression. Near-field and cabin interior noise is of great importance to the civil
operator, whilst the need for acoustic treatment of the fuselage to reduce an unacceptable level of
noise, will impose a weight penalty on both civil and military aircraft. Sweeping the outer section of the
blades lowers the sections critical Mach number and reduces the noise from the supersonic tips. The
sweep also alters the phase of the noise generated by each radial section along the blade, causing a
certain amount of phase interference which helps in noise reduction.

Fig 21-26: Experimental aircraft with VTOL:


Lockheed-XFV
FIS Book 1: Aerodynamics 412
413

CHAPTER 22

APPLICATION OF AERODYNAMICS TO SPECIFIC PROBLEMS OF FLYING

1. While the previous chapters have presented the detailed parts of the general field of
aerodynamics, there remain various problems of flying which require the application of principles from
many parts of aerodynamics. The application of aerodynamics to these various problems of flying will
assist the aviator in understanding these problems and developing good flying techniques.

Primary Control of Airspeed and Altitude

2. For the conditions of steady flight, the aeroplane must be in equilibrium. Equilibrium will be
achieved when there is no unbalance of force or moment acting on the aeroplane. If it is assumed that
the aeroplane is trimmed so that no unbalance of pitching, yawing, or rolling moments exists, the
principle concern is for the forces acting on the aeroplane, i.e., lift, thrust, weight and drag.

Angle of Attack versus Airspeed

3. In order to achieve equilibrium in the vertical direction, the net lift must equal the aeroplane
weight. This is a contingency of steady, level flight or steady climbing and descending flight when the
flight path inclination is slight. A refinement of the basic lift equation defines the relationship of speed,
weight, lift coefficient, etc., for the condition of lift equal to weight.

L = W = CL ½ρ V(TAS)2S
½ ½
V(TAS) = __2 W__ = √ 2_ ___W/S___ (Multiply and divide by √ρO)
CL ρ S √ ρO CL (ρ / ρO)
½
V(TAS) = 1.278 _W / S_ (∵ ρ / ρO = σ)
CL σ
½
∴V(TAS) X √ σ = V(EAS) = 1.278 _W/S__
CL
Where W = Gross weight
S = Wing surface area
W/S = Wing loading
σ = Altitude density ratio
CL = Lift coefficient

4. From this relationship it is appreciated that a given configuration of aeroplane with a specific
wing loading, W/S, will achieve lift equal to weight at particular combinations of velocity, V, and lift
coefficient, CL. In steady flight, each equivalent airspeed demands a particular value of CL and each
FIS Book 1: Aerodynamics 414

value of CL demands a particular equivalent


airspeed to provide lift equal to weight. Fig
22-1 illustrates a typical lift curve for an
aeroplane and shows the relationship
between CL and angle of attack. For this
relationship, some specific value of α will
create a certain value of CL for any given
aerodynamic configuration.

5. For the conditions of steady flight


with a given aeroplane, each angle of attack
corresponds to a specific airspeed. Each
angle of attack produces a specific value of
CL and each value of CL requires a specific
value of equivalent airspeed to provide lift
equal to weight. Hence, angle of attack is
the primary control of airspeed in steady
flight. If an aeroplane is established in
steady, level flight at a particular airspeed,
any increase in angle of attack will result in
some reduced airspeed common to the Fig 22-1: Primary Control of Airspeed and
Altitude
increased CL. A decrease in angle of attack
will result in some increased airspeed common to the decreased CL. As a result of the change in
airspeed, the aeroplane may climb or descend if there is no change in power setting but the change in
airspeed was provided by the changes in angle of attack. The state of the aeroplane during the
change in speed will be some transient condition between the original and final steady state
conditions.

6. Primary control of airspeed in steady flight by angle of attack is an important principle. With
some configuration of aeroplanes, low speed flight will bring about a low level of longitudinal stick
force stability and possibility of low aeroplane static longitudinal stability. In such a case, the “feel” for
airspeed will be light and may not furnish a ready reference for easy control of the aeroplane. In
addition, the high angles of attack common to low speed flight are likely to provide large position
errors to the airspeed indicating systems. Thus, proper control of airspeed will be enhanced by good
“attitude” flying or when the visual reference field is poor by an angle of attack indicator.

7. Rate of Climb and Descent. In order for an aeroplane to achieve equilibrium at constant
altitude, lift must be equal to weight and thrust must be equal to drag. Steady, level flight requires
equilibrium in both the vertical and horizontal directions. For the case of climbing or descending flight
conditions, a component of weight is inclined along the flight path direction and equilibrium is
415 Application of AD to
Specific Problems of Flying

achieved when thrust is not equal to the drag. When the aeroplane is in a steady climb or descent, the
rate of climb is related by the following expression:

ROC = ( Pa - Pr ) Where: ROC = Rate of climb


W Pa = Propulsive power available
Pr = Power required for level flight
W = Gross weight

From this relationship it is appreciated that the rate of climb in steady flight is a direct function of the
difference between power available and power required. If a given aeroplane configuration is in lift-
equal-to-weight flight at some specific airspeed and altitude, there is a specific power required to
maintain these conditions. If the power available from the power plant is adjusted to equal the power
required, the rate of climb is zero (Pa - Pr = 0). This is illustrated in Fig 22-1 where the power
available is set equal to the power required at velocity (A). If the aeroplane were in steady level flight
at velocity (A), an increase in power available would create an excess of power which will cause a
rate of climb. Of course, if the speed were allowed to increase by a decreased angle of attack, the
increased power setting could simply maintain altitude at some higher airspeed. However, if the
original aerodynamic conditions are maintained, speed is maintained at (A) and an increased power
available results in a rate of climb. Also, a decrease in power available at point (A) will produce a
deficiency in power and result in a negative rate of climb (or a rate of descent). For this reasons, it is
apparent that power setting is the primary control of altitude in steady flight. There is the direct
correlation between the excess power (Pa-Pr), and the aeroplane rate of climb, ROC.

8. Flying Technique. Since the conditions of steady flight predominate during a majority of all
flying, the fundamentals of flying technique are the principles of steady flight :

(a) Angle of attack is the primary control of airspeed.


(b) Power setting is the primary control of altitude, i.e., rate of climb/descent.

With the exception of the transients conditions of flight which occur during manoeuvres and
aerobatics, the conditions of steady flight will be applicable during such steady flight conditions as
cruise, climb, descent, take-off, approach, landing etc. A clear understanding of these two principles
will develop good, safe flying techniques applicable to any sort of aeroplane.

9. The primary control of airspeed during steady flight conditions is the angle of attack. However,
changes in airspeed will necessitate changes in power setting to maintain altitude because of the
variation of power required with velocity. The primary control of altitude (rate of climb/descent) is the
power setting. If an aeroplane is being flown at a particular airspeed in level flight, an increase or
decrease in power setting will result in a rate of climb or descent at this airspeed. While the angle of
attack must be maintained to hold airspeed in steady flight, a change in power setting will necessitate
a change in attitude to accommodate the new flight path direction. These principles form the basis for
FIS Book 1: Aerodynamics 416

“attitude” flying technique, i.e., “attitude plus power equals performance”, and provide a background
for good instrument flying technique as well as good flying technique for all ordinary flying conditions.

10. One of the most important phases of flight is the landing approach and it is during this phase
of flight that the principle of steady flight is so applicable. If, during the landing approach, it is realized
that the aeroplane is below the desired glide path an increase in nose up attitude will not insure that
the aeroplane will climb to the desired glide path. In fact, an increase in nose up altitude may produce
a greater rate of descent and cause the aeroplane to sink more below the desired glide path. At a
given airspeed, only an increase in power setting can cause a rate climb (or lower rate of descent)
and an increase in nose up attitude without the appropriate power change only brings the aeroplane
to a lower speed.

11. Region of Reversed Command. The variation of power or thrust required with velocity
defines the power settings necessary to maintain steady level flight at various airspeeds. To simplify
the situation, a generality could be assumed that the aeroplane configuration and altitude define a
variation of power setting required (jet thrust required or prop power required) versus velocity. This
general variation of required power
setting versus velocity is illustrated at
Fig 22-2. The curve illustrates the fact
that at low speeds near the stall or
minimum control speed the power
setting required for steady level flight is
quite high. However, at low speeds, an
increase in speed reduces the required
power setting until some minimum value
is reached at the conditions for
maximum endurance. Increased speed
beyond the conditions for maximum
Fig 22-2: Region of Normal and
endurance will then increase the power Reversed Command
setting required for steady level flight.

Regions of Normal and Reversed Command

12. This typical variation of required power setting with speed allows a sort of terminology to be
assigned to specific regimes of velocity. Speeds greater than the speed for maximum endurance
require increasingly greater power setting to achieve steady, level flight. Since the normal command
of flight assumes “a higher power setting will achieve a greater speed”, the regime of flight speeds
greater than the speed for minimum power required setting is termed the “region of normal
command”. Obviously induced drag or induced power predominates in this regime to produce the
increased power setting required with increased velocity. Of course, the major items of aeroplane
417 Application of AD to
Specific Problems of Flying

flight performance take place in the region of normal command. Flight speed below the speed for
maximum endurance requires “power settings to increase with decrease in speed”. Since the increase
in required power setting with decreased velocity is contrary to the normal command of flight, this
regime of flight speeds between the speed for minimum required power setting and the stall speed (or
minimum control speed) is termed the “region of reversed command”. In this regime of flight, a
decrease in airspeed must be accompanied by an increased power setting in order to maintain steady
flight. One fact should be made clear about the region of reversed command. Flight in the region of
“reversed” command does not imply that a decreased power setting will bring about a higher airspeed
or an increased power setting will produce a lower airspeed. To be sure, the primary control of
airspeed is not the power setting. Flight in the region of reversed command only implies that a higher
airspeed will require a lower power setting and a lower airspeed will require a higher power setting to
hold altitude.

13. Because of the variation of required power setting throughout the range of flight speeds, it is
possible that one particular power setting may be capable of achieving steady, level flight at two
different airspeeds. As shown in Fig 22-2, one given power setting would meet the power
requirements and allow steady, level flight at both points 1 and 2. At speeds lower than point 1, a
deficiency of power would exist and a rate of descent would be incurred. Similarly, at speeds greater
than point 2, a deficiency of power would exist and the aeroplane would descend. The speed range
between points 1 and 2 would provide an excess of power and climbing flight would be produced.

Features of Flight in the Normal and Reversed Regions of Command

14. The region of reversed command is encountered primarily in the low speed phases of flight
during take-off and landing. The characteristics of flight in the region of normal command are
illustrated at point A of Fig 22-3. If the aeroplane is established in steady, level flight at point A, lift is
equal to weight and the power
available is set equal to the power
required. When the aeroplane is
disturbed to some airspeed slightly
greater than point A, a power
deficiency exists and, when the
aeroplane is disturbed to some
airspeed slightly lower than point A, a
‘power excess’ exists. This
relationship provides a tendency for
the aeroplane to return to the
equilibrium of point A and resume the
original flight condition following a Fig 22-3: Features of Flight in Normal and Region of
disturbance. Also, the static Reversed Command
FIS Book 1: Aerodynamics 418

longitudinal stability of the aeroplane tends to return the aeroplane to the original trimmed CL and
velocity corresponding to this CL. The phugoid usually has most satisfactory qualities at low values of
CL so the high speed of the region of normal command provides little tendency of the aeroplane’s
airspeed to vary or wander about.

15. With all factors considered, flight in the region of normal command is characterized by a
relatively strong tendency of the aeroplane to maintain the trim speed quite naturally. However, flight
in the region of normal command can lead to some unusual and erroneous impressions regarding
proper flying technique. For example, if the aeroplane is established at point A in steady level flight, a
controlled increase in airspeed without a change in power setting will create a deficiency of power and
cause the aeroplane to descend. Similarly, a controlled decrease in airspeed without a change in
power setting will create an excess of power and cause the aeroplane to climb. This fact, coupled with
the transient motion of the aeroplane when the angle of attack is changed rapidly, may lead to the
impression that rate of climb and descent can be controlled by changes in angle of attack. While such
is true in the region of normal command, for the conditions of steady flight, primary control of attitude
remains the power setting and the primary control of airspeed remains the angle of attack. The
impressions and habits that can be developed in the region of normal command can bring about
disastrous consequences in the region of reversed command.

16. The characteristics of flight in the region of reversed command are illustrated at point B of Fig
22-3. If the aeroplane is established in steady, level flight at point B, lift is equal to weight and the
power available is set to equal to power required. When the aeroplane is disturbed to some airspeed
slightly greater than point B, an excess of power exists and, when the aeroplane is disturbed to some
airspeed slightly lower than point B, a deficiency of power exists. This relationship is basically
unstable because the variation of excess power to either side of point B tends to magnify any original
disturbance. While the static longitudinal stability of the aeroplane tends to maintain the original
trimmed CL and airspeed corresponding to that CL, the phugoid usually has the least satisfactory
qualities at the high values of CL corresponding to low speed flight.

17. When all factors are considered, flight in the region of reversed command is characterised by
a relatively weak tendency of the aeroplane to maintain the trim speed naturally. In fact it is likely that
aeroplane will exhibit no inherent tendency to maintain the trim speed in this regime of flight. For this
reason, the pilot must give particular attention to precise control of airspeed when operating in the low
flight speeds of the region of reversed command.

18. While flight in the region of normal command may create doubt as to the primary control of
airspeed and altitude, operating in the region of reversed command should leave little doubt about
proper flying techniques. For example, if the aeroplane is established at point B (Fig 22-3) in level
flight, a controlled increase in airspeed (by reducing angle of attack) without changes in power setting
will create an excess of power at the higher airspeed and cause the aeroplane to climb. Also, a
419 Application of AD to
Specific Problems of Flying

controlled decrease in airspeed (by increasing angle of attack) without a change of power setting will
create a deficiency of power at the lower airspeed and cause the aeroplane to descend. This
relationship should leave little doubt as to the primary control of airspeed and altitude.

19. The transient conditions during change in airspeed in the region of reversed command are of
interest from the standpoint of landing flare characteristics. Suppose the aeroplane is in steady flight
at point B and the aeroplane angle of attack is increased to correspond with the value for the lower
airspeed of point C. The aeroplane would not instantaneously develop the lower speed and rate of
descent common to point C but would approach the conditions of point C through some transient
process depending on the aeroplane characteristics. If the aeroplane characteristics are low wing
loading, high L/D, and high lift curve slope, the increase in angle of attack at point B will produce a
transient motion in which curvature of the flight path demonstrates a definite flare. That is, the
increase in angle of attack creates a momentary rate of climb (or reduction of rate of descent) which
would be accompanied by a gradual loss of airspeed. Of course, the speed eventually decreases to
point C and the steady state rate of descent is achieved. If the aeroplane characteristics are high wing
loading, low L/D, and low lift curve slope, the increase in angle of attack at point B may produce a
transient motion in which the aeroplane does not flare. That is, the increase in angle of attack may
produce such rapid reduction of airspeed and increase in rate of descent that the aeroplane may be
incapable of a flaring flight path without an increase in power setting. Such characteristics may
necessitate special landing technique.

20. Operation in the region of reversed command does not imply that great control difficulty and
dangerous conditions will exist. However, flight in the region of reversed command does amplify any
errors of basic flying technique. Hence, proper flying technique and precise control of the aeroplane
are most necessary in the region of reversed command.

The Angle of Attack Indicator

21. The usual errors during the take-off and landing phases of flight involve improper control of
airspeed and altitude along some desired flight path. The angle of attack indicators assists the pilot
during the phase of take-off and landing and allows more consistent, precise control of the aeroplane.
Also many specific aerodynamic conditions exist at particular angles of attack for the aeroplane.
Generally, the conditions of stall, landing approach, take-off, range, endurance, etc., all occur at
specific values of lift coefficient and specific aeroplane angles of attack. Thus, an instrument to
indicate or relate aeroplane angle of attack would be a valuable reference to aid the pilot.

22. When the aeroplane is at high angles of attack it becomes difficult to provide accurate
indication of airspeed because of the possibility of large position errors. In fact, for low aspect ratio
aeroplane configuration at high angles of attack, it is possible to provide indications of angle of attack,
FIS Book 1: Aerodynamics 420

which are more accurate than indications of airspeed. As a result, an angle of attack indicator can be
of great utility at the high angles of attack.

23. A particular advantage of an angle of


attack indicator is that the indicator is not directly
affected by gross weight, bank angle, load
factor, velocity, or density altitude. The typical lift
curve of Fig 22-4 illustrates the variation of lift
coefficient, CL, with angle of attack α. When a
particular aerodynamic configuration is in
subsonic flight, each angle of attack produces a
particular value of lift coefficient. Of course, a
point of special interest on the lift curve is the
maximum lift coefficient, CL max. Angles of attack
greater that for CL max produce a decrease in lift
coefficient and constitute the stalled condition of
Fig 22-4: Variation of CL with α
flight. Since CL max occurs at a particular angle of
attack, any device to provide a stall warning
should be predicated on the function of this critical angle of attack. Under these conditions, stall of
aeroplane may take place at various airspeeds depending on gross weight, load factors etc., but
always the same angle of attack.

24. In order to reduce take-off and landing distances and minimize wear and tear on the wheels,
takeoff and landing will be accomplished at minimum practical speeds. The take-off and landing
speeds must provide sufficient margin above the stall speed (or minimum control speed) and are
usually specified at some fixed percentages of the stall speed. As such, take-off approach and landing
will be accomplished at specific values of lift coefficient and, thus, particular angles of attack. For
example, assume that point A on the lift curve is defined as the proper aerodynamic condition for the
landing approach. This condition exists as a particular lift coefficient and angle of attack for a specific
aerodynamic configuration. When the aeroplane is flown in a steady flight path at the prescribed angle
of attack, the resulting airspeed will be appropriate for the aeroplane gross weight. Any variation in
gross weight will simply alter the airspeed necessary to provide sufficient lift. The use of an angle of
attack indicator to maintain the recommended angle of attack will ensure that the aeroplane is
operated at the approach speed which is not too low or too high an airspeed. In addition to the use of
the angle of attack indicator during approach and landing, the instrument may be used as a principle
reference during take-off. The use of the angle of attack indicators to assume the proper take-off
angle of attack will prevent both over-rotation and excess take-off speed. Also, the angle of attack
indicator may be applicable to assist in control of the aeroplane for conditions of range, endurance,
manoeuvres, etc.
421 Application of AD to
Specific Problems of Flying

The Approach and Landing

25. The specific techniques necessary during the phase of approach and landing may vary
considerably between various types of aeroplanes and various operations. However, regardless of the
aeroplane type or operation, there are certain fundamental principles which will define the basic
techniques of flying during approach and landing. The specific procedures recommended for each
aeroplane type must be followed exactly to insure a consistent, safe landing technique.

26. The Approach. The approach must be conducted to provide a stabilized steady flight path
to the intended point of touchdown. The approach speed specified for an aeroplane must provide
sufficient margin above the stall speed or minimum control speed to allow satisfactory control and
adequate manoeuvrability. On the other hand, the approach speed must not be greatly in excess of
the touchdown speed or a large reduction in speed would be necessary prior to ground contact.
Generally, the approach speed will be from 10 to 30 percent above the stall speed depending on the
aeroplane type and the particular operation.

27. During the approach, the pilot must


attempt to maintain a smooth flight path and
prepare for the touchdown. A smooth, steady
approach to landing will minimize the
transient items of the flight path and provide
the pilot better opportunity to perceive and
orientate the aeroplane along the desired
flight path. Steep turns must be avoided at
the low speeds of the approach because of
the increase in drag and stall speed in the
turn, e.g. while turning on to base leg or for
Fig 22-5: Effect of Steep Turn on
final approach. Fig 22-5 illustrates the typical
Thrust Required
change in thrust required caused by a steep
turn. A steep turn may cause the aeroplane to stall or the large increase in induced drag may create
an excessive rate of descent. In either case, there may not be sufficient altitude to effect recovery. If
the aeroplane is not properly aligned on the final approach, it is certainly preferable to go around
rather than “press on regardless” and attempt to salvage a decent landing from a poor approach.

28. The proper coordination of the controls is an absolute necessity during the approach. In this
sense, due respect must be given to the primary control of airspeed and rate of descent for the
conditions of the steady approach. Thus, the proper angle of attack will produce the desired approach
airspeed. Too low an angle of attack will incur an excess speed while an excessive angle of attack will
produce a deficiency of speed and may cause stall or control problems. Once the proper airspeed and
angle of attack are attained the primary control of rate of descent during the steady approach will be
FIS Book 1: Aerodynamics 422

the power setting. For examples, if it is realised that the aeroplane is above the desired glide path, the
lowering of nose (a more nose-down attitude) without a decrease in power will result in a gain in
airspeed. On the other hand, if it is realised that the aeroplane is below the desired glide path, the
raising of nose (a more nose-up attitude) without an increase in power setting will simply allow the
aeroplane to fly more slowly and in the region of reversed command eventually produce a greater rate
of descent. For the conditions of steady flight, angle of attack is the primary control of airspeed and
power setting is the primary control of rate of climb and descent. This is especially true during the
steady approach to landing. Of course, the ability of the engine to produce rapid changes in thrust will
affect the specific technique to be used. If the engine is not capable of producing immediate controlled
changes in thrust, the operating technique must account for this deficiency. It is most desirable that
the power plant be capable of effecting rapid changes in thrust to allow precise control of the
aeroplane during approach.

Fig 22-6: Various Approach Paths


29. The type of approach path is an important factor since it affects the requirement of the flare,
the touchdown rate of descent, and, to some extent, the ability to control the point of touchdown.
Approach path A of Fig 22-6 depicts the steep, low power approach. Such a flight path generally
involves a low power setting, near idle conditions, and a high a rate of descent. Precise controls of the
aeroplane is difficult and an excess airspeed usually results from an approach path similar to A. A
sudden ‘go-around’ may be difficult because of the required engine acceleration and the high rate of
descent. In addition, the steep approach path with high rate of descent requires considerable flare to
reduce the rate of descent at touchdown. This extreme flare requirement will be difficult to execute
with consistency and will generally result in great variation in the speed, rate of descent and point of
touchdown.

30. Approach path C of Fig 22-6 typifies the long, shallow approach with too small an inclination
of the flight path. Such a flight path requires a relatively high power setting and a deficiency of
airspeed is a usual consequence. This extreme of an approach path is not desirable because it is
423 Application of AD to
Specific Problems of Flying

difficult to control the point of touchdown and the low speed may allow the aeroplane to settle
prematurely short of the intended landing touchdown.

31. Some approach path between the extremes of A and C must be selected e.g., flight path B.
The desirable approach path must not incur excessive speed and rate of descent or require excessive
flaring prior to touchdown. Also, some moderate power setting must be required which will allow
accurate control of the flight path and provide suitable ‘go-around’ characteristics. The approach flight
path cannot be too shallow for excessive power setting may be required and it may be difficult to
judge and control the point of touchdown.

The Landing Flare and Touchdown

32. The specific techniques of landing flare and touchdown will vary considerably between
various types of aeroplanes. The landing speed should be the lowest practical speed above the stall
or minimum control speed to reduce landing distance and arresting loads. Generally, the landing
speed will be from 5 to 25 % above the stall speed depending on the aeroplane type and the
particular operation.

33. The technique required for the landing will be determined in great part by the aerodynamic
characteristics of the aeroplane. If the aeroplane characteristics are low wing loading, high L/D, and
relatively high lift curve slope, the
aeroplane usually will have good
landing flare characteristics. If the
aeroplane characteristics are high
wing loading, low L/D, and relatively
low lift curve slope, the aeroplane
may not possess desirable flare
characteristics and landing
technique may require a minimum
of flare to touchdown. These
extremes are illustrated by the lift Fig 22-7: Typical Lift Curves
curves of Fig 22-7.

34. In preparation for the landing, several factors must be accounted for because of their effect on
landing distance, landing loads, and braking loads. These factors are:

(a) Landing Gross Weight. This must be considered because of its effect on landing
speed and landing loads. Since the landing is accomplished at a specific angle of attack or
margin above the stall speed, gross weight will define the landing speed. In addition, the
gross weight is an important factor in determining the landing distance and energy dissipating
requirements of the brakes. There will be a maximum design landing weight specified for
FIS Book 1: Aerodynamics 424

each aeroplane and this limitation must be respected because of critical landing loads, or
brake requirement. Any aircraft therefore will have a limiting touch down rate of descent
specified with the maximum landing weight and the principal landing load limitation will be
defined by the combination of gross weight and rate of descent at touch down.

(b) Surface Winds. The surface winds must be considered because of the large effect
of a headwind or tailwind on the landing distance. In the case of the crosswind, the
component of wind along the runway will be the effective headwind or tailwind velocity. Also,
the crosswind components across the runways will define certain requirements of lateral
control power. The aeroplane, which exhibits large dihedral effect at high lift coefficient, is
quite sensitive to crosswind and a limiting crosswind components will be defined for the
configuration.

(c) Pressure, Altitude and Temperature. These factors will affect the landing
distance because of the effect on the true air speed. Thus, pressure altitude and temperature
must be considered to define the density altitude.

(d) Runway Condition. The runway condition must be considered for its effect on
landing distances. Runway slope of ordinary values will ordinarily favour selection of a runway
for a favourable headwind at landing. The surface condition of the runway will determine
braking effectiveness and ice or water on the runway may result in an increase in the
minimum landing distance.

35. Preparation for the landing must therefore include determination of the landing distance of the
aeroplane and comparison with the runway length available. Use of the angle of attack indicator will
assist the pilot in effecting touchdown at the desired location with the proper airspeed. Of course, the
landing is not completed until the aeroplane is slowed to turn off the runway. Control of the aeroplane
must be maintained after the touchdown and proper technique must be used to decelerate the
aeroplane.

Typical Errors on Approach

36. There are many undesirable consequences when basic principles and specific procedures
are not followed during the approach and landing. Some of the typical errors involved in landing
accidents are outlined in the following discussion.

37. The Steep, Low Power Approach. This type of approach leads to an excessive rate of
descent and the possibility of a hard landing. This is particularly the case for the modern, low aspect
ratio, swept wing aeroplane configuration, which incurs very large induced drag at low speeds and
does not have very conventional flare characteristics for e.g. the Mig 21 aircraft. For this type of
aeroplane in a steep, low power approach, an increased angle of attack without a change of power
425 Application of AD to
Specific Problems of Flying

setting may not cause a reduction of rate of descent and may even increase the rate of descent at
touchdown. For this reason, a moderate stabilized approach is necessary and the principle changes in
rate of descent must be controlled by changes in power setting and principle changes in airspeed
must be controlled by changes in angle of attack.

38. Excessive Angle of Attack During Approach and Landing. An excessive angle of attack
during the approach and landing implies that the aeroplane is being operated at too low an airspeed.
Of course, excessive angle of attack may cause the aeroplane to stall or spin and the low altitude may
preclude recovery. Also, the low aspect ratio configuration at an excessively low airspeed will incur
very high induced drag and will necessitate a high power setting or otherwise incur an excessive rate
of descent. An additional problem is created by an excessive angle of angle for the aeroplane, which
exhibits a large dihedral effect at high lift coefficients. In this case, the aeroplane would be more
sensitive to crosswinds and adequate lateral control may not be available to affect a safe landing at a
critical value of crosswind. A high attitude will also restrict forward visibility in some aircraft.

39. Excess Airspeed. An excess airspeed at landing is just as undesirable as a deficiency of


airspeed. An excessive airspeed at landing will produce an undesirable increase in landing distance
and the energy to be dissipated by the brakes. In addition, the excess airspeed is a corollary of too
low an angle of attack and the aeroplane may touch down on the runway nose wheel first and cause
damage to the nose wheel or begin a proposing of the aeroplane. During a flare to landing, any
excess speed will be difficult to dissipate due to the reduction of drag due to ground effect. Thus, if the
aeroplane is held off with excess airspeed the aeroplane will “float” with the consequence of a
considerable runway distance used before touchdown and the subsequent difficulty in deceleration
and even barrier engagement.

40. A fundamental requirement for a good landing is a well planned and executed approach. The
possibility of errors during the landing process is minimized when the aeroplane is brought to the point
of touchdown with the proper glide path and airspeed. With the proper approach, there is no need for
drastic changes in the flight path, angle of attack, or power setting to accomplish touchdown at the
intended point on the runway.

The Takeoff

41. As in the case of landing, the specific technique necessary may vary greatly between various
types of aeroplanes and various operations but certain fundamental principle will be common to all
aeroplanes and all operations. The specific procedures recommended for each aeroplane type must
be followed exactly to insure a consistent, safe takeoff flying technique.

42. Takeoff Speed and Distance. The takeoff speed of any aeroplane is some minimum
practical airspeed which allows sufficient margin above stall and provides satisfactory control and
FIS Book 1: Aerodynamics 426

initial rate of climb. Depending on the aeroplane characteristics, the takeoff speed will be some value
5 to 25 % above the stall or minimum control speed. As such, the takeoff will be accomplished at a
certain value of lift coefficient and angle of attack specific to each aeroplane configuration. As a result,
the takeoff airspeed (EAS or CAS) of any specific aeroplane configuration is a function of the gross
weight at takeoff. Too low an airspeed at takeoff may cause stall, lack of adequate control, or poor
initial climb performance. An excess of speed at takeoff may provide better control and initial rate of
climb but the higher speed requires additional distance and may provide critical conditions for the
tyres.

43. The take off distance of an aeroplane is affected by many different factors other than
technique and, prior to take off, the take off distance must be determined and compared with the
runway length available. The principal factors affecting the takeoff distance are as follows:

(a) Gross Weight. The gross weight of the aeroplane has a considerable effect on
takeoff distance because it affects both takeoff speed and acceleration during takeoff roll.

(b) Surface Winds. The surface winds must be considered because of the powerful
effect of a headwind or tailwind on the takeoff distance. In the case of the crosswind, the
component of wind along the runway will be the effective head wind or tailwind velocity. In
addition, the component of wind across the runway will define certain requirements of lateral
control power and the limiting component wind must not be exceeded.

(c) Pressure Altitude and Temperature. These factors can cause a large effect on
takeoff distance, especially in the case of the turbo jet powered aeroplane. Density altitude
will determine the true airspeed at takeoff and can affect the takeoff acceleration by altering
the engine thrust. The effect of temperature alone is important in the case of the jet engine
aircraft since inlet air temperature will affect engine thrust. It should be noted that a typical
turbojet aeroplane may be approximately twice as sensitive to density altitude and five to ten
times as sensitive to temperature as a representative reciprocating engine powered
aeroplane.

(d) Specific Humidity. This must be accounted for in the case of the piston engine
powered aeroplane. High water vapour content in he air will cause a definite reduction in take-
off power and takeoff acceleration.

(e) Runway Condition. The runway condition will deserve consideration when the
takeoff acceleration is basically low. The runway slope must be compared carefully with the
surface winds because ordinary values of runway slope will usually favour choice of the
runway with headwind and upslope rather than down slope and tailwind. The surface
condition of the runway has little bearing on takeoff distance as long as the runway is a hard
surface.
427 Application of AD to
Specific Problems of Flying

44. Each of these factors must be accounted for and the takeoff distance properly computed for
the existing conditions. Since obstacle clearance distance is generally a function of the same factors
which affect take off distance, the obstacle clearance distance is usually related as some proportion of
the takeoff distance. Of course, the takeoff and obstacle clearance distances related by the pilots
notes will be obtained by the techniques and procedure outlined in the pilots notes.

Typical Errors on Take Off

45. The takeoff distance of an aeroplane should be computed for each takeoff. A most
inexcusable error would be to attempt takeoff from a runway of insufficient length. Familiarity with the
aeroplane performance data and proper accounting of weight, wind, altitude, temperature, etc., are
necessary parts of flying. Conditions of high gross weight, high pressure altitude and temperature,
and unfavourable winds create the extreme requirements of runway length, especially for the jet
engine aeroplane. Under these conditions, use of the aeroplane performance data is mandatory and
no guesswork can be tolerated.

46. Premature or Excess Pitch Rotation during Takeoff. One typical error of takeoff
technique is the premature or excess pitch rotation of the aeroplane. Premature or excess pitch
rotation of the aeroplane may seriously reduce the takeoff acceleration and increase the takeoff
distance. In addition, when the aeroplane is placed at an excessive angle of attack during takeoff, the
aeroplane may become airborne at too low a speed and the result may be a stall, lack of adequate
control (especially in a crosswind) or poor initial climb performance. In fact, there are certain low
aspect ratio configurations of aeroplanes, which, at an excessive angle of attack, will not fly out of
ground effect. Thus, over-rotation of the aeroplane during takeoff may hinder takeoff acceleration or
the initial climb. It is quite typical for an aeroplane to be placed at an excess angle of attack and
become airborne prematurely then settle back to the runway. When the proper angle of attack is
assumed, the aeroplane simply accelerates to the takeoff speed and becomes airborne with sufficient
initial rate of climb. In this sense, the appropriate rotation and takeoff speeds or an angle of attack
indicator must be used.

47. If the aeroplane is subject to a sudden pull-up or steep turn after becoming airborne, the
result may be a stall, spin, or reduction in initial rate of climb. The increased angle of attack may
exceed the critical angle of attack or the increase in induced drag may be quite large. For this reason,
any turns made immediately after takeoff must be well within the capabilities of the aeroplane.

48. In order to obviate some of the problems of a deficiency of airspeed at takeoff usual result can
be an excess of airspeed at takeoff. The principal effect of excess takeoff airspeed is the greater
takeoff distance. The general effect is that each 1% excess takeoff velocity incurs approximately 2%
additional takeoff distance. Thus, excess speed must be compared with the additional runway
required to produce the higher speed. In addition, the aircraft tyres may be subject to critical loads
FIS Book 1: Aerodynamics 428

when the aeroplane is at very high rolling speeds and speeds in excess of a basically high takeoff
speed may produce damage or failure of the tyres. As with the conditions of landing, excess velocity
or deficiency of velocity at takeoff is undesirable. The proper takeoff speeds and angle of attack must
be utilized to assure satisfactory takeoff performance.

GUSTS AND WIND SHEAR

49. The variation of wind velocity and direction throughout the atmosphere is important because
of its effect on the aerodynamic forces and moment on an aeroplane. As the aeroplane traverses this
variation of wind velocity and direction during flight, the changes in airflow direction and velocity
create changes in the aerodynamic forces and moments and produce a response of the aeroplane.
The variation of airflow velocity along a given direction exists with shear parallel to the flow direction.
Hence, the velocity gradients are often referred to as the “wind shear”.

50. The effect of the


vertical gust has important
effects on the aeroplane at
high speed because of the
possibility of damaging flight
loads. The mechanism of
vertical gust is illustrates in Fig 22-8: Effect of Vertical Gust
Fig 22-8 where the vertical
gust velocity is added vertical to the flight velocity to produce some resultant velocity. The principle
effect of the vertical gust is to produce a change in aeroplane angle of attack, e.g., a positive (up) gust
causes an increase in angle of attack while a negative (down) gust causes a decrease in angle of
attack. Change in angle of attack will effect a change in lift and therefore load factor. If some critical
combination of high gust intensity and high flight speed is encountered, the change in lift may be large
enough to cause structural damage, as explained in detail in the annexure to the chapter of
manoeuvres.

51. At low flight speeds during approach, landing and takeoff, the effect of the vertical gust is due
to the same mechanism of the change in angle of attack. However, at these low flight speeds, the
problem is of possible incipient stalling and sinking rather than overstress. When the aeroplane is at
high angle of attack, a further increase in angle of attack due to a gust may exceed the critical angle
of attack and cause an incipient stalling of the aeroplane. Also, a decrease in angle of attack due to a
gust will cause a loss of lift and allow the aeroplane to sink. For this reason, any deficiency of
airspeed will be quite critical when operating in gusty conditions.

52. The effect of the horizontal gust differs from the effect of the vertical gust in that the
immediate effect is a change of airspeed rather then a change in angle of attack. In this sense, the
429 Application of AD to
Specific Problems of Flying

horizontal gust is of little consequence in the


major aeroplane airloads and strength
limitation. Of greater significance is the
response of the aeroplane to horizontal
gusts and wind shear when operating at low
flight speeds. The possible conditions in
which an aeroplane may encounter
horizontal gusts and wind shear are
illustrated in Fig 22-9.

53. As the aeroplane traverses a shear


of wind direction, a change in headwind
component will exist. Also, a climbing or Fig 22-9: Horizontal and Vertical Variation of
descending aeroplane may traverse a shear Wind
of wind velocity, i.e., a wind profile in which the wind velocity varies with altitude.

54. The response of an aeroplane is much dependent upon the aeroplane characteristics but
certain basic effects are common to all aeroplanes. Suppose that an aeroplane is established in
steady, level flight with lift equal to weight, thrust equal to drag, and trimmed so there is no unbalance
of pitching, yawing, or rolling moment. If the aeroplane traverses a sharp wind shear equivalent to a
horizontal gust, the resulting change in airspeed will disturb such equilibrium. For example, if the
aeroplane encountered a sharp horizontal gust which reduces the airspeed by 20%, the new airspeed
(80% of the original value) produces lift and drag at the same angle of attack which are 64% of the
original value. The change in these aerodynamic forces would cause the aeroplane to accelerate in
the direction of resultant
unbalance of force (Fig
22-10). That is, the
aeroplane would
accelerate down and
forward until a new
equilibrium is achieved.
In addition, there would
be a change in pitching
which would produce a
response of the
Fig 22-10: Acceleration due to Unbalanced Forces
aeroplane in pitch.

55. The response of the aeroplane to a horizontal gust will differ according to the gust gradient
and aeroplane characteristics. Generally, if the aeroplane encounters a sharp wind shear which
reduces the airspeed, the aeroplane tends to sink and incur a loss of altitude before equilibrium
FIS Book 1: Aerodynamics 430

conditions are achieved. Similarly, if the aeroplane encountered a sharp wind shear which increases
the airspeed, the aeroplane tends to float and incur a gain of altitude before equilibrium conditions are
achieved.

56. Significant vertical and horizontal gusts may be due to the terrain or atmospheric conditions.
The proximity of an unstable front or thunderstorm activity in the vicinity of the airfield is likely to
create significant wind shear and gust activity at low altitude. During gusty conditions every effort must
be made for precise control of airspeed and flight path and any changes due to gusts must be
corrected by proper control action. Under extreme gusts conditions, it may be advisable to utilize
approach, landing, and takeoff speeds slightly greater than normal to provide margin for adequate
control.

57. Ground Effect. When an aeroplane in flight nears the ground (or water) surface, a change
occurs in the three dimensional flow pattern because the local airflow cannot have a vertical
component at the ground plane.
Thus, the ground plane will
furnish a restriction to the flow
and alter the wing upwash,
downwash, and tip vortices.
These general effects due to
the presence of the ground
Fig 22-11: An Aircraft Designed to Make
plane are referred to as “ground Use of Ground Effect
effect.”

Aerodynamic Influence of Ground Effect

58. While the aerodynamic


characteristics of the tail and fuselage are
altered by ground effects, the principal
effects due to proximity of the ground plane
are the changes in the aerodynamic
characteristic of the wing. As the wing
encounters ground effect and is maintained
at a constant lift coefficient, there is a
reduction in the upwash, downwash, and
the tip vortices. These effects are illustrated
by the Fig 22-12. As a result of the reduced
tip vortices, the wing in the presence of
ground effect will behave as if it were of a
Fig 22-12: Modification of Wing Tip Vortices Due
greater aspect ratio. In other words, the to Ground Effect
431 Application of AD to
Specific Problems of Flying

induced velocities due to the tip (or trailing) vortices will be reduced and the wing will incur smaller
values of induced drag coefficient, CD i, and induced angle of attack, αi, for any specific lift coefficient,
CL.

59. In order for ground


effect to be of a significant
magnitude, the wing must be
quite close to the ground plane.
Fig 22-13 illustrates one of the
direct results of ground effect
by the variation of induced drag
coefficient with wing height
above the ground plane for a Fig 22-13: Aeroplane in Ground Effect

representative unswept wing at


constant lift coefficient. Notice
that the wing must be quite
close to the ground for a
noticeable reduction in induced
drag (Fig 22-14). When the
wing is at a height equal to the
span (h/b = 1.0), the reduction
in induced drag is only 1.4%.
However, when the wing is at a
height equal to one-fourth the
span (h/b = 0.25), the reduction
in induced drag is 23.5% and,
when the wing is at a height Fig 22-14: Variation of Induced Drag with Height
Above Surface
equal to one-tenth the span
(h/b = 1.0), the reduction in induced drag is 47.6%. Thus, a large reduction in induced drag will take
place only when the wing is very close to the ground. Because of this variation, ground effect is most
usually recognized during the Lift off of takeoff or prior to touchdown on landing.

60. The reduction of the tip or trailing vortices due to ground effect alters the span wise lift
distribution and reduces the induced angle of attack. In this case, the wing will require a lower angle of
attack in ground effect to produce the same lift coefficient. This effect is illustrated by the lift curves of
Fig 22-15, which shows that the aeroplane in ground effect will develop a greater slope of the lift
curve. For the wing in ground effect, a lower angle of attack is necessary to produce the same lift
coefficient or, if a constant angle of attack is maintained, an increase in lift coefficient will result.
FIS Book 1: Aerodynamics 432

61. Fig 22-15 also illustrates


the manner in which ground
effect will alter the curve of thrust
required versus velocity. Since
induced drag predominates at
low speeds, the reduction of
induced drag due to ground effect
will cause the most significant
reduction of thrust required (Zero Fig 22-15: Variation of CL and Thrust in Ground Effect
Lift Drag plus induced drag) only at low speeds. At high speeds where Zero Lift Drag predominates,
the induced drag is but a small part of the total drag and ground effect causes no significant changes
in thrust required. Because ground effect involves the induced effects of aeroplane when in close
proximity to the ground, its effect are of greatest concern during the takeoff and landing. Normally,
these are the only phases of flight in which the aeroplane would be in close proximity to the ground.

Ground Effect on Specific Flight Conditions

62. The overall influence of ground effect is best realized by assuming that the aeroplane
descends into ground effect while maintaining a constant lift coefficient and, thus a constant dynamic
pressure and equivalent airspeed. As the aeroplane descends into ground effect, the following effect
will take place:

(a) Because of the reduced induced angle of attack and change in lift distribution, a
smaller wing angle of attack will be required to produce the same lift coefficients. If a constant
pitch altitude is maintained as ground effect is encountered, an increase in lift coefficient will
be incurred.

(b) The reduction is induced flow due to ground effect causes a significant reduction in
induced drag but causes no direct effect on zero lift drag. As a result of the reduction in
induced drag, the thrust required at low speeds will be reduced.

(c) The reduction in downwash due to ground effect will produce a change in longitudinal
stability and trim. Generally, the reduction in downwash at the horizontal tail increases the
contribution to static longitudinal stability. In addition, the reduction of downwash at the tail
usually requires a greater up elevator to trim the aeroplane at a specific lift coefficient. For the
conventional aeroplane configuration, encountering ground effect will produce a nose-down
change in pitching moment. Of course, the increase in stability and trim change associated
with ground effect provide a critical requirement of adequate longitudinal control power for
landing and takeoff.

(d) Due to the change in upwash, downwash, and tip vortices, there will be a change in
position error of the airspeed systems, associated with ground effect. In the majority of cases,
433 Application of AD to
Specific Problems of Flying

ground effect will cause an increase in the local pressure at the static source and produce a
lower indication of airspeed and altitude.

63. During the landing phase of flight, the effect of proximity to the ground plane must be
understood and appreciated. If the aeroplane is brought into ground effect with a constant angle of
attack, the aeroplane will experience an increase in lift coefficient and reduction in thrust required.
Hence, a “floating” sensation may be experienced. Because of the reduced drag and power-off
deceleration in ground effect, any excess speed at the point of flare may incur a considerable “float”
distance. As the aeroplane nears the point of touchdown on the approach, ground effect will be most
realized at altitudes less than the wingspan. During the final phases of approach as the aeroplane
nears the ground, reduced power setting is necessary or the reduced thrust required would allow the
aeroplane to climb above the desired glide path.

64. An additional factor to consider is the aerodynamic drag of the aeroplane during the landing
roll. Because of the reduced induced drag when in ground effect, aerodynamic braking will be of
greatest significance only when partial stalling of the wing can be accomplished. The reduced drag
when in ground effect accounts for the fact that the brakes are the most effective source of
deceleration for the majority of aeroplane configurations.

65. During the takeoff phase of flight ground effect produces some important relationships. Of
course, the aeroplane leaving ground effect encounters just the reverse of the aeroplane entering
ground effect, i.e., the aeroplane leaving ground effect will:

(a) Require an increase in angle of attack to maintain the same lift coefficient.
(b) Experience an increase in induced drag and thrust required.
(c) Experience a decrease in stability and a nose-up change in moment.
(d) Usually a reduction in static source pressure and increase in indicated airspeed.

These general effects should point out the possible danger in attempting takeoff prior to achieving the
recommended takeoff speed. Due to the reduced drag in ground effect the aeroplane may seem
capable of takeoff below the recommended speed. However, as the aeroplane rises out of ground
effect with a deficiency of speed, the greater induced drag may produce marginal initial climb
performance. In the extreme conditions such as high gross weight, high density altitude, and high
temperature, a deficiency of airspeed at takeoff may permit the aeroplane to become airborne but be
incapable of flying out of ground effect. In this case, the aeroplane may become airborne initially with
a deficiency of speed, but later settle back to the runway. It is imperative that no attempt be made to
force the aeroplane to become airborne with a deficiency of speed. The recommended takeoff speed
is necessary to provide adequate initial climb performance.
FIS Book 1: Aerodynamics 434

Interference Between Aeroplanes in Flight

66. During close formation flying and in flight refuelling, aeroplanes in proximity to one another
will produce a mutual interference of the flow patterns and alter the aerodynamic characteristics of
each aeroplane. The principal effects of this interference must be appreciated since certain factors
due to the mutual interference may enhance the possibility of a collision.

67. One example of interference


between aeroplanes in flight is shown first in
Fig 22-16 with the effect of lateral separation
of two aeroplanes flying in line abreast. A
plane of symmetry would exist halfway
between two identical aeroplanes and would
furnish a boundary of flow across which
there would be no lateral components of
flow. As the two aeroplane wing tips are in
proximity, the effect is to reduce the strength
of the tip or trailing vortices and reduce the Fig 22-16: Lift Distribution about the Plane of
induced velocities in the vicinity of wing tip. Symmetry
Thus, each aeroplane will experience a local
increase in the lift distribution as the tip vortices are reduced and a rolling moment is developed which
tends to roll each aeroplane away from the other. This disturbance may provide the possibility of
collision if other aeroplanes are in the vicinity and there is delay in control correction or over control. If
the wing tips are displaced in a fore-and-aft direction, the same effect exists but generally it is of a
lower magnitude.

68. The magnitude of the interference effect due to lateral separation of the wing tips depends on
the proximity of the wing tips and the induced flow. This implies that the interference would be
greatest when the tips are very close and the aeroplanes are operating at high lift coefficients. An
interesting ramification of this effect is that several aeroplanes in line abreast with the wing tips quite
close will experience a reduction in induced drag.

69. An indirect form of interference can be encountered from the vortex systems created by a
preceding aeroplane along the intended flight path. The vortex sheet rolls up a considerable distance
behind an aeroplane and creates considerable turbulence for any closely following aeroplane. This
wake can prove troublesome if aeroplanes taking off and landing are not provided adequate
separation. The rolled-up vortex sheet will be strongest when the preceding aeroplane is large, high
gross weight, and operating at high lift coefficient. At times this turbulence may be falsely attributed to
prop-wash or jet-wash.
435 Application of AD to
Specific Problems of Flying

70. Another important


form of direct interference
is common when the two
aeroplanes are in a trail
position and stepped down
i.e. line astern position. As
shown in Fig 22-17, the
Fig 22-17: Interference between Aeroplane in Line Astern
single aeroplane in flight
develops up wash ahead of the wing and downwash behind and any restriction accorded the flow can
alter the distribution and magnitude of the up wash and downwash. When the trailing aeroplane is in
close proximity aft and below the leading aeroplane a mutual interference takes places between the
two aeroplanes. The leading aeroplane above will experience an effect which would be somewhat
similar to encountering ground effect, i.e., a reduction in induced drag, a reduction in downwash at the
tail and a change in pitching moment nose down. The trailing aeroplane below will experience an
effect, which is generally the opposite of the aeroplane above. In other words, the aeroplane below
will experience an increase in induced drag, an increase in downwash at the tail, and change in
pitching moment nose up. Thus, when the aeroplanes are in close proximity, a definite collision
possibility exists because of the trim change experienced by each aeroplane. The magnitude of the
trim change is greatest when the aeroplanes are operating at high lift coefficients, e.g., low speed
flight, and when the aeroplanes are in close proximity.

71. In formation flying, this sort of interference must be appreciated and anticipated. In crossing
under another aeroplane, care must be taken to anticipate the trim change and adequate clearance
must be maintained, otherwise a collision may result. The pilot of the leading aircraft will know of the
presence of the trailing aeroplane by the trim change experienced. Obviously, some anticipation is
necessary and adequate separation is necessary to prevent a disturbing magnitude of the trim
change. In a close diamond formation the leader will be able to “feel” the presence of the slot man
even though the aeroplane is not within view. Obviously, the slot man will have a difficult job during
formation manoeuvres because of the unstable trim changes and greater power changes required to
hold position.

72. A common collision problem is the case of an aeroplane with a malfunctioning landing gear. If
another aeroplane is called to inspect the malfunctioning landing gear, great care must be taken to
maintain adequate separation and preserve orientation. Many instances such as this have resulted in
a collision when the pilot of the trailing aeroplane became disoriented and did not maintain adequate
separation.

73. During in-flight refuelling, essentially the same problems of interference exist. As the receiver
approaches the tanker from behind and below, the receiver will encounter the downwash from the
tanker and require a slight, gradual increase in power and pitch attitude to continue approach to the
FIS Book 1: Aerodynamics 436

receiving position. While the receiver may not be visible to the pilot of the tanker, he will anticipate the
receiver coming into position by the slight reduction in power required and nose down change in
pitching moment. Adequate clearance and proper position must be maintained by the pilot of the
receiver for a collision possibility is enhanced by the relative positions of the aeroplanes. A hazardous
condition exists if the pilot of the receiver has excessive speed and runs under the tanker in close
proximity. The trim change experienced by both aeroplanes may be large and unexpected and it may
be difficult to avoid a collision.

74. In addition to the forms of interference previously mentioned, there exists the possibility of
strong interference between aeroplanes in supersonic flight. In this case, the shockwaves from one
aeroplane may strongly affect the pressure distribution and rolling, yawing, and pitching moments of
an adjacent aeroplane. It is difficult to express general relationship of the effect except that magnitude
of the effects will be greatest when in close proximity at low altitude and high α. Generally, the trailing
aeroplane will be most affected.

Braking Performance

75. For the majority of aeroplane configurations and runway conditions, the aeroplane brakes
furnish the most powerful means of deceleration. While specific techniques of braking are required for
specific situations, there are various fundamentals which are common to all conditions.

76. Solid friction is the resistance to relative motion of two surfaces in contact. When relative
motion exists between the surfaces, the resistance to relative motion is termed “kinetic” or “sliding”
friction. When no relative motion exists between the surfaces, the resistance to the impending relative
motion is termed “static” friction. The minute discontinuities of the surface in contact are able to mate
quite closely when relative motion impends rather than exists, so static friction will generally exceed
kinetic friction. The magnitude of the friction force between two surfaces will depend in great part on
the type of surfaces in contact and the magnitude of force pressing the surfaces together. A
convenient method of relating the friction characteristics of surface in contact is a proportion of the
friction force to the normal (or perpendicular) force pressing the surfaces together. This proportion
defines the coefficient of friction, µ.

µ = F
N
Where: µ = Coefficient of friction (mu).
F = Friction force.
N = Normal force.

The coefficient of friction of tyres on a runway surface is a function of many factors. Runway surface
condition, rubber composition, tread, inflating pressure, surface friction shearing stress, relative slip
speed, etc., all are factors which affect the coefficient of fiction. When the tyre is rolling along the
437 Application of AD to
Specific Problems of Flying

runway without the use of brakes, the friction force resulting is simple rolling resistance. The
coefficient of rolling friction is of an approximate magnitude of 0.015 to 0.030 for dry, hard runway
surface.

77. The application of brakes supplies a torque to


the wheel, which tends to retard wheel rotation,
however this is balanced by the increase in friction
force which produces a driving or rolling torque (Fig
22-18). When the braking torque is equal to the rolling
torque, the wheel experiences no acceleration in
rotation and the equilibrium of constant rotational
speed is maintained. Thus, the application of brake
develops a retarding torque and causes an increase Fig 22-18: Forces on a Wheel
in friction between the tire and runway surface. A
common problem of braking technique is application of excessive brake pressure, which creates a
braking torque greater than the maximum possible rolling torque. In this case, the wheel loses
rotational speed and decelerates until the wheel is stationary and the result is a locked wheel with the
tyre surface subject to a full slip condition.

78. The effect of slip velocity on the coefficients of friction is illustrated by the graph of Fig 22-19.
The conditions of zero slip corresponds to the rolling wheel without brake application while the
condition of full, 100% slip corresponds to
the locked wheel where the relative velocity
between the tyre surface and the runway
equals the actual velocity. With the
application of brakes, the coefficient of
friction increases but incurs a small but
measurable apparent slip. Continued
increase in friction coefficient is obtained
until some maximum is achieved then
decreases as the slip increases and
approaching the 100% slip conditions.
Actually, the peak value of coefficient of
friction occurs at an incipient skid condition
and the relative slip apparent at this point
consists primarily of elastic shearing
deflection of the tyre structure. Fig 22-19: Effect of Slip Velocity on µ

79. When the runway surface is dry, brush-finished concrete, the maximum value for the
coefficient of friction for most aircraft tyres is of the order of 0.6 to 0.8. Many factors can determine
FIS Book 1: Aerodynamics 438

small differences in this peak value of friction coefficient for any surface conditions. For example, a
soft gum rubber composition can develop a very high value of coefficient of friction but only for low
values of surface shearing stress. At high values of surface shearing stress, the soft gum rubber will
shear or scrub off before high values of friction coefficient are developed. The higher strength
compounds used in the production of aircraft tyres produce greater resistance to surface shear and
scrubbing but the harder rubber has lower intrinsic friction coefficients. The high performance
aeroplane tyres will be of relatively hard rubber and will operate at or near the rated load capacities.
As a result, there will be little deference between the peak values of friction coefficient for the dry,
hard surface runway for the majority of aircraft tires.

80. If high traction on dry surfaces were the only consideration in the design of tires, the result
would be a soft rubber tire of extreme width to create a large footprint and reduce surface shearing
stresses, e.g., driving tires on a drag racer. However, such a tire has many other characteristics,
which are undesirable such as high roiling friction, large size, poor side force characteristics, etc.

81. When the runway has water or ice on the surface, the maximum value for the coefficient of
friction is reduced greatly below the value obtained for the dry runway condition. When water is on the
surface, the tread design becomes of greater importance to maintain contact between the rubber and
the runway and prevent a film of water from lubricating the surfaces. When the rainfall is light, the
peak value for friction coefficient is of the order of 0.5. With heavy rainfall it is more likely that
sufficient water will stand to form a liquid film between the tire and the runway. In this case, the peak
coefficient of friction rarely exceeds 0.3. In some extreme conditions, the tyre may simply plane along
the water without contact of the runway and the coefficient of friction is much lower than 0.3. Smooth
clear ice on the runway will causes extremely low values for the coefficient of friction. In such a
condition, the peak value for the coefficient of friction may be of the order of 0.2 or 0.15.

82. Note that immediately past the incipient skidding condition the coefficient of friction decreases
with increased slip speed, especially for the wet or icy runway conditions. Thus, once skid begins, a
reduction in friction force and rolling torque must be met with a reduction in braking torque, otherwise
the wheel will decelerate and lock. This is an important factor to consider in braking technique
because the skidding tire surface on the locked wheel produces considerably less retarding force than
when at the incipient skid condition which causes the peak coefficient of friction. If the wheel locks
from excessive braking, the sliding tire surface produces less than the maximum retarding force. Stop
distance will increase and it may be difficult if not impossible to control the aeroplane when full slip is
developed. In addition, at high rolling velocities on the dry surface runway, the immediate problem of a
skidding tyre is not necessarily the loss of retarding force but the imminence of tire failure. The pilot
must insure that the application of brakes does not produce some excessive braking torque, which is
greater than the maximum rolling torque, and particular care must be taken when the runway
conditions produce low values of friction coefficient and when the normal force on the braking
439 Application of AD to
Specific Problems of Flying

surfaces is small. When it is difficult to perceive or distinguish a skidding condition, the value of an
antiskid or automatic braking system will be appreciated.

83. Braking Technique. It must be clearly distinguished that the techniques for minimum stop
distance may differ greatly from the technique required to minimize wear and tear on the tyres and
brakes. For the majority of aeroplane configuration, brakes will provide the most important source of
deceleration for all but the most severe of icy runway conditions. Of course, aerodynamic drag is very
durable and should be utilized to decelerate the aeroplane if the runway is long enough and the drag
high enough. Aerodynamic drag will be of importance only for the initial 20 to 30% of speed reduction
from the point of touchdown. At speeds less than 60 to 70% of the landing speed, aerodynamic drag
is of little consequence and brakes will be the principle source of deceleration regardless of the
runway surface. For the condition of minimum landing distance aerodynamic drag will be principal
source of deceleration only for the initial portion of landing roll for very high drag configurations on
very poor runway conditions. These cases are quite limited so considerable importance must be
assigned to proper use of the brakes to produce maximum effectiveness.

84. In order to provide the maximum


possible retarding force, effort must be
directed to produce the maximum normal
force on the braking surface (Figure 22-
20). The pilot will be able to influence the
normal force on the braking surface during
the initial part of the landing roll when
dynamic pressure is large and Fig 20-20: Braking Performance

aerodynamic forces and moments are of consequence. During this portion of the landing roll the pilot
can control the aeroplane lift and the distribution of normal force to the landing gears.

85. First to consider is that any positive lift will support a part of the aeroplane weight and reduce
the normal force on the landing gear. For the purpose of braking friction, it would be of advantage to
create negative lift but this is not the usual capability of the aeroplane with the tricycle landing gear.
Since the aeroplane lift may be considerable immediately after landing, retraction of flaps or extension
of spoiler immediately after touchdown will reduce the wing lift and increase the normal force on the
landing gear. With the retraction of flaps, the reduced drag is more than compensated for by the
increased braking friction force afforded by the increased normal force on the braking surfaces.

86. A second factor to control braking effectiveness is the distribution of normal force to the
landing gear surfaces. The nose wheel of the tricycle landing gear configuration usually has no brakes
and any normal force distributed to this wheel is useful only for producing side force for control of the
aeroplane. Under conditions of deceleration, the nose down pitching moment created by the friction
force and the inertia force tends to transfer a significant amount of normal force to nose wheel where
FIS Book 1: Aerodynamics 440

it is unavailable to assist in creating friction force. For the instant after landing touchdown, the pilot
may control this condition to some extent and regain or increase the normal force on the main wheels.
After touchdown, the nose is lowered until the nose wheel contacts the runway then brakes are
applied while the stick is eased back without lifting the nose wheel back off the runway. The effect is
to minimise the normal force on the nose wheel and increase the normal force on braking surfaces.
While the principal effect is to transfer normal force to the main wheels, there may be a significant
increase in normal force due to a reduction in net lift, i.e., tail download is noticeable. This reduction in
net lift tends to be particular to tailless or short coupled aeroplane configurations.

87. The combined effect of flap retraction and aft stick is a significant increase in braking friction
force. The flaps should not be retracted while still airborne and aft stick should be used just enough
without lifting the nose wheel off the runway. These techniques are to no avail if proper use of the
brakes does not produce the maximum coefficient of friction. The incipient skid condition will produce
the maximum coefficients of friction but this is difficult to recognize and maintain without an antiskid
system. Judicious use of the brakes is necessary to obtain the peak coefficient of friction but not
develop a skid or locked wheel, which could cause tire failure, loss of control, or considerable
reduction in the friction coefficient.

88. The capacity of the brakes must be sufficient to create adequate braking torque and produce
the high coefficients of friction. In addition, the brakes must be capable of withstanding the heat
generated without fading or losing effectiveness. The most critical requirements of the brakes occur
during landing at the maximum allowable landing weight.

89. Typical Errors of Braking Technique. Errors in braking technique are usually coincident
with errors of other sorts. For example, if the pilot lands an aeroplane with excessive airspeed, poor
braking technique could accompany the original error to produce an unsafe situation. One common
error of braking technique is the application of braking torque in excess of the maximum possible
rolling torque. The result will be that the wheel decelerates and locks and the skid reduces the
coefficient of friction, lowers the capability for side force, and enhances the possibility of tire failure. If
maximum braking is necessary, caution must be used to modulate the braking torque to prevent
locking the wheel and causing a skid. On the other hand, maximum coefficient of friction is obtained at
the incipient skidding condition so sufficient brake torque must be applied to produce maximum
friction force. Intermittent braking serves no useful purpose when the objective is maximum
deceleration because the periods between brake applications produce only slight or negligible cooling.
Brake should be applied smoothly and braking torque modulated at or near the peak value to insure
that skid does not develop.

90. One of the important factors affecting the landing roll distance is landing touchdown speed.
Any excess velocity at landing causes a large increase in the minimum stop distance and it is
necessary that the pilot control the landing precisely so as to land at the appropriate speed. When
441 Application of AD to
Specific Problems of Flying

landing on the dry, hard surface runway of adequate length, a tendency is to take advantage of any
excess runway and allow the aeroplane to touchdown with excess speed. Of course, such errors in
technique cannot be tolerated and the pilot must strive for precision in all landings. Immediately after
touchdown, the aeroplane lift may be considerable and the normal force on the braking surfaces quite
low. Thus, if excessive braking torque is applied, the wheel may lock easily at high speeds and tire
failure may take place suddenly.

91. Landing on a wet or icy runway requires judicious use of the brakes because of the reduction
in the maximum coefficient of friction. Because of reduction in the maximum attainable value of the
coefficient of friction, the pilot must anticipate an increase in the minimum landing distance above that
applicable for the dry runway conditions. When there is considerable water or ice on the runway, an
increase in landing distance of the order of 40 to 100% must be expected for similar conditions of
gross weight, density altitude, wind, etc. Unfortunately, the conditions likely to produce poor braking
action also will cause high idle thrust of the turbojet engine and the extreme case (smooth, glazed ice
or heavy rain) may dictate shutting down the engine to effect a reasonable stopping distance.

AQUAPLANING

92. Description. As an un-braked pneumatic tyre rolls on a water logged runway, it contacts
and displaces the stationary water. The resulting change in momentum of the water creates
aerodynamic pressure, which reacts on the runway and tyre surfaces. This hydrodynamic pressure
tends to increase as the square of the ground speed. As the ground speed increases the inertia of the
water tends to retard its escape from the tyre/ground contact area and a wedge of water forms, which
begins to lift the tyre from the ground. Further increase of ground speed increases the hydrodynamic
lift until the lift developed equals the weight supported by the wheel. Any further increase in ground
speed will result in the tyre being lifted off the runway surface. This is the condition of total
aquaplaning.

93. There are three types aquaplaning :

(a) Dynamic Aquaplaning. In total dynamic aquaplaning, water standing on the


runway exerts pressure between the tires and the runway. The tires are lifted and are not in
contact with the runway surface. The rolling coefficient of friction (and also brake
effectiveness) is reduced to nearly nothing. This means also that steering is not effective. A
strong crosswind can cause added problems of control. Dynamic aquaplaning starts at high
speeds and in standing water on the runway.

A thumb rule for predicting the minimum dynamic aquaplaning speed (knots) is 8.6 X
√(tyre pressure, psi) or 60 X √(tyre pressure, Kg / cm2) and the resultant speed in
kmph. At a tyre pressure of 25 Ibs the expected minimum dynamic aquaplaning
FIS Book 1: Aerodynamics 442

speed is 43 k (8.6 X √25 = 8.6 X 5 = 43). You could expect problems above this
speed.

(b) Viscous Aquaplaning. When the runway has painted areas or rudder deposit that
make it smooth, the tire can’t fully displace the moisture film. You can feel this effect in driving
a car when your car slips momentarily as you cross an extra thick painted highway centre line
covered with rain or dew. While a large area of the runway or taxiway is involved, you could
lose steering and braking ability. This can occur at a much lower speed than dynamic
aquaplaning.

(c) Reverted Rubber Aquaplaning. Suppose you are touching down on a wet runway
and (wrongly) apply brakes immediate after touchdown. The aeroplane starts dynamic
aquaplaning because the brakes are locked. The aeroplane slows down, the dynamic
aquaplaning decreases, and the locked tyres heat up because of added friction. A layer of
steam occurs between the tires and the runway and the rubber melts. This prevents water
dispersal because the tire is riding on a layer of steam and molten rubber. This is the worst of
the aquaplaning variations because it can happen down to zero speed. Locking the brakes for
prolonged periods causes reverted (melted) rubber aquaplaning. Some of the newer runways
are grooved to cut down on aquaplaning effects, but you should be ready for it anytime you
are taking off or landing on a wet runway. Think of braking or directional controls problems,
and avoid excessive use of the rudder pedals or brakes. You might assume that the liquid
water will have the friction properties of ice(which it will, under the conditions just mentioned).

Spin-down of Unbraked Wheel

94. When aquaplaning occurs, perhaps the most unexpected feature is the condition in which free
rolling (unbraked) wheel slows down or stop completely. This wheel spin-down arises from
hydrodynamic lift effects which combine to provide a total wheel spin down moment in excess of the
wheel spin-up moment caused by all tyre drag sources.

Factors Affecting Aquaplaning

95. Effect of Weight and tyre Pressure. Changing the weight on the tyre appears to have little
effect on the speed at which total aquaplaning occurs. As the weight on the tyre changes so the
contact area changes and the radio of weight to area remains constant; this is essentially due to the
tyre inflation pressure. It is this pressure that the hydrodynamic lift pressure must equal over the entire
contact area before total aquaplaning occurs. A simple expression based on hydrodynamic lift theory
can be used to predict total aquaplaning speed ; this is :

V = 8.6 X √P Where V = total aquaplaning speed (in knots)


P = Tyre pressure (in psi)
443 Application of AD to
Specific Problems of Flying

It must be remembered however, that aquaplaning is a progressive phenomenon and significant


losses in braking and directional control will occur below this speed. Once aquaplaning has started,
the wheel spins down and stops fairly rapidly. However, if ground speed is reduced, the time for the
wheels to spin up again is much longer and aquaplaning will persist to a lower speed than that given
by the above mentioned equation.

96. Others Factors. The actual depth of fluid required is not well defined and varies
considerably with tyre tread design and runway surface and smoothness. Aquaplaning of smooth
tyres can occur on a smooth runway with an average fluid depth of 0.15 in, ribbed tyres on a rough
surface will require a depth far in excess of this.

97. In aquaplaning conditions at ground speeds above tyre aquaplaning speed on take-off, the
coefficient of ground friction will reduce to nil. Additionally, the wheels may spin down rapidly and
stop. The futility of applying brakes in these conditions is obvious, brake effectiveness will be zero and
there be no directional control.
FIS Book 1: Aerodynamics 444
445

BASIC PRINCIPLES OF HELICOPTER FLIGHT

CHAPTER 23

BASICS OF ROTORDYNAMICS

Introduction

1. The same basic laws govern the flight of both fixed and rotary wing aircraft and, equally, both
types of aircraft share the same fundamental problem; namely that the aircraft is heavier than air and
must, therefore, produce an aerodynamic lifting force to overcome the weight of the aircraft before it
can leave the ground. In both types of aircraft the lifting force is obtained from the aerodynamic
reaction resulting from a flow of air over an aerofoil section. The important difference lies in the
relationship of the aerofoil to the fuselage. In a fixed-wing aircraft, the aerofoil is fixed to the fuselage
as a wing while in a helicopter, the aerofoil has been removed from the fuselage and attached to a
centre shaft, which, by one means or another, is given a rotational velocity.

2. Helicopters have rotating wings, which are engine-driven in normal flight. The rotor provides
both lift and horizontal thrust.

Rotor Systems

3. Helicopters may be single


or multi-rotor, each rotor having
several blades, usually varying
from two to six in number. The
rotor blades are attached by a
rotor head to a rotor shaft which
extends approximately vertically
from the fuselage. They form the
rotor, which turns independently
through the rotor shaft, (Fig 23-1).
The axis of rotation is the axis
through the rotor head and about
which the rotor blades are
permitted to rotate. The plane of
rotation is at right angles to the Fig 23-1: Rotor Head Arrangement
axis of rotation at the head of the
main rotor shaft (Fig 23-2).
FIS Book 1: Helicopter Flight 446

Definitions

4. General. Certain terms are used in explanation of helicopter aerodynamics which differs
from those used in fixed-wing flight. It is therefore necessary to define these terms before considering
the principles of helicopter flight.

5. Shaft Axis. The shaft axis is


the axis through the main rotor shaft,
about which the blades are permitted
to rotate (Fig 23-2).

6. Axis of Rotation. The axis


of rotation is the axis through the head
of the main rotor shaft about which the
blades actually rotate. Under ideal
conditions, the axis of rotation will
coincide with the shaft axis. However,
this will not always be the case since
Fig 23-2: Terms Related to Rotor
the rotor disc is permitted to tilt under
certain conditions of flight (Fig 23-2).

7. Plane of Rotation. The plane of rotation is at right-angles to the axis of rotation at the
head of the main rotor shaft, and is parallel to the tip path plane (Fig 23- 2).

8. Tip path Plane. The tip path plane is the plane described by blade tips during rotation
and is at right-angles to the axis of rotation. The area contained within this path is referred to as the
rotor disc. (Fig 23- 2)

9. Lift. The lift produced from the


wing of a fixed wing aircraft results from
a combination of many factors and is
commonly expressed in the formula CL
½ p V² S. Lift from a helicopter rotor
blade can generally be expressed in the
same terms but because the blade
moves independently of the fuselage,
the velocity (V²) in hovering conditions
(in still air) is purely the result of the
Fig 23-3: Lift / TR on the Blade
blade rotation (Fig 23-3)
447 Basics of Rotordynamics

10. Blade Pitch. The angle


between the blade element chord
and the plane of rotation is called
the blade pitch angle (Fig 23-4).

11. If the blade had a


constant value of pitch throughout
its length, blade-loading problems
would arise because each section Fig 23-4: Blade Pitch Angle

of the blade would have a different


rotational velocity and would
therefore produce a different value
of lift. As lift is proportional to V2
each time the speed is doubled,
the lift would quadruple and the lift
pattern or loading, on the blade
would be as shown by the red line
in Fig 23-5. To avoid this
Fig 23-5: Lift Variation with Wash In / Out
considerable load variation, lift
must be increased at the root end and decreased at the tip; the blade is therefore either tapered,
given a washout, or a combination of both. Lift from the blade will still have its greatest value near the
tip but its distribution along the blade will be more uniform (the green line in Fig 23-5). The Sycamore
had taper and wash-out, being -10 at the tip and +3 ½0 at the root while the Chetak rotor blade has a
wash-out of 6030’.

12. Relative Airflow. If a rotor blade is moved horizontally through a column of air, the effect
will be to displace some of the air downwards. If a number of rotor blades are travelling along the
same path in rapid succession with a three-blades rotor systems rotating at 240 rotor rpm (Rrpm), a
blade will be passing a
given point every twelfth of
a second, then the column
of still air will eventually
become a column of
descending air (Fig 23-6).
This downward motion of
Fig 23-6: Induced flow
the air is known as
induced flow. The direction of the airflow relative to the blade is the resultant of the blade’s horizontal
travel through the air and the induced flow.
FIS Book 1: Helicopter Flight 448

13. Rotor Thrust and Rotor Drag. If the force acting on the aerofoil (total reaction) is split into
the components of lift and induced drag, then lift, which is at right-angles to the relative airflow (RAF),
is providing a force in direct
opposition to the weight, as in the
case of the fixed-wing aircraft. The
lifting component of the total
reaction must therefore be that
part of it which is acting along the
axis of rotation. This component is
known as rotor thrust. The other
component of total reaction will be
in the blade’s plane of rotation and
is known as rotor drag (Fig.23-7). Fig 23-7: Forces on an Aerofoil

14. Total Rotor Thrust. Provided each blade produces equal rotor thrust, the total rotor thrust
will act through the hub at right angles to the plane of rotation.

15. Control. A helicopter is able to climb and descend vertically, move horizontally in any
direction and while hovering over a spot on the ground, turn onto any selected heading.

16. Vertical Movement. To achieve vertical movement, the total rotor thrust must be increased
by increasing the pitch angle of each and every blade, which, in turn increases the angle of attack.
The pitch angle is increased collectively by use of the collective pitch lever. The reverse takes place in
a vertical descent.

17. Control of Rrpm. An increase in total rotor thrust necessitates an increase in total reaction,
which, in turn, will give an increase in rotor drag. Engine power must therefore be increased to
maintain Rrpm when increasing total rotor thrust, and vice versa. The designer normally incorporates
some form of cam-operated linkage between the collective lever and the engine in order to maintain
Rrpm. A twist-grip type throttle is also provided on the lever for fine manual rpm adjustments.

18. Horizontal movement. The thrust required to move the helicopter horizontally must be
obtained from the total rotor thrust. This can be achieved by tilting the disc so that the rotor disc is
tilted in the direction of the required movement. To enable the disc to tilt, the pitch angle on one side
of the disc must, at the same time, be decreased by the same amount, causing the blades to
descend. To keep the rotor disc in the tilted position, the pitch must vary throughout the blades 3600
cycle of travel. This change in pitch is therefore known as a cyclic pitch change and is achieved by the
pilot moving a cyclic pitch stick
449 Basics of Rotordynamics

19. Torque Reaction. Unless blanked in some way, the fuselage will yaw in the opposite
direction to the main rotor as a result of torque reaction. There are a number of ways by which this
reaction can be overcome, but the only method considered here would be the fitting of a tail rotor.
When the moment of the tail rotor thrust equals the torque reaction couple, then the fuselage will
maintain a constant direction. As the torque reaction is not constant, some means must be provided to
vary the thrust from the tail rotor. This is achieved by the pilot moving yaw pedals which collectively
change the pitch, and thereby the angle of on the tail rotor blades.

20. Additional Tail Rotor Functions. The tail rotor has the following additional functions:

(a) To alter the direction of the fuselage while hovering.

(b) To maintain a balanced condition of fight.

(c) To stop the fuselage rotating in power-off (autorotative) flight. When the rotors are
being turned purely by the reaction of the air and without any assistance from the engine,
there will be no torque reaction; friction will cause the fuselage to rotate in the same direction
as the rotor. Directional control is maintained by changing the pitch on the tail rotor to such a
degree that the tail rotor produces a thrust in a direction opposite to the required when the
rotor is being driven by engine power. The tail rotor blades are symmetrical in shape and
must be capable of being turned to produce plus or minus values of pitch.

21. Flapping. Flapping is the angular movement of the blade above and below the plane of the
hub. Flapping relieves bending stresses at the root of the blade which might otherwise be caused by
cyclic and collective pitch changes or
changes in the speed and direction of
the airflow relative to the disc. In a rigid
rotor system bending stresses are
absorbed by designed deformation of
the rotor / hub combination. In an
articulated rotor, bending stresses are Fig 23-8: Flapping Hinge
avoided by allowing the blade to flap
about the flapping hinge (Fig 23-8)

22. Coning. Rotor thrust will


cause the blades to rise about the
flapping hinges until they reach a
position where their upward movement
is balanced by the outward force of
centrifugal reaction being produced by Fig 23-9: Centrifugal Reaction

the rotation of the blades, (Fig 23-9). In normal operation the blades are said to be coned upwards,
FIS Book 1: Helicopter Flight 450

the coning angle being measured between the span-wise length of the blade and the blades tip path
plane. The coning angle will vary with combinations of rotor thrust and rotor rpm Nr, (Fig 23-9). If
rotor thrust is increased and Nr remains constant, the blades cone up. If Nr is reduced, centrifugal
force decreases and if rotor thrust remains constant, the blades again cone up. The weight of the
blade will also have some effect but for any given helicopter this will be constant.

Limits of Rotor RPM

26. Because the area of the rotor disc reduces as the coning angle increases, thereby decreasing
thrust the coning angle must never be allowed to become too big. As centrifugal force gives a
measure of control of the coning angle through rotor rpm, provided the rotor rpm is kept above a laid
down minimum, the coning angle will always be within safe operating limits. There will also be an
upper limit to Nr due to transmission considerations and blade root loading stresses. Compressibility,
due to high blade tip speeds, is also a limiting factor. rotor rpm limits are to be found in the appropriate
Aircrew Manual.

Overtorqueing

27. Overtorqueing is possible on turbine driven helicopters when the transmission system can not
absorb the high torque that turbine engines are capable of producing. Overtorqueing can be avoided
by careful monitoring of the torque gauge and careful use of the helicopter controls.

Overpitching

28. Overpitching is a dangerous condition reached following the application of pitch to the rotor
blades without sufficient engine power to compensate for the extra rotor drag. The rotor rpm decays,
coning angle increases, disc area and rotor thrust reduce. A need is felt to increase the collective
further which again decays rotor rpm. This can be prevented only by reducing the pitch angle of the
blade by reducing collective. This entails a loss of height which can be dangerous close to ground.
451

CHAPTER 24

HOVERING AND HORIZONTAL MOVEMENT

HOVERING

Lift -Off and Climb to a Free Air Hover

1. To lift a helicopter off the ground, a force must be produced greater than the weight, which
acts vertically downwards through the aircraft's centre of gravity (CG). On the ground with minimum
pitch set, the total rotor thrust is small, and on some aircraft can even be negative, and the aircraft
remains on the ground. As the collective
lever is raised blade pitch and the angle of
attack are increased and the total rotor
thrust becomes equal to AUW and the
helicopter is resting only lightly on the
ground. A further increase in angle of
attack causes total rotor thrust to exceed
the AUW (Fig 24-1 a) and the helicopter
accelerates vertically (in still air
conditions).
Fig 24-1a: Climb to hover

2. As the Rate of Climb (ROC)


increases there is a relative airflow down
through the rotor. This adds to and
increases the induced airflow (IF). The
angle of attack and total rotor thrust (TRT)
are automatically reduced by the
increased IF and the acceleration
Fig 24-1b: Steady Hover
decreases until a steady ROC is achieved
with TRT = AUW, (Fig 24-1b).

3. In the climb the Total Reaction


Vector is tilted away from the axis of
rotation because the direction of the RAF
has changed. Rotor drag is increased and
more power is required to maintain rotor
rpm (Fig 24-2).
Fig 24-2: TRT Equal to AUW
FIS Book 1: Helicopter Flight 452

4. To stop the climb, collective pitch and angle of attack are reduced and the TRT is now less
than AUW. The helicopter's ROC decreases, IF reduces, angle of attack re-increases and TRT
increases until a steady hover is achieved with TRT equal to AUW. The helicopter is now said to be
in a Free Air Hover.

Vertical Descent

5. At low rates of descent the


sequence is the reverse of the vertical
climb, that is, due to downward
movement, IF will be opposed and angle
of attack will increase, see Fig 24-3. At
higher rates of descent airflow is more
complex.

6. When climbing or descending Fig 24-3: Vertical Descent

there will be some parasite drag from


the fuselage but the amount is small
since a ROC or ROD of 1200 ft / min is
barely 12 Kt.

Ground Effect

7. In a free air hover the airflow


through the rotor disc begins at zero
velocity some distance above and
accelerates through the disc and into the
air below. There is little resistance to Fig 24-4: Divergent Duct
the downward movement of air. If the
helicopter is hovered close to the ground
the downwash meets the ground, is
opposed, and escapes horizontally. A
divergent duct is produced causing an
increase in pressure (Fig 24-4). The
increased pressure of the air beneath
the helicopter opposes and reduces the
induced flow so that angle of attack and
Fig 24-5: AOA / TR Thrust Increase
hence TRT are increased for a given
pitch setting ( Fig 24-5). In order to remain at a constant height the collective pitch must be reduced,
to reduce the angle of attack and keep the TRT equal to AUW. The TR will have moved closer to the
453 Hovering and Horizontal Movement

axis of rotation producing a reduction in rotor drag and in power required to hover in Ground Effect.
Helicopters are said to hover Inside Ground Effect (IGE) or, when in free air hover, Outside Ground
Effect (OGE).

8. Factors affecting Ground Effect. Ground effect is affected by the following factors:

(a) Height. The reduction in IF is greater when the rotor is close to the ground. Ground
effect reduces with increase in height until it is negligible above 2 / 3 rotor diameter distance
from the ground.

(b) Slope. On sloping ground much of the air flows downhill and there is reduced
ground effect because there is no development of a divergent duct.

(c) Nature of the Ground. Rough ground will tend to disrupt the air flow preventing a
divergent duct from being formed.

(d) Wind. The downwash is displaced downwind reducing ground effect. However, as
wind speed increases IF is reduced by translational lift, which is described in chapter 25.

Recirculation

9. Whenever a helicopter is hovering


near the ground some of the air passing
through the disc is re-circulated and it
would appear that the re-circulated air
increases speed as it passes through
the disc a second time, see Fig 24-6.
This local increase in IF near the tips
gives rise to a loss of rotor thrust. Some
recirculation is always taking place, but Fig 24-6: Recirculation

over a flat, even surface the loss of rotor


thrust due to recirculation is more than
compensated for by ground effect. If a
helicopter is hovering over tall grass or
similar types of surface the loss of lift due
to recirculation will increase and, in some
cases the effect will be greater than
ground effect and more power would be
required to hover near the ground than in
Fig 24-7: Increased IF Near Tips
free air (Fig 24-7), heavy helicopters can
FIS Book 1: Helicopter Flight 454

experience this phenomenon hovering over


water.

10. Recirculation will increase when any


obstruction on the surface or near where the
helicopter is hovering prevents the air from
flowing evenly away. Hovering close to a
building, wire link fencing or cliff face may cause
Fig 24-8: Recirculation near obstruction
severe recirculation (Fig 24-8).

HORIZONTAL MOVEMENT

Cyclic Pitch Changes

11. For a helicopter to move


horizontally the rotor disc must be
tilted so that the total rotor thrust
vector has a component in the
direction required (Fig 24-9). To
enable the rotor disc to tilt, the
swash plates are tilted so that the
pitch angle on one side of the disc
increases causing the blade to
rise, while the pitch angle on the
other side of the disc must, at the
same time, be decreased by the Fig 24-9: Horizontal Flight
same amount, causing the blade to descend. The pilot moving the cyclic stick controls the tilting of
the swash plates.

Flapping to Equality

12. A cyclic pitch change does not markedly


alter the magnitude of total rotor thrust but simply
changes the disc attitude. The blades flapping to
equality of rotor thrust achieve this. If a blade in a
hover has an angle of attack (24-10), a cyclic stick
movement will decrease the blade pitch and,
assuming that initially the direction of the RAF
remains unchanged, the reduction in pitch will
Fig 24-10: Flapping to Equality
reduce both the blade's angle of attack (a) and
455 Hovering and Horizontal Movement

rotor thrust (Fig 24-11a). The blade cannot maintain


horizontal flight and will now begin to flap down,
causing an automatic increase in the blade's angle of
attack. When the angle returns to the blade thrust will
return to its original value and the blade will continue
to follow the new path required to keep the angle of
attack constant, see Fig 24-11b. Thus cyclic pitch will
alter the plane in which the blade is rotating, but the
angle of attack remains unchanged. The reverse
takes place when a blade experiences an increase in
cyclic pitch. It should be remembered that when a
cyclic pitch change is made, the blades continuously
flap to equality as they travel through 360° of
movement.
Fig 24-11 : Flapping to Equality
Control Orbit

13. In its simplest form of operation, movement of the cyclic stick causes a flat plate, or non
rotational swash plate, mounted centrally on the rotor shaft to tilt, the direction being controlled by the
direction in which the cyclic stick is moved. Rods of equal length, known as pitch operating arms
(POA) or pitch change rods connect the swash plate to the rotor blades. When the swash plate is
tilted the pitch operating arms move
up or down, increasing or
decreasing the pitch on the
blades (Fig 24-12). The amount by
which the pitch changes, and which
blades are affected, depends on the
amount and direction in which the
swash plate is tilted. The swash
Fig 24-12: Control Orbit
plate can be more accurately
described as a control orbit because
it represents the plane in which the
pitch operating arms are rotating.

14. Now consider the effect of


the movement of the POA when the
control orbit has been tilted
(assuming that the control orbit tilts
in the same direction as the stick is Fig 24-13 a: Pitch Operating Arm Movement
being moved) as in Fig 24-13. A
FIS Book 1: Helicopter Flight 456

Plan view of the same shows


clearly the amount by which the
control orbit has been tilted at
four positions, A, B, C and D. If
the movement of the POA
through 360° of travel is plotted
on a simple graph, the result
Fig 24-13: Movement of Pitch Operating Arms Through 3600
would be as shown in Fig 24-13.
The rate at which the POA is moving up and down is
not uniform. This can be shown more clearly as a
comparison is made between the control orbit in plan
view and the control orbit in side elevation; and noting
how much movement takes place in each 30° of travel
over a range of 90°, see Fig 24-14.

15. Resultant Change in Disc Attitude. In


order to determine the resultant change in disc
attitude, the movement of each blade is followed
through four points A,B,C and D during 360° of
movement. The control orbit has been tilted by the
Fig 24-14: Rate of Movement of Pitch
cyclic stick and hence the pitch operating arms move Operating Arms
so that a maximum pitch of +2° is applied at point B; a
minimum pitch, -2°, at point D, and zero pitch at points A
and C, see Fig 24-16. As the blade moves clockwise
from A it will experience an increase in pitch and the
blade will begin to flap up. The rate of flapping will vary
with the amount of pitch change so the blade will be
experiencing its greatest rate of flapping as it passes B,
the point of maximum pitch change. In its next 90° of
travel the pitch is returned from +2° to 0 at point C and
the rate at which the blade is flapping will slowly reduce
to reach zero at point C. Flapping up, however, will have
continued past B and the blade will be at its highest point
at C. The exact reverse will take place after C, resulting
in the blade being at its lowest at point A. The disc will
now be tilted along the axis B-D. This is 90° out of
phase with the maximum and minimum pitch positions
(Fig 24-15).
Fig 24-15: High and Low Blade
Position
457 Hovering and Horizontal Movement

Phase Lag

16. When cyclic pitch is applied the blades will automatically flap to equality and, in so doing, the
disc attitude will change, the blade reaching its highest and lowest positions 90° later than the point
where it experiences the maximum increase
and decrease of cyclic pitch. The variation
between the tilt of the control orbit in
producing this cyclic pitch change and
subsequent tilt of the rotor is known as
phase lag. Phase lag will also occur when
the blades experience a cyclic variation
resulting from a change in speed or direction
of the RAF, as occurs in horizontal flight.

Advance Angle

17. Phase lag, if uncorrected, would


have the effect that movement of the cyclic
stick would cause the rotor to tilt in a
direction 90° out of phase with the direction Fig 24-17: 900 Advance Angle
in which the cyclic stick is moved. Thus
moving the cyclic stick forward would have
the effect of moving the helicopter sideways.
This undesirable feature is overcome by
arranging for the blade to receive the
maximum alteration in cyclic pitch change
90° before the blade is over the highest and
lowest points on the control orbit, see Fig
24-17. The angular distance that the POA is
positioned on the control orbit in advance of
the blade to which it relates is known as the
advance angle. When the control orbit tilts to
follow the stick, to compensate fully for
phase lag, the advance angle would have to
be 90°. If the control orbit is 45° out of
phase with stick movement, then the
advance angle needs to be only 45° to
make full compensation for phase lag,
Fig 24-18: 450 Advance Angle
see Fig 24-18.
FIS Book 1: Helicopter Flight 458

Dragging

18. Dragging is the freedom


given to each blade to allow it to
move in the plane of rotation
independently of the other
blades. To avoid bending
stresses at the root, the blade is
allowed to lead or lag about a
dragging hinge, see Fig 24-18,
but rate of movement is
Fig 24-18: Dragging Hinge
restricted by some form of drag
damper to avoid undesirable oscillations. Dragging is caused by:

(a) Periodic Drag Changes When the helicopter moves horizontally, the blade's angle
of attack is continually changing during each complete revolution to provide symmetry of rotor
thrust. The variation in angle of attack results in variation in rotor drag and consequently the
blade will lead or lag about the dragging hinge.

(b) Conservation of Angular Momentum If a helicopter is stationary on the ground


in still air conditions, rotor running, the radius of the blade's CG relative to the axis of
rotation/shaft axis will be constant. If the cyclic stick is now moved the blades will flap to
produce a change in disc attitude. The axis of rotation will no longer be coincident with the
shaft axis and this results in a continuous change of the CG radius relative to the shaft
axis through 360° of
travel. The radius
variation will cause
the blades to speed
up or slow down
depending on
whether the radius is
reducing or
increasing, Fig 24-19.
Fig 24-19: Variation in Radius of Blade CG Resulting from
Flapping
(c) Hooke's
Joint Effect. Hooke's joint effect is the movement of a blade to reposition itself relative to
the other blades when cyclic stick is applied; its effect is very similar to the movement of the
blades CG relative to the hub. If a rotor is hovering in still air, see Fig 24-20a and b, when
viewed from above the shaft axis, the blades A,B,C and D appear equally spaced relative to
the shaft axis. When a cyclic tilt of the disc occurs, Fig 24-20c and d, the cone axis will have
459 Hovering and Horizontal Movement

tilted but, if still viewed from the shaft axis, which has not tilted, blade A will appear to have
increased its radius and blade C decreased its radius. Blades B and D must maintain position
as in Fig 24-20c in order to achieve their true positions on the cone. It follows therefore that
they must move in the plane of rotation to position themselves as in Fig 24-20 d.

Fig 24-20: Hooke’s Joint Effect


FIS Book 1: Helicopter Flight 460
461

CHAPTER 25

CONTROL IN FORWARD FLIGHT

Introduction

1. When torque is applied to the rotor shaft of a helicopter there is an equal and opposite torque
reaction applied to the helicopter by the rotor shaft. If the torque reaction is not balanced the
helicopter fuselage will turn in the opposite direction to the rotor. In this chapter torque reaction and
the solution to it will be discussed. The forces in the hover and in forward flight, and transition from
forward flight to the hover will also be discussed.

Torque Reaction

2. The torque reaction on a


single rotor helicopter is shown in
Fig 25-1. A torque compensating
force at the tail is the most
common method of balancing
torque reaction and the force is
provided by a tail rotor or
shrouded tail rotor (Fenestron).

3. The Tail Rotor. The tail


rotor is mounted vertically at the
rear of the fuselage and clear of
the main rotor, see Fig 25-2. It is Fig 25-1: Torque Reaction
driven from the main gearbox by
a tail rotor drive shaft and geared
such that the shaft revolves at a
very high speed compared to the
main rotor. The reason for an SPIDER ASSEMBLY
increase in the rpm of the tail rotor
drive shaft is to allow the
CONTROL ROD
construction of it to be flimsier
because the torque, which is
directly proportional to rpm, is
reduced. It is also easier to Fig 25-2: Conventional Tail Rotor
balance the shaft if it rotates at
high rpm.
FIS Book 1: Helicopter Flight 462

4. The Shrouded Tail Rotor. The shrouded


Cross-section of
tail rotor, or Fenestron, is a high speed, variable
Cambered Fin
pitch ducted fan mounted in a cambered fin (Fig 25-
3). It has many features in common with a propeller
but it has control characteristics similar to a tail rotor.

5. Control Mechanism. When the moment


of the tail rotor thrust equals the torque reaction
couple, then the fuselage will maintain a constant
direction. As the torque reaction is not constant
some means must be provided to vary the thrust
from the tail rotor. This is achieved by the pilot
moving yaw pedals which collectively change the Fig 25-3: Shrouded Tail Rotor
pitch and, therefore, the thrust from the tail rotor. (Fenestron)

6. Additional Tail Rotor Functions. The tail rotor has the following additional functions:

(a) Heading control in the hover is achieved by increasing or decreasing tail rotor thrust
so that torque reaction is not balanced and the helicopter is able to turn about the rotor shaft.

(b) Balance in forward flight is adjusted by tail rotor thrust in a similar fashion to the
rudder control of an aeroplane.

(c) In power off flight (autorotation) there is no torque reaction. The rotor is turning and
there is friction in the transmission which tends to turn the helicopter in the same direction as
the rotor. The turn is prevented by negative pitch on the tail rotor which produces thrust
opposite to that in powered flight.

Tail Rotor Compensation

7. Tail Rotor Drift. If a


fuselage is being turned by a
couple YY1, about a point, the
rotation will stop if a couple ZZ1,
of equal value, acts in the
opposite direction, Fig 25-4a.
The rotation would also stop if a
single force ZZ1 was used to
Fig 25-4: Tail Rotor Compensation
produce a moment equal to the
463 Control in Forward Flight

couple YY1, Fig 25-4b, but there would now be a side force X on the pivot point, Fig 25-4c. This side
force is known as tail rotor drift and, unless corrected, it would result in the helicopter moving
sideways over the ground.

8. Correcting For Tail Rotor Drift. Tail rotor drift can be corrected by tilting the rotor disc
away from the direction of the drift. This can be achieved by:

(a) The pilot making a movement of the cyclic stick.

(b) Rigging the controls so that when the stick is in the centre the disc is actually tilted by
the correct amount.

(c) By mounting the gearbox so that the drive shaft to the rotor is offset.

9. Tail Rotor Roll. If the tail rotor is


mounted on the fuselage below the level of the
main rotor the tail rotor drift corrective force
being produced by the main rotor will create a
rolling couple with the tail rotor thrust, causing
the helicopter to hover one wheel or skid low.
The amount of roll depends upon the value and
angle of the tail rotor thrust and the vertical
separation between main and tail rotors. In the
hover, the helicopter will roll about the horizontal
couple until the movement is balanced by the
couple of the vertical component of total rotor
thrust and the helicopter all up weight (25-5). Fig 25-5: Tail Rotor Roll

Rotor Configurations

10. It is possible to counteract torque reaction by using twin main rotors which may be mounted
co-axially and revolve in opposite directions (Kamov), or in a fore and aft configuration (Chinook) or
even side-by-side. In all cases synchronization of the rotors is vital for the maintenance of directional
control.

Forces in the Hover and in Forward Flight

11. Forces in Balance – Hover. In a free air hover the total rotor thrust will be acting vertically
upwards through the axis about which the blades are rotating and at right angles to the tip path plane.
Weight will be acting vertically downwards through the CG, Fig 25-6a. If the helicopter is loaded to
position the CG immediately below the blades’ axis of rotation, and discounting downdraft on any
FIS Book 1: Helicopter Flight 464

horizontal surfaces, no change in


fuselage attitude will occur when the
helicopter leaves the ground, Fig 25-6b.
If, however, the CG is not below the axis
of rotation, Fig 25-6c, a couple will exist
between total rotor thrust and the
weight, and the fuselage will pitch until
both forces are in line, Fig 25-6d. It
should be noted that a helicopter in the
hover often adopts a nose-up attitude in
any case, irrespective of the position of Fig 25-6: Forces in Balance - Hover
the CG. This happens because downwash from the main rotor exerts a force on the tail stabilizer
causing a tail-down moment. In still air conditions the nose-up attitude is quite marked but as wind
speed increases the vertical component of rotor downwash is reduced and the helicopter adopts a
more level attitude. Hovering attitude is also affected by flapback, which is discussed in para 28.

12. Forces in Balance - Forward Flight. If a helicopter moves from a free air hover into
forward flight with no change in the fuselage attitude, the rotor disc will be tilted forward and the
disposition of forces will be as shown in Fig 25-7a. Total rotor thrust is now inclined forward and
produces a nose-down pitching moment about the CG. The vertical components of TRT and AUW
remain in line but a couple now exists between the horizontal component of TRT and fuselage
parasite drag as the aircraft speed increases. The fuselage will pitch forward but the moment will now
be opposed by the vertical component of TRT and Wt with the forces resolved as in Fig 25-7b. The
fuselage will only pitch forward until the couples are in balance. This will occur when TRT is in line
with the CG. CG therefore controls the position of the fuselage in relation to the disc. This
relationship is affected in forward flight by the negative lift effect of the tail stabilizer and the moment
exerted by it.

Fig 25-7a: Level Attitude with Pitching Fig 25-7 b: Forces in Balance Transition
Moment
465 Control in Forward Flight

13. To achieve forward flight the rotor disc is tilted so that TRT produces not only a vertical force
to balance the weight but also a horizontal force in the direction of flight. The change of state from the
hover to forward flight, or from forward flight to the hover, is known as transition.

14. The Sequence of Events during Transition. As the helicopter moves initially from the
hover the disc, and hence TRT, is tilted. The vertical component of TRT is reduced and becomes less
than Wt and to prevent the helicopter from descending TRT is increased with more collective pitch.
The power required increases, see Fig 25-8. As the aircraft accelerates the fuselage acts
pendulously below the main rotor and pitches nose down, Fig 25-9.

Fig 25-8: Forces during Transition Fig 25-9: Fuselage Pendulosity

Translational Lift

15. When a helicopter is in a free air hover in still air conditions, for a given rotor rpm (Rrpm) a
certain value of collective pitch, say 8°, will be required to support it in the air. A column of air, the
induced flow, will be continually moving down towards the rotor disc, and this downward flow of air
must be considered when determining the direction of the airflow in relation to the blades. It will be
noted that the angle of
attack, say 4°, is less than
the pitch angle. The
angle of attack depends
on the value of the
induced flow; if the
induced flow is removed,
the angle of attack
Fig 25-10: Induced Flow from Vertically Above
becomes the same as the
pitch angle (Fig 25-10).
FIS Book 1: Helicopter Flight 466

16. If the effect of a helicopter facing into a 20 Kt wind is considered, and it is assumed that it is
possible for it to maintain the hover without tilting the disc, the horizontal flow of air (wind) will blow
across the vertically induced column of air and deflect it down-wind before it reaches the disc. The
column of air which was flowing down towards the disc will, therefore, be modified and gradually be
replaced by a mass of air which is moving horizontally across the disc. The rotor will act on this air
mass to produce an induced flow but the velocity of the induced flow will be greatly reduced, see Fig
25-11 a and b.
Therefore, an airflow
parallel to the disc will
reduce the value of
the induced flow,
increase the angle of
attack and, therefore,
Fig 25-11: Induced Flow with Air Moving Horizontally
rotor thrust.

17. To maintain the hover condition when facing into wind, the disc must be tilted forward. The
horizontal flow of air will not now be parallel to the disc, and a component of it can now be considered
to be actually passing through
the disc at right angles to the
plane of rotation, effectively
increasing the induced flow,
see Fig 25-12. To consider
the extreme case if the rotor
disc were tilted 90 to this
horizontal flow of air, then all
of it would be passing through Fig 25-12: Induced Flow with Disc Tilted Forward
the disc at right-angles to the
plane of rotation.

18. As described in para 16, the effect of the horizontal airflow across the disc when hovering into
wind is to reduce the induced flow but, because the disc has had to be tilted forward, (para 17) a
component of this horizontal airflow will now be passing through the disc, effectively increasing the
induced flow; both of these effects must now be taken into consideration to give the total flow towards
the disc and to determine the direction of the airflow relative to the blades. Provided the reduction in
the induced flow caused by the flow parallel to the disc is greater than the increase caused by the
component of horizontal airflow passing through the disc, then the relative airflow will be nearer the
plane of rotation than when the helicopter is in the hover, the angle of attack will increase and the
aircraft will climb. Therefore, the collective pitch can be decreased to say, 7 , while maintaining an
angle of attack of 5 . As the relative airflow moves nearer the plane of rotation, the total reaction must
467 Control in Forward Flight

move forward. There will, therefore, be less rotor drag, and rotor rpm can be maintained with less
power.

19. The reduction in induced flow, translational lift, first takes effect when air moves towards the
disc at approximately 12Kt. The reduction is appreciable at first, and although it continues to reduce
as the velocity of the horizontal airflow increases, the rate at which it reduces becomes progressively
less because there is less induced flow to be influenced.

20. The rotor disc has to be tilted forward to provide


a thrust component equal to parasite drag. Parasite
drag is low at low forward speed so only a small tilt of
the disc is required to provide a balancing amount of
thrust and, with only a small tilt of the disc, only a small
component of the horizontal airflow will be passing
through the disc at right angles to the plane of rotation.
Because the parasite drag increases as the square of
the speed, the greater must be the amount that the disc
must be tilted to provide the necessary increase in thrust
and, as the horizontal airflow approaching the disc
increases, the greater will be the component of it
passing through the disc at right angles to the plane of
rotation, see Fig 25-13b. If the curves in Figs 13a and b
are now transferred to one graph it will be seen that the
total flow of air at right angles to the plane of rotation at
first decreases and then increases again, becoming a
minimum when the two airflows have the same value,
see Fig 25-14. As the flow of air through the disc Fig 25-13: Airflow thought Disc
decreases, less collective pitch and power will be
required to maintain the required angle of attack. When
the flow of air through the disc begins to increase again,
then collective pitch and power must be increased if the
required angle of attack is to be maintained.

Summary of Transition

21. The sequence of events as a helicopter moves Fig 25-14: Variation of Total Airflow
through the Disc with
into forward flight is summarized as follows: Forward Flight

(a) The cyclic stick is moved forward to tilt the disc and the TRT forward.
FIS Book 1: Helicopter Flight 468

(b) The vertical component of TRT is reduced and the collective pitch must be increased
to maintain height. More power is required.

(c) As airspeed increases the disc flaps back. The disc attitude is maintained by
increasing forward cyclic control.

(d) As airspeed increases inflow roll tilts the disc to the advancing side. The disc attitude
is maintained by cyclic control to the retreating side.

(e) As airspeed increases the TRT increases with increased translational lift and the pilot
lowers the collective to maintain height. Less power is required.

(f) During power changes the changing torque reaction is balanced by movement of the
yaw pedals.

Transition From Forward Flight to Hover

22. In order to decelerate a helicopter from steady level flight to the hover the balance of forces
must be changed. The general method of coming to the hover from forward flight is by the pilot
executing a flare by tilting the disc in the opposite direction to that in which the helicopter is moving.
The handling techniques needed to control the manoeuvre differ from those required for a more gentle
transition.

23. The Flare. To execute a flare the cyclic stick is moved in the opposite direction to that in
which the helicopter is moving. The harshness of the flare depends upon how far and how fast the
stick is moved. The flare will produce a number of effects.

24. Flare Effects. The following effects occur


during the flare:

(a) Thrust Reversal. By tilting the


disc away from the direction of flight the
horizontal component of total rotor thrust will
now act in the same direction as parasite
drag causing the helicopter to slow down
very rapidly, see Fig 15a. The fuselage will
respond to this rapid deceleration by
pitching up because reverse thrust is being
maintained whilst parasite drag decreases.
If no corrective action is taken the disc will Fig 25-15: Change in Relative Airflow
be tilted further still, adding to the deceleration effect, Fig 25-15b.
469 Control in Forward Flight

(b) Increase in Total Rotor Thrust. Another effect of tilting the disc back whilst the
helicopter is moving forward is to change the airflow relative to the disc, Fig 25-16. As was
explained in paragraph 19, translational lift, a component of the horizontal airflow (due to the
forward movement of the ( helicopter) is passing through the disc at right angles to the plane
of rotation, opposing the induced flow. The result is an increase in total rotor thrust. To
prevent a climb the collective lever must be lowered.

Fig 25-16: Increase in Total Rotor Thrust

(c) Increase in Rotor RPM. Unless power is reduced when collective pitch is reduced,
the Rrpm will rise. It will also increase rapidly in the flare for two other reasons, conservation
of angular momentum and reduction in rotor drag.

(i) Conservation of Angular Momentum. An increase in total rotor thrust


causes the blades to cone up. The radius of the blades’ CG from the shaft axis
decreases and the rotational velocity will automatically rise.

(ii) Reduction in Rotor Drag. Rotor drag is reduced in the flare because the
total reaction moves towards the axis of rotation. This results from the changed
direction of the relative airflow. The forward movement of the total reaction vector
causes the rotor drag component to be reduced, Fig 25-16b.As a result of the flare,
the speed reduces rapidly and the flare effects disappear. Collective pitch and power
which had been reduced during the flare must be replaced and, in addition, more
collective pitch and power must be used to replace the loss of translational lift caused
by the speed reduction, otherwise the aircraft would sink. The cyclic stick must also
be moved forward to level the aircraft and to prevent the helicopter moving
backwards. The power changes necessary during the flare have an effect on the
aircraft in the yawing plane. Therefore, yaw pedals must be used to maintain heading
throughout.
FIS Book 1: Helicopter Flight 470

Landing

25. If collective pitch is reduced slightly in a hover IGE, the helicopter will descend but settle at a
height where ground effect has increased total rotor thrust to again equal all up weight. Therefore a
progressive lowering of the collective lever is required to achieve a steady descent to touchdown.
When the helicopter is close to the ground the tip vortices are larger and unstable causing variation in
the thrust around the rotor disc and turbulence around the tail and makes control difficult. For this
reason, and to help to prevent ground resonance, the helicopter is normally landed firmly to decrease
the chance of drifting when touching down.

Symmetry and Dissymmetry of Rotor Thrust

26. Symmetry of Rotor Thrust. If a helicopter


is stationary on the ground in still air conditions, rotor
turning and some collective pitch applied, then the
rotor thrust produced by each blade will be uniform.
The speed of the relative airflow over each blade will
be equal to the speed of rotation of the blade, and if
a given section on each blade of a four-bladed rotor
is considered, the vector showing the relative airflow
will have the same value irrespective of the position
of the blade during its 360° of travel, see Fig 25-17.
As the velocity of this airflow is equal to the blade’s
Fig 25-17: Relative Airflow
speed of rotation, this airflow will be referred to as - Still Air
VR.

27. Dissymmetry of Rotor Thrust. If the


conditions change and the helicopter now faces into
a wind, during the blade’s rotation through 360° half
the time it will be moving into wind and the
remainder of the time it will be moving with the wind.
The disc can therefore be divided in half, one half
being the advancing side and the other the
retreating side, see Fig 25-18. When the blade is at
right angles facing into wind (position B), the velocity
of the relative airflow will be a maximum and if the
value of the wind speed is referred to as VW, then at Fig 25-18: Relative Airflow - Wind
Conditions
position B the velocity of the relative airflow will be
VR + VW. As the blade continues to rotate, the effect of VW will decrease and when the blade
reaches position D the velocity of the relative airflow will have become VR – VW. If no change has
471 Control in Forward Flight

taken place in the blade’s plane of rotation, the rotor thrust being produced by the advancing blade at
position B will be greatest and, for the retreating blade at position D, least. The value of rotor thrust
across the disc will no longer be uniform and unless some method is employed to provide equality,
the helicopter will roll towards the retreating side. This condition, where one side of the disc produces
more rotor thrust than the other, is known as dissymmetry of rotor thrust.

Flapback

28. To maintain control of the helicopter dissymmetry must be prevented; one method of doing
this is to decrease the angle of attack of the advancing blade and increase the angle of attack of the
retreating blade so that each blade again produces the same value of rotor thrust. With the fully
articulated rotor head this change in angle of attack takes place automatically by flapping but, as a
result, the disc attitude changes. The manner in which it changes and the reason why this change in
attitude prevents dissymmetry can be seen by following the movement of a blade through 360° of
travel.

29. Starting at position A of Fig 25-18, the blade starts to travel on the advancing side and the
relative airflow will increase. Rotor thrust begins to increase and, because it is free to do so, the blade
will begin to flap up about the flapping hinge. As the blade flaps up the angle of attack will begin to
decrease, rotor thrust decreases and the blade will proceed to follow a path to maintain the same
value of rotor thrust as it was producing before it began to flap up. The blade, in fact, is flapping to
equality. The further round that the blade progresses on the advancing side, the greater will be the
velocity of the relative airflow; therefore, to maintain a constant value of rotor thrust, the rate at which
the blade is flapping will
steadily increase, with the
maximum rate of flapping
and, therefore, minimum
angle of attack occurring
when the blade reaches
position B. For the next
90° of travel the velocity
of the relative airflow
begins to decrease, so
the rate of flapping will
decrease. When the
blade reaches position C,
the relative airflow will
Fig 25-19: Flapback
have the same value as at
position A, so the rate of flapping dies out completely but, because the blade has been rising all the
time from position A, the blade will reach its highest position at C. The reverse will take place on the
FIS Book 1: Helicopter Flight 472

retreating side, with the blade having its maximum rate of flapping down and, therefore, its maximum
angle of attack at position D, reaching its lowest position at A. In flapping to equality, the blade will
have flapped away from the wind. This change of disc attitude, which has occurred without any
control movement by the pilot, is known as flapback, see Fig 25-19.

30. Figs 25-20a and b show that when the helicopter is on the ground and the disc is subject to
wind, the disc attitude is altered, although no cyclic stick has been applied. The disc has flapped back
relative to the wind and to the control orbit, and the blades are moving about their flapping hinges.
However, the rotor thrust being produced will be the same value as before the disc flapped back,
but tilted in direction.

31. If the pilot now moves the stick forward to return the disc to its original position (Fig 25-20c) it
will be seen that the disc is now flapped back only in relation to the control orbit and not to the wind,
and that movement is no longer taking place about the flapping hinges. Thus flapback has been
counteracted by cyclic feathering, and, since the cyclic stick only changes the disc attitude, the
value of the rotor thrust force remains unchanged. When the helicopter is airborne and moving in any
horizontal direction, the effect will be the same as has been described for a helicopter on the ground
facing into wind, with flapback being prevented by cyclic feathering. The first movement of the cyclic
stick will tilt the disc to initiate horizontal flight, then a second movement will be necessary to prevent
the disc from flapping back when the aircraft moves and gains speed. It should be noted however
that some movement about the flapping hinges will still take place if the CG of the helicopter is not in
the ideal position.

Fig 25-20: Relationship between Disc, Control Orbit and Stick Resulting from Flapback:
Centrifugal Reaction

Inflow Roll

32 The effect of moving air horizontally across the disc causes a reduction in the induced flow.
However, this reduction is not uniform because air passing across the top of the disc is being
continually pulled down by the action of the rotors. Thus air which is moving horizontally towards the
disc will cause the greater reduction in induced flow at the front of the disc, and the smallest reduction
at the rear of the disc The reduction in induced flow for the disc as a whole will produce an increase in
473 Control in Forward Flight

rotor thrust but because the increase in the angle of attack is not uniform, it will also produce a
change in the attitude of the disc. Assuming the flapback has been corrected, see Fig 25-21, the
effect of a cyclic variation in angle of attack for a blade starting at position B, Fig 25-21b, must be
considered. As the blade moves towards position C, the increased angle of attack will cause the blade
to flap to equality. The rate of flapping up will be a maximum as the blade passes position C because
this is the point where there has been the greater reduction in induced flow. In the next 90° of travel
the rate of flapping will slow down, dying out completely when the blade is at position D. Thus the
blade will be rising all the time it is travelling from position B to reach its highest position at D. The
reverse will take place for the next 180° of travel, with the blade having its maximum rate of flapping
down at A and its lowest position at B. As a result of the inflow the disc will, therefore, tilt about axis
AC towards the advancing side. The combined effect of inflow roll and flapback is, therefore, to tilt the
disc about axis ZZ1, Fig 25-21c. As inflow roll will have its greatest effect at low speed, and flapback
its greatest effect at high speed, the axis about which the disc will tilt will vary with forward speed. In
general, the cyclic stick has to be positioned forward towards the retreating side to correct these
effects in forward flight.

Fig 25-21: Combined Effect of Flap back and Inflow Roll

Factors Affecting Maximum Forward Speed

33. There are several factors which must be taken into account when trying to increase the
maximum speed of a helicopter.

34. Cyclic Control Limits. To achieve forward flight the cyclic stick is moved to tilt the disc
forward, the disc tilting by the same amount that the control orbit has been moved. As the airspeed
increases the rotor disc flaps back relative to the cyclic control position and the attitude of the disc is
maintained by moving the cyclic stick forward. There will be a speed at which the cyclic stick is fully
forward and no further acceleration is possible. The amount of forward cyclic control is reduced if the
helicopter’s centre of gravity is aft.
FIS Book 1: Helicopter Flight 474

35. Power Available. In level flight VMAX is limited by power available. A higher speed may
be possible in descent .

36. Structural Strength. As speed increases, both, the forces on the rotor and transmission
and the levels of vibration, increase. Apart from the limitation of the strength of the airframe and other
components against these forces, the combination of stress and vibration causes fatigue. It is
impractical to make components so strong that they do not suffer fatigue and, therefore, the level of
vibration must be kept below that at which the failure of components may occur. This will set a limit to
the maximum speed.

37. Airflow Reversal. The speed of


rotation of the retreating blade is high at
the tip and low at the root, but the airflow
from forward flight will have an equal value
for the whole length of the blade and,
where the airflow from forward flight is
greater than the blade’s rotational velocity,
eg at the root end, the airflow will be from
the trailing edge to the leading edge,
causing a loss of rotor thrust. At higher
airspeeds the airflow is reversed over a
progressively large section of the blade
leading to a greater loss of thrust, see Fig
Fig 25-22: Airflow Reversal
25-22. The reduction of rotor thrust on the
retreating blade by airflow reversal is
countered by greater cyclic control and
hence the retreating blade operates at an
increasingly higher pitch angle and hence
angle of attack.

38. Retreating Blade Stall. As the


angle of attack of the retreating blade is
steadily increased with increasing forward
airspeed there will be a speed at which
the airflow breaks away and the blade
Fig 25-23: Retreating Blade Stall
stalls. The large sudden loss of rotor
thrust will cause the blade to flap down, but instead of flapping to equality the effect will be simply to
stall the blade even further. The stall starts at the tip first and spreads inboard as shown in Fig 25-23.
The variation in angle of attack along the blade will be offset to some extent by washout but in all
conditions of forward flight the highest angle of attack will be at the tip.
475 Control in Forward Flight

39. Characteristics of Retreating Blade Stall. The approach of the retreating blade stall can
be detected by:

(a) Rotor roughness.


(b) Erratic stick forces.
(c) Stick shake.

40. If these conditions are ignored, a pitch up tendency will develop followed by a roll to the
retreating side. There will be a substantial loss of control and if the stall is severe, control may be lost
completely.

41. Causes of Retreating Blade Stall. Retreating blade stall can occur as a result of:

(a) High forward speed.


(b) High G manoeuvres.
(c) Rough, abrupt or excessive control movements
(d) Flying in turbulent air.

42. A high all up weight/high density altitude will also aggravate the situation. Recovery action
will depend upon which of the above in-flight conditions are prevailing when the stall symptoms are
recognized. Recovery will normally be made by reducing forward speed, reducing collective pitch,
reducing the severity of the manoeuvre or by a combination of these recovery actions.

43. Compressibility. As an example, the speed of rotation of the tip of a Gazelle rotor blade is
approximately 400Kt. In forward flight at 150 Kt the advancing blade tip has a relative velocity of
550Kt. The velocity of sound at sea level is 660Kt. Compressibility is therefore significant. The main
effects of compressibility are:

(a) A reduction in the lift/drag ratio, requiring more power for the same total rotor thrust.

(b) An increase in the pitching moment on the aerofoil, which is normally very small. This
requires greater control forces and leads to vibration.

(c) The production of shock waves which increase vibration and noise. The effects can
be reduced by using a high speed aerofoil section or sweep back at the blade tips. Any such
solutions have penalties at low speeds.
FIS Book 1: Helicopter Flight 476
477

CHAPTER 26

POWER REQUIREMENTS

Introduction

1. The power required to maintain level flight in a helicopter will vary from the hover to maximum
forward speed, and this chapter considers in detail how and why these requirements vary.

Work

2. If a body is to be moved from one position to another, then a force must be applied to
overcome the resistance to movement. When the body is moved, work is said to have been done,
and it is calculated by multiplying the force used by the distance that the body has been moved. The
resistance set up by the rotor blades to be turned, or the resistance caused by moving the fuselage
through the air, is termed drag, and, since in any state of equilibrium, force equals drag, then work
must equal drag x distance.

Power

3. Power is defined as the rate of doing work, or the ratio of the work done to the time taken.
Therefore:

Power = Work = Drag X Distance


Tine Time
But Distance = Velocity, and therefore
Time
Power = Drag X Velocity = Drag X TAS

The equation for calculation drag is:

Drag = CD½ρ V2S


∴ Power = CD½ρ V2S X Velocity
Assuming CD½ρ S is constant, (K)
Power = KV2 X V
= KV3

The resistance, or drag, of a body moving through the air will vary as the square of the speed, but the
power required to balance the drag will vary as the cube of the speed. Power is normally expressed
in terms of kilowatts (1 kw is equal to 737.6 foot pounds force/sec).
FIS Book 1: Helicopter Flight 478

POWER REQUIRED

Introduction

4. The power required by the rotor to maintain level flight throughout the helicopter's speed
range can be considered under three headings:

(a) Rotor Profile Power.


(b) Induced Power.
(c) Parasite Power.

Rotor Profile Power

5. Rotor Profile power is the power required to drive the rotor at minimum pitch at a constant Nr
plus the power required to drive the tail rotor and all ancillary equipment. With minimum pitch applied
there is drag on the blades as they rotate. As the speed of the airflow past the rotor increases, the
profile drag (Zero Lift Drag, see Part 1, Sect 1, Chap 4) of the advancing blade is increased and that
of the retreating blade is reduced. There is, however, an imbalance because the amount by which the
drag is increased on the advancing blade is greater than the amount by which the drag is reduced on
the retreating blade and so, as airspeed increases, the power required to maintain Nr will also
increase. Furthermore, since power increases in proportion to speed cubed the graph representing
rotor profile power might be expected to rise very steeply. This is not the case, however, because in
the early stages of the increase in airspeed, the tail rotor experiences translational lift and therefore
less pitch and less power are required to keep the aircraft straight. The conventional tail rotor also
flaps back and so obtains flare effect leading to a small further saving in power. As forward speed
increases, the rotor profile power curve rises only slowly at first but rises more rapidly in the higher
speed range as the beneficial effects of the
conventional tail rotor are over-ridden by the
increasing drag, Fig 26-1. The fenestron is
different in that the aerofoil section of the
cambered fin provides thrust in the required
direction as airspeed increases, and hence
the tail rotor requires less power as airspeed
Fig 26-1: Rotor Profile Power
increases. This applies up to about 120 Kts.

Induced Power

6. When the collective pitch is minimum there is virtually no rotor thrust being produced. In
order to increase rotor thrust it is necessary to increase blade pitch and this leads to an increase in
rotor drag. To maintain Nr the power must be increased to overcome the rising drag. This increase in
power is known as the induced power because it is the extra power required to overcome the rise in
479 Power Requirements

drag when the blades are inducing air to flow down through the rotor. As explained in Chap 3,
translational lift, induced flow diminishes with forward speed and less collective pitch is needed to
produce the required angle of attack. The curve of induced power will start at a position on the
vertical axis of the graph at Fig 26-2 and will fall
rapidly at first due to the onset of translational lift,
and then fall more slowly as forward speed
increases. The ground effect, shown by the dotted
line, will also reduce power required to hover.
Induced power accounts for approximately 60% of
the power required to hover.

Parasite Power
Fig 26-2: Induced Power

7. As the helicopter speed increases so does


fuselage parasite drag and the rotor disc needs to
be tilted progressively further forward to provide an
increasing horizontal component of total rotor thrust
to balance the parasite drag. The further forward
the rotor disc is tilted the greater the horizontal
airflow through the disc becomes. This component
adds to, and increases, the induced airflow hence
increasing rotor drag. Parasite power is the power
Fig 26-3: Parasite Power
required to overcome this increasing rotor drag.
Parasite power increases as V3, see Fig 26-3.

Power Required

8. The power required to maintain the


helicopter in steady straight and level flight at any
given forward speed will be the combination of rotor
profile power, induced power and parasite power for
that speed, see Fig 26-4.

Power Available
Fig 26-4: Power Required

9. For a helicopter, the power available is considered to be the power which is available to the
rotor and not that which is available from the rotor. For any given altitude this power will remain more
or less constant and it, therefore, appears on the power graph as a straight line, see Fig 26-5.
FIS Book 1: Helicopter Flight 480

10. Performance. The


performance of a helicopter will
lie in the relationship between
the power available and the
power required - the greater the
difference between them the
greater the margin of power.
From Fig 26-5 it can be seen
that a surplus of power available
over the power required exists
over the greater part of the
speed range. The greater the
power margin the more power
can be used for manoeuvring or Fig 26-5: Power Available/Power Required
for climbing.

11. Significant Speeds. Significant speeds are marked on Fig 26-5.

(a) The best rate of climb speed is at point 1, the maximum power margin.

(b) Vmax occurs at point 2 where there is no longer power available to accelerate the
helicopter in level flight.

(c) Minimum power required, and also minimum fuel consumption occur at point 1. This
is the endurance speed and for most helicopters is at about 60-70 Kts.

(d) The range speed occurs at point 3 where a tangent from the origin to the curve
indicates the best ratio of
power required to airspeed.

Effect of Limited Power

12. With changes in air density,


weight and altitude, the power
available and power required curves
will move close together, and power
available may eventually be sufficient
to hover only with the assistance of
ground effect; in extreme conditions,
there may be insufficient power to Fig 26-6: The Effect of Reduced Power Available on
Helicopter Performance
481 Power Requirements

hover at all. Under these conditions there will be a minimum speed below which, even with ground
effect, the helicopter cannot maintain height (Fig 26-6).

Power Checks

13. Conditions at the take-off and landing areas may differ, and in order that the pilot may make
an airborne assessment of the power available before committing himself to a landing, a simple power
check can be carried out. When flying straight and level at a predetermined speed and with landing
Rrpm, the torque required to maintain that speed is noted. The difference between this torque and
the maximum available represents the power margin from which, by reference to prepared data, can
be used to determine the slow speed capabilities of the helicopter. A similar check can be carried out
while hovering and before moving into forward flight, in order to assess the take-off capabilities.

14. In making the check of power available some allowance must be made if the helicopter is
operating above the altitude where the rotor is most efficient. Information on this is available from the
Operating Data Manual for the aircraft.

Best Climbing Angle

15. When operating with limited power, the helicopter must be moving forward in order to climb.
To assess the steepest climbing angle, it is necessary to find the best rate of climb/forward speed
ratio. This can be determined by drawing a line from the point when the power available curve cuts
the vertical axis of the graph tangential to the power required curve (see Fig 6). The point of tangency
indicates the speed for maximum climbing angle, and this will always be less than the speed for
maximum rate of climb.

Turning

16. In addition to providing a component to balance the weight and a thrust force to maintain
speed, the total rotor thrust must supply a further component to change the direction of the helicopter
in a balanced turn, and the greater the angle of bank, the greater this force must be. Its effect is
similar to an increase in weight; with 30° of bank the apparent weight increases by 15%, with 60° of
bank, the apparent weight will increase by 100%. More collective pitch and, therefore, more power
will be required to maintain height in the turn, and the effect on the power required curve is to cause it
to move up the graph. The maximum angle of bank to maintain a level turn is reached when full
power is applied and best climbing speed is maintained. If bank is increased beyond this point, any
attempt to maintain height by use of lever will result in loss of Rrpm, due to overpitching, see Part 2,
Sect 1, Chap 1 and Vol 5, Pt 2, Sect 3, Chap 1.
FIS Book 1: Helicopter Flight 482
483

CHAPTER 27

AUTOROTATION

AUTOROTATION IN STILL AIR

Introduction

1. In powered flight the rotor drag is overcome with engine power but, when the engine fails or is
deliberately disengaged from the rotor system, some other force must be used to maintain the rotor
rpm. This is achieved by allowing the helicopter to descend and by lowering the collective lever fully
so that the resultant airflow strikes the blades in such a manner that the airflow itself provides the
driving force. When the helicopter is descending in this manner, the rate of descent becomes the
power equivalent and the helicopter is said to be in a state of autorotation.

2. Although most autorotation are carried out with forward speed, the explanation as to why the
blades continue to turn when in auto rotation can best be seen if it is considered that the helicopter is
auto rotating vertically downwards in still air. Under these conditions, if the various forces involved
are calculated for one blade, the
calculations will be valid for all
the other blades irrespective of
where the blade is positioned in
its 360° of travel. The various
airflows and angles, which will
be referred to, are shown in Fig
26-1.
Fig 26-1: Autorotation - Terms Used

3. It will be noted that the inflow has been determined from the blades' rotational velocity and the
airflow arising from rate of descent. This is not strictly true as the action of the blades slows down the
rate of descent airflow, producing, in effect, an induced flow, making the inflow angle smaller than has
been shown in Fig 26-1; the fact that it is smaller and how this affects the blade is considered later.

Autorotative Force/Rotor Drag

4. Consider three sections A, B and C of a rotor blade (Fig 26-2). The direction of the relative
airflow for each section can be determined from the rotational velocity and the helicopter's rate of
descent. The rate of descent will have a common value for each section but the rotational velocity will
decrease from the tip towards the root. Comparing sections A, B and C, the inflow angle must
therefore be progressively increasing (Fig 26-3). Because of the washout incorporated in the blade,
FIS Book 1: Helicopter Flight 484

the pitch angle is also increasing and as


the blade's angle of attack is the pitch
angle plus the inflow angle, the blade's
maximum angle of attack will be at the
root.

5. If the angle of attack for each


section of the blade is known, the
lift/drag ratio for these angles of attack Fig 26-2: Distribution of Rotational Forces

can be ascertained by referring to the


aerofoil data tables, and, by adding lift
and drag vectors in the correct ratio, the
position of the total reaction can be
determined (Fig 26-3). Relating total
reaction position to the axis of rotation
(see Fig 26-3b) at section A, the total
reaction lies behind the axis; at section
B it is on the axis and at section C it is in
front of the axis. Having
determined the position of the
total reaction, it can now be
considered in terms of rotor thrust and
rotor drag (Fig 26-3b). At section A, the
condition is the same as in powered
flight and the component of total
reaction in the plane of rotation opposes
rotation and is continually trying to
decelerate the blade. At section B no
part of the total reaction is acting in the Fig 26-3: Effect of Forward Speed on Rate of
plane of rotation and it is all rotor thrust; Descent
at section C the component of total reaction in the plane of rotation assists rotation and is continually
trying to accelerate the blade. Under these conditions it is no longer referred to as rotor drag, but as
the autorotative force.

6. Considering the blade as a whole, the section producing an autorotative force will be
accelerating the blade, whilst the section producing rotor drag will endeavor to slow it down. To
maintain a constant rotor rpm, the autorotative section must be sufficient to balance the rotor drag
section of the blade, plus the drag set up by the ancillary equipment, tail rotor shaft and tail rotor, all of
which continue to function in autorotation.
485 Autorotation

7. In normal conditions with the lever lowered, the blade geometry is such that the autorotative
rpm are in the correct operating range, provided an adequate rate of descent exists. If the lever is
raised during autorotation the pitch angles increases on all sections (Figs 26-2 and 26-3). Section B
will tend towards section A and section C will tend towards B, thus the autorotative section moves
outwards. However, section D at the root becomes stalled and the extra drag generated causes a
decrease in the size of the autorotative section and therefore rpm decreases, stabilizing at a lower
figure. This continues with further raising of the lever until such time as the blade is no longer able to
autorotate.

8. Autorotative descent from high altitudes or at a high all-up weight leads to high rates of
descent. Inflow angles will be higher and autorotative sections will be further outboard on the blades;
rpm will be higher in autorotation under these conditions. It should be noted, however, that descent
into more dense air decreases rate of descent and rpm for a constant lever position.

Rate of Descent

9. If the engine fails during a hover in still air and the collective pitch is reduced, the helicopter
will accelerate downwards until such time as the angle of attack is producing a total reaction to give
an autorotative force to maintain the required rotor rpm and a rotor thrust equal to the weight. When
this condition has been established, the acceleration will stop and the helicopter will continue
downwards at a steady rate of descent. If some outside influence causes the angle of attack to
increase, there will be an automatic reduction in the rate of descent, the reverse taking place if the
angle of attack is decreased.

10. Compared with a vertical autorotation in still air, the rate of descent will initially decrease with
forward speed, but beyond a certain speed the rate of descent will start to increase again. The cause
of this variation of rate of descent with forward
speed is the changing direction of the relative
airflow which occurs throughout the speed range
in autorotation.

11. Relative Airflow - Vertical Autorotation.


Consider a helicopter of a given weight requiring a
mean angle of attack of 8° to provide the required
rotor thrust and autorotative forces to maintain it in
a vertical autorotation, and assume that this angle
of attack is obtained when the rate of descent is
2,000 fpm. If the inflow angle is determined from
rate of descent and a mean rotational velocity, it
Fig 26-4 Inflow Angle and Rate of Descent
will be found to have a value of, say 10° (Fig 26-
Relationship
FIS Book 1: Helicopter Flight 486

4a) but because the action of the blades slows down the airflow coming from below the disc, the
actual inflow angle will be less, say, only 6° (Fig 26-4b). If the mean pitch value of the blade is 2°,
then the angle of attack will be 8°, which is the angle required. So 2,000 fpm rate of descent is
required by this particular helicopter to produce an inflow angle of 6°.

AUTOROTATION WITH FORWARD SPEED

Relative Airflow - Forward Autorotation

12. In determining the direction of the relative airflow when the helicopter is in a forward
autorotation, three factors must be taken into account. The effect of these factors on the inflow angle
will first be considered individually and then collectively.

13. Individual Effect.

(a) Factor A. To achieve forward


autorotation the disc must be tilted forward.
If the effective airflow from rate of descent
remains unchanged then the inflow angle
must decrease (Fig 26-5). The angle of
attack and therefore the rotor thrust must
also decrease, causing an increased rate of Fig 26-5 Inflow Angle - Disc Tilted
descent. Forward

(b) Factor B. When the helicopter is


moving forward, the disc will be subjected to
not only the descent airflow, but also to a
horizontal airflow. Because the disc is tilted
to this horizontal airflow, it will further reduce
the inflow angle (Fig 26-6). The angle of
Figure 26-6: Inflow Angle effect of
attack is further decreased therefore, horizontal airflow
causing an increased rate of descent.

(c) Factor C. When the helicopter


moves forward, the disc is moving into air
which has not been slowed down by the
action of the blades to the same extent as it
is when the helicopter is descending
vertically, therefore the effective rate of
descent airflow will increase, which will Fig 26-7: Inflow Angle - Effect of
Forward Speed
487 Autorotation

result in the inflow angle increasing (Fig 26-7). The angle of attack and rotor thrust increases,
giving a decreased rate of descent.

14. Combined Effect. At low forward speed only a small tilt of the disc is required and the
effect of factor C will be greater than the combined effects of factors A and B, so the inflow angle will
increase. Angle of attack, and therefore rotor thrust, will increase and the rate of descent will
decrease. As the rate of descent reduces, the inflow angle will decrease and the rate of descent will
stabilize again when the angle of attack is such that the value of rotor thrust equals the weight. As
forward speed is progressively increased, the effect of factor C will continue to increase the inflow
angle, but, similar to the induced flow in powered level flight, its effect is large initially but diminishes
with increasing forward speed. Since the disc has to be tilted more and more to overcome the rising
parasite drag from the fuselage, the combined effects of factors A and B rapidly increase with forward
speed. Therefore, a forward speed is eventually reached where the combined effects of factors A and
B equal C and balance out. When this occurs the helicopter will be flying at the speed to give
minimum rate of descent. Beyond this speed the effects of factors A and B will be greater than factor
C, inflow angle will therefore reduce and the required rotor thrust can only be obtained from a higher
rate of descent.

Rate of Descent Requirements in Autorotation

15. In autorotation, a rate of descent will be required to:

(a) Produce a rotor thrust equal to the weight.


(b) Provide an autorotative force for the selected rotor rpm.
(c) Produce a thrust component equal to parasite drag.

Autorotation for Endurance and Range in Still Air

16. Autorotation to give the maximum


time in the air must be at the speed to give
the minimum rate of descent. The speed for
endurance will therefore correspond to the
lowest part on the rate of descent curve (Fig
26-8). Maximum range will be achieved
when the helicopter is descending along its
shallowest flight path. This will be achieved
when flying at the best forward speed/rate of
descent ratio. Relating this to the rate of
descent curve, the optimum ratio will be at
the speed where a line drawn from the point Fig 25-8: Autorotation for Range and Endurance
FIS Book 1: Helicopter Flight 488

of origin of the graph is tangential to the rate of descent curve. For both range and endurance, rotor
rpm should be as quoted in the Aircrew Manual.

Flare

17. The flare effect in autorotation will be exactly the same as for a flare in powered flight. Rotor
rpm will rise because the increased inflow angle will cause the autorotative section to move further out
towards the tip, and increased rotor thrust will reduce the rate of descent while flare effects last.

Avoid Area for Autorotation

18. To establish fully developed autorotation, following power failure, it is vital to lower the lever
immediately, probably fully depending on forward speed and how quickly the lever is lowered after
power loss is detected. At low forward speed it may also be necessary to gain forward speed.
Lowering the lever and gaining forward speed will require considerable height loss before full
autorotation is established at a safe speed to execute an engine off landing. If power failure occurs
above optimum autorotation speed, flare may be used to recover Nr and reduce height loss as
autorotation is established. At high airspeed and low level there may be insufficient time to reduce
speed for a safe landing, despite the use of the lever and flare to maintain Nr and reduce height loss
as autorotation is established. Avoid areas, determined by test flying, are published in the relevant
aircraft Aircrew Manual; Fig 26-10 shows
an example. Power failure when
operating inside the avoid areas may
result in an unsuccessful engine off
landing as the aircraft may be too low
and too slow, or too low and too fast, to
establish full autorotation at a safe
speed for landing. Operation within the
avoid areas should be kept to a
minimum. The relevant Aircrew manual
should be consulted for specific
techniques following power failure. Fig 26-9: Typical Autorotation Avoid Areas

19. Autorotative Landing. When engine failure occurs at height, the aircraft has potential
energy to dissipate and this is converted into kinetic energy during the descent process in
autorotation. When near the ground, the kinetic energy stored in the rotor by virtue of its rpm is
converted into work, in the form of a large increase in rotor thrust, by use of the collective lever, with a
consequent rapid decay in Rrpm as the kinetic energy is used.

Вам также может понравиться