Vous êtes sur la page 1sur 115

Some notes on aircraft and spacecraft

stability and control


Michael Carley, m.j.carley@bath.ac.uk
A lonely impulse of delight
Drove to this tumult in the clouds; . . .
Before we begin . . .
These notes contain most, but not all, of the content of the course. You will also need:
ESDU, Lift-curve slope and aerodynamic centre position of wings in inviscid
subsonic ow, ESDU 70011.
Culick, F. E. C., The Wright brothers: First aeronautical engineers and test pi-
lots, AIAA Journal, 41(6):9851006, 2003.
Hefey, Robert K. and Jewell, Wayne F., Aircraft handling qualities data, NASA
CR-2144, 1972.
Thompson, Ambler and Taylor, Barry N., Guide for the use of the international
system of units (SI), NIST Special Publication 811, 2008.
Gratton, Guy and Newman, Simon, Towards the tumble resistant microlight,
In European Symposium of Society of Experimental Test Pilots, Dresden, Germany,
2125 June 2006. Available from http://eprints.soton.ac.uk/43858/
Hamilton-Paterson, James, Empire of the Clouds: When Britains Aircraft Ruled
the World, Faber & Faber, 2010.
You are responsible for nding and obtaining these documents. They will not be
distributed to the class. They can all be downloaded via the universitys systems. You
should print out the rst two. The third is long and we will not be using all of it so
wait until we come to use it before printing out the parts you need. You will only
need Appendix B of NIST SP-811. The microlight paper has a lot of information on
instability of both exible and rigid wing aircraft. Empire of the Clouds is an account
of British aeronautical engineering in the decade or so after 1945, and is well worth
reading as background to the present state of the industry.
You must also ll in and return the registration form for the ight test course.
Flight tests this year will be on the 30th and 31st of October. The ight test is a com-
pulsory element of the degree and essential for accreditation.
Contents
Contents i
List of Figures iv
I Static stability 1
1 How aircraft y 3
1.1 Equilibrium and stability . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Functions of aircraft controls . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Forces and moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Trim and stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Aerodynamic centre and neutral point . . . . . . . . . . . . . . . . . . . 7
Denitions of static and c.g. margins . . . . . . . . . . . . . . . . . . . . 9
1.5 Basic aerofoil and control characteristics . . . . . . . . . . . . . . . . . . 10
Aerofoils and wings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Control forces and moments . . . . . . . . . . . . . . . . . . . . . . . . . 11
Control hinge moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2 Longitudinal static stability 15
2.1 Some basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Downwash . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3 Stick xed stability and c.g. margins . . . . . . . . . . . . . . . . . . . . 18
2.4 Stick free stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.6 Stick free stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3 Flight testing 23
3.1 K
n
elevator angle to trim . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 What does the pilot feel? . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.3 K
n
tab angle to trim . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
4 Tailless aircraft 29
4.1 Stick xed stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 Static margin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
i
ii CONTENTS
5 Stick forces 33
5.1 Analysis to calculate stick forces . . . . . . . . . . . . . . . . . . . . . . . 33
5.2 More ight testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.3 Modication of stick forces . . . . . . . . . . . . . . . . . . . . . . . . . . 36
6 Manoeuvre stability 39
6.1 Analysis of a steady pullout . . . . . . . . . . . . . . . . . . . . . . . . . 39
6.2 Stick xed manoeuvre point . . . . . . . . . . . . . . . . . . . . . . . . . 42
6.3 Stick xed manoeuvre stability . . . . . . . . . . . . . . . . . . . . . . . 43
6.4 Stick free manoeuvre stability . . . . . . . . . . . . . . . . . . . . . . . . 43
6.5 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6.6 Tailless aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
6.7 Tailless aircraft manoeuvre point . . . . . . . . . . . . . . . . . . . . . . 46
6.8 Tailless aircraft manoeuvre margins . . . . . . . . . . . . . . . . . . . . . 47
6.9 Relationships between static and manoeuvre margins . . . . . . . . . . 47
Conventional aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
Tailless aircraft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6.10 Modication of stick free neutral and manoeuvre points . . . . . . . . . 48
7 Compressibility effects 51
7.1 High speed effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
II Dynamic stability 55
8 Dynamic behaviour of aircraft 57
8.1 Axes and notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
8.2 Aerodynamic derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
8.3 Longitudinal symmetric motion . . . . . . . . . . . . . . . . . . . . . . . 61
9 Normal modes of aircraft 63
Phugoid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Short period oscillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
9.1 Lateral motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
Dutch roll . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Spiral mode and roll subsidence . . . . . . . . . . . . . . . . . . . . . . . 66
9.2 Dihedral effect and weathercock stability . . . . . . . . . . . . . . . . . 67
Dihedral effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Weathercock stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
III Spacecraft dynamics and control 71
10 Getting around: Orbits 73
10.1 The two-body problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
CONTENTS iii
Elliptical orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
10.2 Orbital maneouvres . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Hohmann transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Orbital capture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
11 Getting things done: Spacecraft control 81
11.1 Attitude control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Gravity gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
Sun-synchronous orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
11.2 Manouevring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
IVProblems 85
12 Problems 87
Basic equations 101
List of Figures
1.1 Phases of ight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2 Terminology for study of stability . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Axes and sign conventions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Sign conventions for longitudinal stability . . . . . . . . . . . . . . . . . . . 7
1.5 Trim and stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Centre of pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.7 Incremental loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.8 Centre of gravity and aerodynamic centre relationships . . . . . . . . . . . 9
1.9 Lift curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.10 Loading due to control deection . . . . . . . . . . . . . . . . . . . . . . . . 11
1.11 Measurement of control/tab deections . . . . . . . . . . . . . . . . . . . . 12
1.12 Pressure distributions due to deections, a
1
> a
2
> a
3
. . . . . . . . . . . . . 12
1.13 Measurement of control surface area . . . . . . . . . . . . . . . . . . . . . . 13
1.14 Control hinge moments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.1 Stick xed stability (conventional aircraft) . . . . . . . . . . . . . . . . . . . 16
2.2 Trailing vortices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3 Effect of downwash on tailplane . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4 Stick free elevator conditions . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.1 A typical weight and balance envelope . . . . . . . . . . . . . . . . . . . . . 23
3.2 Elevator angle to trim at various lift coefcients . . . . . . . . . . . . . . . . 24
3.3 Measurement of neutral point location . . . . . . . . . . . . . . . . . . . . . 25
3.4 What the pilot experiences . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.5 Tab angle to trim at varying lift coefcients . . . . . . . . . . . . . . . . . . 27
3.6 Measurement of stick free neutral point location . . . . . . . . . . . . . . . 27
4.1 Canard conguration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 Control surfaces for tailless aircraft . . . . . . . . . . . . . . . . . . . . . . . 30
4.3 Representation of tailless aircraft . . . . . . . . . . . . . . . . . . . . . . . . 30
5.1 Aerodynamic assistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
6.1 Manoeuvre conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
6.2 Manoeuvre conditions for a tailless aircraft . . . . . . . . . . . . . . . . . . 45
iv
LIST OF FIGURES v
6.3 Aerodynamic forces during pitching motion . . . . . . . . . . . . . . . . . 46
6.4 Modication of neutral and manoeuvre points . . . . . . . . . . . . . . . . 48
6.5 Effects of positive springs and bob-weights . . . . . . . . . . . . . . . . . . 49
6.6 Modication of stick force per g . . . . . . . . . . . . . . . . . . . . . . . . . 49
7.1 Compressibility effects on lift curve slope and aerodynamic centre . . . . . 52
7.2 Compressibility effects on zero lift pitching moment and zero lift angle . . 52
7.3 Aeroelastic effects on lift curve slope . . . . . . . . . . . . . . . . . . . . . . 53
7.4 Aeroelastic effects on tailplane and elevator . . . . . . . . . . . . . . . . . . 53
7.5 Variation of downwash with Mach number . . . . . . . . . . . . . . . . . . 54
7.6 Variation of pitching moment with Mach number . . . . . . . . . . . . . . 54
7.7 Variation of stick forces with Mach number . . . . . . . . . . . . . . . . . . 54
8.1 Notation for analysis of dynamic stability . . . . . . . . . . . . . . . . . . . 58
9.1 Phugoid oscillation trajectory . . . . . . . . . . . . . . . . . . . . . . . . . . 64
9.2 Short period oscillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
9.3 Rolling subsidence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
9.4 Stability of the lateral modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
9.5 Dihedral effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
9.6 Wing sweep effects on L
v
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
9.7 Wing-fuselage interference effects on L
v
. . . . . . . . . . . . . . . . . . . . 69
9.8 Use of twin ns at high speed . . . . . . . . . . . . . . . . . . . . . . . . . . 70
9.9 Intake effects on N
v
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
10.1 The two-body problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
10.2 Polar coordinate system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
10.3 The geometry of an ellipse . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
10.4 Walter Hohmann and his transfer from low earth to geostationary Earth
orbit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
11.1 Two approaches for a docking spacecraft . . . . . . . . . . . . . . . . . . . 83
12.1 Aircraft with different centres of gravity . . . . . . . . . . . . . . . . . . . . 87
12.2 Full-scale and model aircraft. C
M
p
is measured by the balance, which re-
strains the model in pitch. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Part I
Static stability
1
Chapter 1
How aircraft y
Aircraft y by generating a lift greater than or equal to their weight. They do this by
holding a wing at a certain angle of attack, or incidence. Longitudinal control is the
study of how to set, maintain or change that angle of attack; stability is the study of
whether and how that angle of attack will remain xed when the aircraft is subjected
to small perturbations, due to atmospheric turbulence, for example.
One way to think about this is to look at the phases of aircraft ight, Figure 1.1.
When an aircraft takes off, its speed is quite low and it must generate lift greater than
its weight in order to leave the ground. In cruise, the aircraft operates at a constant
speed and constant lift. Finally, when the aircraft lands, it needs to reduce its speed
without losing too much lift.
In each case, the issue for control of the aircraft is how it can maintain its inci-
dence at a given speed. When it takes off or lands, it does so by rotatingraising
or lowering its nosein order to change the incidence of the wing, altering the rela-
tionship between speed and lift. Aircraft control is the study of how a pilot can x the
relationship between speed and incidence.
When an aircraft cruises, it is desirable that it do so at constant speed and inci-
dence, so the controls are at a xed setting. Aircraft stability is the study of how an
aircraft responds to small disturbances in ight and how it can be designed so that it
remains at a xed incidence and speed without overworking the pilot.
In each case, the basic question is how to generate a moment on the aircraft so that
it rotates and changes the wing incidence or so that the net moment is zero and the
aircraft ies at constant incidence. This is done via the aircraft controls. Before we go
any further, we need to clarify what we mean by some basic terms.
1.1 Equilibrium and stability
The requirements of an aircraft control system are that it be able to bring the aircraft
into some required equilibrium and that it be able to maintain that equilibrium stably.
This statement of requirements contains some terms which have precise mean-
ings:
3
4 CHAPTER 1. HOW AIRCRAFT FLY
Take-off: the incidence increases to generate more lift at low speed.
Cruise: the aircraft ies at constant incidence and speed
Landing: the aircraft increases incidence to allow it to slow down.
Figure 1.1: Phases of ight
equilibrium a system is in equilibrium when the sums of all of the forces and mo-
ments acting on it are identically zero;
static stability a system is statically stable if, when disturbed from equilibrium, it
initially tends to return to the equilibrium conguration;
dynamic stability a system is dynamically stable if, when disturbed from equilib-
rium, it does eventually return to the equilibrium conguration.
The distinction between the static and dynamic stability of a system is simple,
though subtle. If we disturb a system, static stability deals with the question of what
the system does in the very short time just after the disturbance has been applied;
dynamic stability is the study of what happens after that, over long periods of time.
Figure 1.2 shows the response of systems which are statically and dynamically stable
and/or unstable, in various combinations, including the case of a system which is
statically stable but dynamically unstable. It also includes the case of neutral stability,
where the system remains in whatever conguration it has been shifted to.
1.2 Functions of aircraft controls
The function of an aircraft control system is to provide a means of changing the mo-
ments on an aircraft, to control, in this case, its incidence. There are many ways of
1.2. FUNCTIONS OF AIRCRAFT CONTROLS 5
R
e
s
p
o
n
s
e
Time
Neutral stability
Statically and dynamically stable
Statically stable and dynamically unstable
Statically unstable
Figure 1.2: Terminology for study of stability
doing this, but for the conventional aircraft we consider for now, this is done by mov-
ing surfaces in order to change the aerodynamic forces on some part of the aircraft,
thereby changing the overall moment. These control surfaces are the:
elevator this changes the total lift on the tail when it is deected, causing a change
in pitching moment on the aircraft. This allows the pilot to adjust the aircraft
incidence;
ailerons these change the lift on each wing when they are deected. They move in
opposite directionsone goes up when the other goes down so that the lift on
one wing increases and the other decreases. This generates a change in rolling
moment and allows the aircraft to rotate about its axis to initiate turns, or allows
it to oppose disturbances due to crosswind or gusts;
rudder changes the side force on the vertical tailplane (or n), generating a change in
yawing moment, rotating the aircraft about a vertical axis. This can be used to
resist yawing moments due to engine failure and crosswind and to aid in spin
recovery and turn co-ordination.
In steady, level ight in still air, the rudder and ailerons will be undeected while
the elevator will probably be at some deection which depends on the aircraft load-
ing. Under other conditions, or during a manouevre, all three controls may be used
simultaneously. The controls can be operated directly by the pilot, through a system
of mechanical actuators, possibly with aerodynamic or power assistance, or controls
may be fully powered using a hydraulic or electrical system. These systems can be
mechanically or electronically controlled (y by wire or y by light).
The sign conventions for the controls and motions are shown in Figure 1.3.
6 CHAPTER 1. HOW AIRCRAFT FLY
Positive up
Positive down
Positive left
Positive down
z
Yaw
x
Roll
y
Pitch
Figure 1.3: Axes and sign conventions
1.3 Forces and moments
Forces and moments on an aircraft are due to the mass of the aircraft, a function of
how it is built and loaded, and the aerodynamic forces generated in ight. The mass
distribution of the aircraft gives rise to inertial forces and moments as described by
Isaac Newton and to gravitational forces.
You should already know that the basic aerodynamic forces include the lift and
drag on the aircraft. More generally, we consider the static forces and moments due
to linear velocities; the damping forces and moments caused by angular velocities
(such as pitching and rolling); the control forces generated when the pilot operates a
control surface.
By considering the lateral symmetry of most aircraft, it is clear that in forward
level ight, the forces on the aircraft act in the plane of symmetry. This means that
any symmetric disturbance, such as operation of the elevator, will generate horizontal
and vertical motion of the aircraft and rotation in the vertical plane only. At this
point, then, we study longitudinal symmetric motion of the aircraft and consider
longitudinal stability.
1.4 Trim and stability
Figure 1.4 shows the basic conguration for the study of stability of an aircraft, la-
belled with the forces and moments and showing the corresponding sign conven-
tions. The orientation of the aircraft is labelled with two angles, and . You should
1.4. TRIM AND STABILITY 7
not confuse these. The angle is the inclination of the aircraft which is the angle be-
tween the direction of ight and the horizontal; the angle is the incidence, or angle
between the direction of ight and the Zero Lift Line (ZLL). When is zero, the Zero
Lift Line is aligned with the ight direction and there is no lift acting on the aircraft,
whatever might be its inclination .
Horizontal
F
lig
h
t
d
ir
e
c
t
io
n

Z
ero
lift
lin
e

W
L
D
T
Figure 1.4: Sign conventions for longitudinal stability
To examine the equilibrium and stability of the aircraft, we resolve forces parallel
and perpendicular to the aircraft axis:
Parallel: T D W sin = 0, (1.1)
Perpendicular: L W cos = 0, (1.2)
Moments about the c.g.: M
cg
= 0. (1.3)
Moments about the centre of gravity (c.g.) cannot be due to the mass of the aircraft
(by denition). This means that if M
cg
= 0, the aerodynamic moments on the aircraft
are in equilibrium and the aircraft is said to be trimmed or in trim. This is the basic
problem of control: is it possible, using the tailplane, to generate a pitching moment
so that the overall moment about the centre of gravity is zero?
The basic problem of static stability is then: when an aircraft in trim is subjected
to a disturbance which changes its incidence, does it tend to return to the equilibrium
position?
This can be restated: if the aircraft pitches nose up, the change in aerodynamic
moment about the centre of gravity M
cg
should be negative in order to push the
nose back down or, M
cg
/ < 0.
Figure 1.5 shows various ways M
cg
can vary with , including how it is possible
to trim an unstable aircraft and how is possible for an aircraft to be stable without
being able to trim at a useful incidence.
Aerodynamic centre and neutral point
The forces and moments acting on an aircraft depend on the shape of the aircraft and
not on the position of the centre of gravity so we can consider the aerodynamic loads
8 CHAPTER 1. HOW AIRCRAFT FLY
C
M
c
g

Neutrally stable
cannot trim
C
M
/ > 0
C
M
/ < 0
C
M
/ < 0
C
L
< 0
Figure 1.5: Trim and stability
separately from the gravitational. Figure 1.6 shows a pressure distribution on an
aerofoil section. The loads can be considered to act at a point, the centre of pressure,
where the total aerodynamic moment is zero. The centre of pressure, however, moves
as the incidence varies, so it is not very useful as a reference point in calculations
involving changing incidence.
L
Figure 1.6: Centre of pressure
To make life easier, we can give up the requirement that the moment about our
reference point be zero and, instead, allow it to have some nite value as long as the
reference point is xed and the moment is constant. We can do this by looking at the
incremental pressure distribution, sketched in Figure 1.7.
When the incidence is increased by some small amount, the incremental aerody-
namic load can be considered to act through a certain point, generating no change
in moment about that position. This point is the aerodynamic centre and is the point
about which dM/d 0.
If we now think about the basic question of stability, we can consider what hap-
pens to an aircraft which pitches slightly nose up. Depending on the position of the
centre of gravity, relative to the aerodynamic centre, the aircraft will be stable, unsta-
ble or neutrally stable, Figure 1.8.
1.4. TRIM AND STABILITY 9
L
Figure 1.7: Incremental loads
W
L
M
0
W
L
M
0
W
L
M
0
Figure 1.8: Centre of gravity and aerodynamic centre relationships
1. If the centre of gravity is forward of the aerodynamic centre, dM
cg
/d is nega-
tive and the aircraft is statically stable.
2. If the centre of gravity is aft of (behind) the aerodynamic centre, dM
cg
/d is
positive and the aircraft is statically unstable.
3. If the centre of gravity is at the aerodynamic centre, dM
cg
/d is zero and the
aircraft is neutrally stable.
When we talk about the properties of a whole aircraft, rather than just a wing or
aerofoil, we use the term neutral point. This is the position of the centre of gravity
for which the aircraft is neutrally stable. It is a purely aerodynamic property, which
depends on the shape of the aircraft. For a wing, the neutral point and the aerody-
namic centre are identical.
Denitions of static and c.g. margins
Abasic measure of stability is howquickly an aircraft responds to a perturbation. This
is equivalent to the gradient C
M
/. When this is large, the aircraft experiences a
large pitching moment when it is perturbed and quickly returns to its equilibrium.
Our measure of the innate stability of an aircraft is then the static margin:
K
n
=
dC
M
cg
dC
R
(1.4)
10 CHAPTER 1. HOW AIRCRAFT FLY
where C
R
= (C
2
L
+C
2
D
)
1/2
, the resultant force.
We can express this in a more easily visualized form by looking at the c.g. margin,
H
n
, which is the distance between the neutral point and the centre of gravity, scaled
on the mean chord c:
H
n
=
dC
M
dC
L
= h
n
h (1.5)
where hc is the displacement of the centre of gravity aft of the reference point and h
n
c
is the displacement of the neutral point aft of the reference point.
At low speed and low inclination, C
R
C
L
and C
L
, C
M
, C
R
are not inuenced by
Mach number or aeroelastic effects so that H
n
K
n
.
1.5 Basic aerofoil and control characteristics
So far, we have considered aircraft stability without considering the behaviour of
aircraft. In order to deal with realistic problems, we need to know something about
how aerodynamic surfaces and bodies behave in ight.
Aerofoils and wings
As always, we work in terms of non-dimensional quantities using standard reference
values velocity V , density , wing planform area S and wing mean chord c:
C
L
=
L
V
2
S/2
, C
D
=
D
V
2
S/2
, C
M
=
M
V
2
Sc/2
. (1.6)
For tailless aircraft, the wing root chord c
0
is often used as a reference length. The
subscript in each coefcient is upper case because the coefcients are those for three
dimensional bodies.

C
L
Figure 1.9: Lift curve
1.5. BASIC AEROFOIL AND CONTROL CHARACTERISTICS 11
Figure 1.9 shows the lift curve slope of a wing. The important point to note is
that we choose a reference incidence such that the lift is zero when = 0. This will
not always be true in other calculations and you should check the conventions used
when you take data from published sources. In this course, we will only consider
linear aerodynamics, i.e. the part of the lift curve where dC
L
/d is constant. This is a
reasonable assumption for most aircraft most of the time, but will not be correct near
stall or in high speed manoeuvres.
Control forces and moments
We control an aircraft by moving a control surface such as an elevator. In order to
have some clue what the result of a control input will be, we need to know how far
the surface must be moved in order to generate, indirectly, the pitching moment we
need and we need to know what moment will be needed to rotate the control surface
about its hinge point.
If the incidence of the wing or tailplane as a whole is held constant, deecting the
control surface in the positive sense is equivalent to introducing extra camber. This
generates extra lift and a pitching moment which is usually nose-down, because of
the form of the incremental pressure distribution, Figure 1.10.
x/c
C
p
Figure 1.10: Loading due to control deection
This means that a positive control deection generates a positive change in lift and
negative pitching moment (if the tail pushes up, it forces the nose down). If there is a
tab, an extra small surface on the end of the elevator, this too will generate positive lift
and negative moment. Figure 1.11 shows the notation for control surface deection.
Since we are working on the basis of linear aerodynamics, each deection con-
tributes linearly to the forces and moments, with the following symbols dened for
convenience:
a
1
=
C
L
T

T
, a
2
=
C
L
T

, a
3
=
C
L
T

, (1.7)
with the corresponding pressure distributions shown in Figure 1.12.
12 CHAPTER 1. HOW AIRCRAFT FLY

Figure 1.11: Measurement of control/tab deections


a
1
a
2
a
3
C
p
x/c
C
p
x/c
C
p
x/c

Figure 1.12: Pressure distributions due to deections, a


1
> a
2
> a
3
.
The total tailplane lift coefcient is then:
C
L
T
= a
1

T
+a
2
+a
3

and
C
Mac
= C
M
0
+
C
M
0

+
C
M
0

.
Control hinge moments
When an aircraft control system is designed, we will need to know what force is
required to move a control, whether it is being moved directly by the pilot or through
a powered actuation system. The force needed depends on the moment required to
rotate the surface through a given deection angle. The hinge moment coefcient is:
C
H
=
M
H
V
2
S

/2
, (1.8)
1.5. BASIC AEROFOIL AND CONTROL CHARACTERISTICS 13
Hinge line
Aerodynamic balance
S

Figure 1.13: Measurement of control surface area


where S

is the control surface area and c

is the control surface mean chord. Both of


these values are measured aft of the hinge line, as shown in Figure 1.13.
As before, we adopt a shorthand notation for the contribution of each deection
to the total hinge moment:
b
1
=
C
H

T
, b
2
=
C
H

, b
3
=
C
H

, (1.9)
and note that for non-symmetric tailplane sections there is usually a hinge moment
when all other deections are zero, given the symbol b
0
. The pressure distributions
associated with each of these terms are shown in Figure 1.14.
The total hinge moment coefcient is then:
C
H
= b
0
+b
1

T
+b
2
+b
3
. (1.10)
14 CHAPTER 1. HOW AIRCRAFT FLY
b
1
b
2
b
3
H
i
n
g
e
l
i
n
e
C
p
x/c
H
i
n
g
e
l
i
n
e
C
p
x/c
H
i
n
g
e
l
i
n
e
C
p
x/c

Figure 1.14: Control hinge moments


Chapter 2
Longitudinal static stability
Given the denitions and background information of Chapter 1, we are in a position
to start doing some calculations for the stability of aircraft. We consider two basic
cases, the stick xed and stick free. In the rst case, conceptually, we move the
elevator to the position required for trim and then x the stick so that the elevator
remains at the set deection, ignoring the question of what force is required to keep
it in place. The stability problem can then be phrased: with the stick xed, how does
the aircraft respond to a perturbation? In the stick free case, we use the tab to adjust
the moment on the elevator so that it comes to an equilibrium position which trims
the aircraft. In this case, zero force is required to keep the control xed. Using this
zero-force case as a reference, we can work out the stick force required for any other
tab setting.
2.1 Some basics
Figure 2.1 shows the problem of stick xed stability for a conventional aircraft, i.e.
one with a tail at the back. The aircraft has two lifting surfaces, the wing which
generates most of the lift, and the tailplane which generates a small amount of lift but
which can be adjusted to change the pitching moment on the aircraft as a whole. The
tailplane is set at angle
T
relative to the aircraft zero lift line.
On the usual assumption of linear aerodynamics, with small and
T
and no wake
effect on the tailplane (but see 2.2):
M
cg
= M
0
L
WBN
(h
0
h)c L
T
((h
0
h)c +l) Tz
T
+Dz
D
= M
0
(h
0
h)c(L
WBN
+L
T
) L
T
l Tz
T
+Dz
D
= M
0
(h
0
h)cL L
T
l Tz
T
+Dz
D
,
where the lift has been broken up into a wing-body-nacelle (WBN) and a tailplane (T)
contribution. Given that lift is much larger than drag (and likewise thrust), we can
neglect Tz
T
and Dz
D
, and
M
cg
= M
0
(h
0
h)cL L
T
l.
15
16 CHAPTER 2. LONGITUDINAL STATIC STABILITY
Datum
W
h c
L
WBN
h
0
c
M
0
L
T
l
T
z
T
D
z
D
Figure 2.1: Stick xed stability (conventional aircraft)
We now non-dimensionalize the equation to write it in terms of the coefcients de-
ned in 1.5, with the addition of the tailplane lift coefcient:
C
L
T
=
L
T
V
2
S
T
/2
(2.1)
where S
T
is the area of the tailplane.
Thus, dividing by Sc(V
2
/2):
C
M
cg
= C
M
0
(h
0
h)C
L
C
L
T
S
T
S
l
c
and dening the tail volume coefcient:
V =
S
T
S
l
c
(2.2)
we nd:
C
M
cg
= C
M
0
(h
0
h)C
L
V C
L
T
. (2.3)
This is the fundamental equation of aircraft stability and control. In control problems,
the aircraft is trimmed with C
M
cg
0, the lift coefcient is known from the operating
conditions and C
M
0
is known from the aircraft geometry. Then, if V is known, C
L
T
can be calculated and from that the elevator deection; if C
L
T
is known (because
the tailplane shape has already been decided), V can be calculated, and the size of
the tailplane xed. The tail volume coefcient represents the effectiveness of the
tailplane at generating a moment. It contains the size of the tailplane S
T
and the lever
arm l which, combined, tell us the moment which the tailplane can generate.
2.2 Downwash
Any lifting wing generates a downwash, due to the trailing vortex system, shown in
Figure 2.2. This needs to be included in the stability calculation because it alters the
2.2. DOWNWASH 17
Figure 2.2: Trailing vortices
incidence at the tailplane and that change in incidence changes with aircraft angle of
attack, Figure 2.3.
Z
L
L
t
a
i
l
p
l
a
n
e
Z
L
L
W
B
N

T
Free stream

R
e
s
u
lt
a
n
t

o
w
Figure 2.3: Effect of downwash on tailplane
From Figure 2.3:

T
= +
T
.
For an untwisted wing, the downwash angle is proportional to the lift on the wing,
meaning that in the linear regime, it is also proportional to :
=
d
d
+
0
,
18 CHAPTER 2. LONGITUDINAL STATIC STABILITY
with
0
only present for a wing where the zero lift angle of attack varies along its
span (i.e. a wing with a varying cross-section or camber along its length or with
twist). Combining the previous equations:

T
= +
T

0
+
d
d

_
=
_
1
d
d
_
+ (
T

0
).
From1.5:
C
L
T
= a
1

T
+a
2
+a
3
,
so that:
C
L
T
= a
1
( +
T
) +a
2
+a
3
,
and:
C
L
T
= a
1

_
1
d
d
_
+a
1
(
T

0
) +a
2
+a
3
.
This can be related to known quantities by including the relationship between inci-
dence and lift coefcient (Figure 1.9):
C
L
= a,
where a is the overall lift curve slope of the aircraft. Upon substitution:
C
L
T
=
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
2
+a
3
. (2.4)
In deriving (2.3), we made no assumptions about how C
L
T
was generated, so (2.4)
can be substituted to give:
C
M
= C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
2
+a
3

_
.
Given a ight condition (speed, aircraft weight, etc.), this equation allows us to cal-
culate the elevator angle to trim, (trim quantities are written as the usual symbol
with an overbar). In designing aircraft, there will be a limit on the maximum elevator
deection. Given this maximum , we can use the ight conditions to estimate V and
so size the tailplane.
2.3 Stick xed stability and c.g. margins
The static margin is dened in 1.4:
K
n
H
n
=
dC
M
dC
L
,
2.4. STICK FREE STABILITY 19
so that to determine the stability of the aircraft, we differentiate (2.3) with respect to
C
L
:

dC
M
dC
L
= (h
0
h) +V
a
1
a
_
1
d
d
_
.
The neutral point is the centre of gravity position where dC
M
/dC
L
0:
h
n
= h
0
+V
a
1
a
_
1
d
d
_
,
where h
0
is the neutral point of the aircraft less tail. Adding a tail has moved the
neutral point back by an amount V (a
1
/a)(1 d/d), increasing the stability (as you
might expect). The fundamental problem of designing the control system of an air-
craft is that of determining, via V , the size of the tailplane such that it makes the whole
aircraft stable and controllable. The stability requirement is specied as a minimum
value of h h
n
; the control requirement is stated, in effect, as a maximum pitching
moment to be generated.
2.4 Stick free stability
The stick-xed analysis of the rst part of this chapter only considers where the el-
evator needs to be in order to trim the aircraft, without looking at the forces needed
to get there. As the rst step towards calculating the forces needed to move a control
surface, we consider the stick-free problem, where the elevator is allowed to op
around until it reaches an equilibrium position where the moment on it, and so the
stick force, is zero. On small aircraft, the moment on the elevator is modied aerody-
namically by moving the tab, on large ones, the elevator is moved by an actuator and
no tab is required. In both cases, however, we need to knowthe forces on the elevator
either to keep them within the limits of a pilots strength or to size the actuators.
2.5 Analysis
The hinge moment coefcient is:
C
H
= b
0
+b
1

T
+b
2
+b
3
,
which is zero if the elevator (stick) is free. In this case, the elevator angle is:
=
b
0
+b
1

T
+b
3

b
2
.
From2.2:
C
L
T
=
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
2
+a
3
,
20 CHAPTER 2. LONGITUDINAL STATIC STABILITY
and
C
L
T
=
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
)
a
2
b
2
(b
0
+b
1

T
+b
3
) +a
3
.
Incorporating the expression for from2.2, page 18)

T
=
C
L
a
_
1
d
d
_
+ (
T

0
),
and
C
L
T
=
_
a
1

a
2
b
1
b
2
__
1
d
d
_
C
L
a
+
_
a
1

a
2
b
1
b
2
_
(
T

0
)
+
_
a
3

a
2
b
3
b
2
_

a
2
b
0
b
2
.
To simplify our notation, we dene two new variables (both given on the data
sheet in the appendix):
a
1
= a
1
_
1
a
2
b
1
a
1
b
2
_
,
a
3
= a
3
_
1
a
2
b
3
a
3
b
2
_
,
so that:
C
L
T
= a
1
_
1
d
d
_
C
L
a
+a
1
(
T

0
) +a
3

a
2
b
0
b
2
.
We already know the tailplane lift coefcient, (2.3), so that we can nd the lift if
the stick is released and the elevator comes to equilibrium. The resulting pitching
moment is:
C
M
= C
M
0
(h
0
h)C
L
V C
L
T
,
and:
C
M
= C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
3

a
2
b
0
b
2
_
,
which allows us to nd the tab angle to trim with zero stick force, .
2.6 Stick free stability
To nd the stick-free stability properties, we differentiate the pitching moment equa-
tion:
dC
M
dC
L
= (h
0
h) V
a
1
a
_
1
d
d
_
,
2.6. STICK FREE STABILITY 21
and the static margin stick free:
K

n
=
dC
M
dC
L
= (h
0
h) +V
a
1
a
_
1
d
d
_
.
This is related to the static margin stick xed:
K
n
= (h
0
h) +V
a
1
a
_
1
d
d
_
,
with the two static margins being equal if:
a
1
= a
1
= a
1
_
1
a
2
b
1
a
1
b
2
_
,
or
a
2
b
1
a
1
b
2
= 0.
Since a
1
and a
2
are both positive, and b
2
is negative for correct feel of the elevator, this
can only happen if b
1
= 0, which can be achieved using aerodynamic balancing or by
moving the elevator hinge line, which will also change b
2
.
a
2
b
1
/a
1
b
2
> 0 a
2
b
1
/a
1
b
2
< 0 a
2
b
1
/a
1
b
2
= 0
Figure 2.4: Stick free elevator conditions
Figure 2.4 shows the three possible cases for b
1
:
when a
2
b
1
/a
1
b
2
> 0, the aircraft is less stable stick freethe elevator is conver-
gent;
when a
2
b
1
/a
1
b
2
< 0, the aircraft is more stable stick freethe elevator is diver-
gent;
when a
2
b
1
/a
1
b
2
= 0, the aircraft is equally stable stick freethe elevator is null.
We dene the neutral point stick free h

n
and the static and c.g. margins stick-free
H

n
in the same way as in the stick-xed case.
Chapter 3
Flight testing
Having designed an aircraft to have given stability characteristics, we must test the
production model to nd what the real behaviour is. In the early stages of design, we
use approximate analyses and semi-empirical methods (for example, ESDUsheets) to
estimate the aerodynamic parameters such as lift curve slopes, largely because early
in design, we have not xed the exact shape and size of the aircraft or its subsystems.
When we have a detailed geometry, we can use computational methods to rene our
estimates. When the rst few aircraft are produced, we must test them to see what
the real behaviour of the real aircraft is.
This information is used in setting the limits to be observed in servicethe ight
envelope of Figure 3.1. Before ight, the aircraft weight and centre of gravity are
plotted on the diagram and must lie within the limits indicated. If they do not, then
the weight must be reduced or the centre of gravity must be moved by adding ballast.
This guarantees that the aircraft will y within the limits set at the design stage.
3.1 K
n
elevator angle to trim
Given that, for a trimmed aircraft:
C
M
= C
M
0
(h
0
h)C
L
V C
L
T
,
h c
m
Aircraft safe to y
Figure 3.1: A typical weight and balance envelope
23
24 CHAPTER 3. FLIGHT TESTING
and,
C
L
T
=
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
2
+a
3
,
C
M
= 0 = C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
2
+a
3

_
.
This can be differentiated (2.3) to examine the stick-free stability:
K
n
H
n
=
C
M
C
L
= (h
0
h) +V
a
1
a
_
1
d
d
_
.
We also know that:
=
1
V a
2
_
C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
3

__
,
and that there is a relationship between and K
n
because:
d
dC
L
=
1
V a
2
_
(h
0
h) +V
a
1
a
_
1
d
d
__
=
K
n
V a
2
.
Figure 3.2 shows plotted against C
L
, while Figure 3.3 shows the relationship
between d/dC
L
and h.
C
L

h
1
h
2
h
3
c.g. forward
Figure 3.2: Elevator angle to trim at various lift coefcients
Given this information, one way of nding the aircraft neutral point stick-xed
is: y the aircraft straight and level at various speeds, recording the elevator angle to
trim. This is repeated for various different centre of gravity positions, yielding a plot
like Figure 3.2. To nd the neutral point, plot the gradients of the lines of Figure 3.2,
as in Figure 3.3. Extrapolating to d/dC
L
gives the centre of gravity position where
K
n
= 0, the neutral point.
3.2. WHAT DOES THE PILOT FEEL? 25
h
d /dC
L
h
1
h
2
h
3
Figure 3.3: Measurement of neutral point location
3.2 What does the pilot feel?
Pilots rarely knowthe lift coefcient of the aircraft: they will have a feel for stick force
and for elevator deection (because they know how far the stick has moved) and for
speed (because they can see out the window or look at the instruments). We can see
how the elevator angle to trim varies with speed, to look at what the pilot feels in
ying the aircraft:
d
dV
=
d
dC
L
dC
L
dV
.
By denition,
C
L
=
L
V
2
S/2
,
so that:
dC
L
dV
=
2L
V
3
S/2
=
2C
L
V
,
and
d
dV
=
2C
L
V
d
dC
L
=
2C
L
V
K
n
V a
2
,
which is sketched in Figure 3.4.
FromFigure 3.4, it is clear that the aircraft is uncontrollable belowsome minimum
ight speedit is not possible to move the elevator far enough to trim. This happens
because at lowspeed, the control surfaces cannot generate enough force to balance the
moment about the centre of gravity. Likewise, above a certain speed, small changes
in lead to large changes in trim speed and the aircraft is also very hard to control.
The useful range of speeds for an aircraft lies between these two limits, although the
limits in question will be a function of the aircraft type and of the skill assumed of the
pilot.
26 CHAPTER 3. FLIGHT TESTING
V

c.g. moving forward
Figure 3.4: What the pilot experiences
3.3 K
n
tab angle to trim
To nd the neutral point stick-free, we can use the same approach as in the stick-xed
case, but using the tab to trim, rather than the elevator. We know that:
C
M
= C
M
0
(h
0
h)C
L
V C
L
T
= 0,
and
C
L
T
=
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
2
+a
3
.
From2.5:
=
b
0
+b
1

T
+b
3

b
2
,
and C
L
T
=
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
3

a
2
b
0
b
2
,
so that:
C
M
= C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
)
a
2
b
0
b
2
_
.
Rearranging to nd the tab angle to trim:
=
1
V a
3
_
C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
3

a
2
b
0
b
2
__
,
and differentiating with respect to C
L
:
d
dC
L
=
1
V a
3
_
h
0
h +V
a
1
a
_
1
d
d
__
.
3.3. K
N
TAB ANGLE TO TRIM 27
We know, however, that:
K

n
= (h
0
h) +V
a
1
a
_
1
d
d
_
,
and that
d
dC
L
=
K

n
V a
3
.
So to nd the neutral point stick free, we vary the aircraft speed at xed centre of
gravity, trimming with the tab, giving us Figure 3.5. We then plot the gradients from
that gure against C
L
, Figure 3.6 to nd h

n
C
L

h
1
h
2
h
3
Figure 3.5: Tab angle to trim at varying lift coefcients
h
d

/dC
L
h
1
h
2
h
3
Figure 3.6: Measurement of stick free neutral point location
Chapter 4
Tailless aircraft
In the previous chapters, we have considered conventional aircraft, those which use
a tail to provide pitching moment control. We can use the same methods to analyse
canard aircraft which have the control surface ahead of the wing. In this case, the
basic equations are the same as in the conventional case, but the tail arm l about the
aerodynamic centre is negative, Figure 4.1.
L
F
L
WBN
l
M
0
W
h c
h
0
c
Figure 4.1: Canard conguration
On a tailless aircraft, there is no separate control surface for pitch control, with the
elevators and ailerons being combined into elevons. These are moved in opposite
directions for roll control and in the same direction for pitch control, Figure 4.2.
4.1 Stick xed stability
Figure 4.3 shows how a tailless aircraft is represented for the study of static stabil-
ity. The most obvious difference from the conventional case is that we have no tail
contribution to include. On tailless aircraft, the pitching moment required to trim is
generated by the control surfaces which also have a large effect on lift. This makes
the control of such aircraft quite complicated, especially on landing.
29
30 CHAPTER 4. TAILLESS AIRCRAFT
Elevon
Rudder
Figure 4.2: Control surfaces for tailless aircraft: the elevons operate together for pitch
control and differentially for roll

L
WBN
M
0
W
h c
h
0
c
Figure 4.3: Representation of tailless aircraft
4.2. STATIC MARGIN 31
The lift coefcient for a tailless aircraft is written:
C
L
= a
1
+a
2
,
with no tab included, since tailless aircraft rarely have them.
Taking moments about the centre of gravity:
M
cg
= M
0
+
M
0

(h
0
h)c
0
L,
and non-dimensionalizing:
C
M
= C
M
0
+
C
M
0

(h
0
h)C
L
,
which can be rearranged to yield the elevon angle to trim.
4.2 Static margin
The denition of static margin is the same as in the conventional aircraft case:
K
n
=
dC
M
dC
L
= h
0
h.
Because tailless aircraft are usually large and have large control surfaces, their
control systems are powered, so that there is no stick free condition: we do not need
to consider this case.
Chapter 5
Stick forces
So far, we have not considered howwhat force or, equivalently, torque, will be needed
to move a control surface into position or to hold it in place. This is an important ques-
tion because it must be possible for the pilot to control the aircraft without requiring
excessive stick force. On the other hand, the aircraft must not be too twitchy, respond-
ing excessively to small control inputs. If the controls are powered, it is also essential
to know what forces the actuators will need to generate, so that the hydraulic sys-
tem can be sized. Table 5.1 shows the maximum forces which can be applied to the
different controls under various circumstances.
Aileron Elevator Rudder
Stick Wheel Stick Wheel (Push)
Maximum all-out effort 2 hands 400 530 800 980 1780 N
Maximum permissible effort 2 hands 360 440 440 890 N
1 hand 220 220 310 310 N
Maximum comfortable effort 2 hands 130 180 270 N
1 hand 90 90 130 130 N
Largest full travel 254 508 230 230 126 mm
Table 5.1: Maximum control forces
It is considered good practice to make sure that the maximum elevator force is
higher than the maximum aileron force and that the maximum rudder force is higher
than both. The controls are said to be harmonized if the aileron, elevator and rudder
forces have the ratio 1:2:4 for a given control response. For example, the rudder force
for a 10

/s yaw is twice the elevator force for a 10

/s pitch.
5.1 Analysis to calculate stick forces
The input from the pilot for a given elevator deection is the stick force, so that the
stick force to trim P
e
, which is found from the moment on the control multiplied by
33
34 CHAPTER 5. STICK FORCES
the gearing ratio between the stick and the control deection m
e
, is:
P
e
= m
e
V
2
2
S

C
H
.
To make the aircraft controllable, then, the stick force to trim must lie within rea-
sonable limits: too high and the pilot will not be able to move the elevator over the
full range of deections needed; too low and a small stick deection will generate
a large acceleration on the aircraft with a risk of overloading the structure. To start
with, we need the hinge moment to trim, which we can derive from our previous
analysis of the tab angle to trim, 2.5.
From the denition of hinge moment coefcient:
C
H
= b
0
+b
1

T
+b
2
+b
3
,
we can rearrange to nd as a function of C
H
= 0:
=
C
H
b
0
b
1

T
b
3

b
2
.
From the data sheet:
C
L
T
= a
1
_
1
d
d
_
C
L
a
+a
1
(
T

0
) +a
2
+a
3
,
and, on substituting :
C
L
T
= a
1
_
1
d
d
_
C
L
a
+a
1
(
T

0
) +a
3
+
a
2
b
2
(C
H
b
0
).
This equation is the general formof one we have already derived for the zero stick
force case. Given that:
C
M
= C
M
0
(h
0
h)C
L
V C
L
T
,
and substituting for C
L
T
:
C
M
= C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
3
+
a
2
b
2
(C
H
b
0
)
_
,
which can be re-arranged to nd :
V a
3
= C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
)
a
2
b
0
b
2
_
, (5.1)
or hinge moment to trim:
V
a
2
C
H
b
2
= C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
3

a
2
b
0
b
2
_
.
(5.2)
5.2. MORE FLIGHT TESTING 35
We could use (5.2) to work out the hinge moments directly, but it is more conve-
nient to use the stick-free case as a reference. Subtracting (5.2) from (5.1):
V
_
a
3

a
2
b
2
C
H
_
= V a
3
,
yielding:
C
H
=
b
2
a
2
a
3
( )
so that:
P
e
= m
e
V
2
2
S

b
2
a
2
a
3
( ),
so that the stick force to trim is proportional to the difference between the actual tab
angle and the tab angle to trim.
5.2 More ight testing
In theory we might use the stick forces to calculate the stick free neutral point, begin-
ning from:
C
H
=
b
2
a
2
a
3
( ),
which, under differentiation with respect to lift coefcient at constant tab angle, yields:
C
H
C
L
=
b
2
a
2
a
3

C
L
.
In 3.3, we found that:
d
dC
L
=
K

n
V
1
a
3
,
so that:
dC
H
dC
L
=
b
2
K

n
V a
2
.
In principle, by measuring the stick force or hinge moment at different ight con-
ditions, we can work out the stick free neutral point. In practice, however, we cannot
measure the force accurately enough for a reliable estimate, due to errors introduced
by such things as friction in the system.
36 CHAPTER 5. STICK FORCES
5.3 Modication of stick forces
There are three main methods which can be used to modify the stick forces to bring
them into the correct range for control of the aircraft:
Gearing between stick and control surface, but this is limited because of the range of
elevator movement required.
Power assistance, which can share the load or supply all of the force required to
move the control, with a feedback system to give the pilot feel for the control
input.
Aerodynamic methods of modifying the loads include adding surface ahead of the
hinge line (a horn balance), moving the hinge line or adding tabs which are
geared to the elevator, Figure 5.1.
Hinge line
Horn balance
S

a: horn balance b: hinge location


c: geared tab d: anti-balance tab
Figure 5.1: Aerodynamic assistance
Aerodynamic balancing is a means of changing the hinge moment required for a
given elevator deection, dP
e
/d:
P
e
= m
e
V
2
2
S

C
H
,
but,
C
H
= b
0
+b
1

T
+b
2
+b
3
,
5.3. MODIFICATION OF STICK FORCES 37
so that
dP
e
d
= m
e
V
2
2
S

b
2
.
To reduce the stick force, we want to reduce b
2
, but dP
e
/d b
2
, must be negative for
correct feel of the controls. Reducing b
2
is useful at high speed (because of the effect
of V
2
) but at low speed, the pilot may not have enough feel for the controls and other
methods of reducing the stick force may be needed.
Chapter 6
Manoeuvre stability
We have now completed our analysis of straight and level static stability. The next
step is to examine longitudinally symmetric manoeuvres (i.e. manoeuvres that affect
the left and right hand side of the aircraft equally). The most straightforward example
of this is a steady pullout at constant velocity. Somewhat surprisingly, the elevator
angle required for pitch trim in a steady banking turn can also be calculated in the
same way. This is because the radius of a typical banked turn is very large. Hence,
the asymmetry in the ow is small once the turn has been initiated.
6.1 Analysis of a steady pullout
Consider two conditions, shown in Figure 6.1:
1. an aircraft in steady, level ight at speed V ;
2. the same aircraft in a steady pullout at speed V .
In the steady pullout the aircraft has a radial (centripetal) acceleration V
2
/r = ng.
The difference in lift between (1) and (2) is nmg = nW. Hence:
C
L
= nC
L
=
nW
V
2
S/2
where C
L
is the lift coefcient in the straight and level case.
The other key difference between the two cases is that in the steady pullout the
aircraft has an angular (pitch) velocity as well as a linear velocity. This angular ve-
locity can easily be found by considering the amount of time that the aircraft would
take to complete a full circle at constant speed. If the aircraft completed a full circle it
would pitch though 2 radians and would therefore cover a distance of 2r, where r
is the radius of the circle. The time taken, t, at speed V would be:
t =
2r
V
,
39
40 CHAPTER 6. MANOEUVRE STABILITY
V
C
L1
, L = W
W = mg
C
L2
= (1 +n)C
L1
, L = (1 +n)W
W = m(1 +n)g
r
1: Steady, level ight 2: Steady pullout
Figure 6.1: Manoeuvre conditions
but by considering the centripetal acceleration we also know that:
r =
V
2
ng
.
Hence,
t =
2V
ng
.
The aircraft has pitched through a total angle of 2 radians in this time. Therefore the
pitch rate, q, is:
q =
2
2V/ng
=
ng
V
.
This pitching will cause the tail of the aircraft to move down relative to the incoming
air. This causes the incidence at the tailplane to increase by an amount:

T
=
ql
T
V
,
where l
T
is the tail arm measured from the centre of gravity. We already have an expres-
sion for q, so we get:

T
=
ngl
T
V
2
.
Unfortunately, this expression has a V
2
term, and hence will vary with the ight
conditions. We can get rid of this termby applying the denition of the lift coefcient,
6.1. ANALYSIS OF A STEADY PULLOUT 41
C
L
, for the straight and level case:
C
L
=
W
V
2
S/2
.
Hence
V
2
=
W
SC
L
/2
.
Substituting this back into the expression for
T
results in:

T
=
gSl
T
W
nC
L
2
,
or

T
=
nC
L
2
1
,
where
1
= W/gSl
T
, the longitudinal relative density.
The change in incidence of the tailplane causes the lift coefcient of the tailplane
to alter by an amount a
1

T
.Therefore, remembering that the lift coefcient of the
aircraft in the steady pullout is (1 +n)C
L
:
C
L
T
=
a
1
a
_
1
d
d
_
(1 +n)C
L
+a
1
(
T

0
) +a
2
+a
3
+a
1

T
.
Hence,
C
L
T
=
a
1
a
_
1
d
d
_
(1 +n)C
L
+a
1
(
T

0
) +a
2
+a
3
+a
1
nC
L
2
1
.
The basic pitching moment equation is still valid, since it makes no assumptions
about the source of the lift and momentsit is simply the result of non-dimensionalizing
a free body diagram. Therefore, this revised expression for C
L
T
can be substituted.
Again, remembering that the lift coefcient in the steady pullout is (1 +n)C
L
:
C
M
= C
M
0
(h
0
h)(1 +n)C
L
V
_
a
1
a
_
1
d
d
_
(1 +n)C
L
+a
1
(
T

0
) +a
2
+a
3
+a
1
nC
L
2
1
_
.
For the straight and level ight of the aircraft, in trim, we have previously derived
the equation:
C
M
= 0 = C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
2
+a
3

_
.
(6.1)
42 CHAPTER 6. MANOEUVRE STABILITY
In trim, the pitching moments acting on the manoeuvring aircraft will be zero if
the aircraft is undertaking a steady manoeuvre. If we now look at the expression for
trim in a steady pullout, and look at the change in elevator angle required for trim,
such that the elevator angle is now + we get:
0 = C
M
0
(h
0
h)(1 +n)C
L

V
_
a
1
a
_
1
d
d
_
(1 +n)C
L
+a
1
(
T

0
) +a
2
( +) +a
3
+a
1
nC
L
2
1
_
. (6.2)
These equations (6.1) and (6.2) are very similar. We can therefore subtract one
from the other and get:
0 = (h
0
h)nC
L
V
_
a
1
a
_
1
d
d
_
nC
L
+a
1
nC
L
2
1
+a
2

_
. (6.3)
This can be rearranged to get the elevator deection/g required for a steady pull-
out:

n
=
C
L
V a
2
_
(h
0
h) +V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
__
.
This must always be negative, otherwise the aircraft would pitch nose-down when
the pilot pulls back.
6.2 Stick xed manoeuvre point
When /n = 0 the c.g. is at the stick xed manoeuvre point. Hence, at h = h
m
:
0 =
C
L
V a
2
_
(h
0
h
m
) +V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
__
,
h
m
= h
0
+V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
_
.
This should be compared with the neutral point location stick xed, h
n
, which we
have previously shown to be:
h
n
= h
0
+V
a
1
a
_
1
d
d
_
.
Therefore, the stick xed manoeuvre point is a distance a
1
c/2
1
aft of the stick
xed neutral point. It is worth noting that the location of the stick xed manoeu-
vre point varies with altitude, since
1
is a function of the air density as well as of
geometry.
6.3. STICK FIXED MANOEUVRE STABILITY 43
6.3 Stick xed manoeuvre stability
The stick xed manoeuvre margin, H
m
, is dened as:
H
m
= h
m
h.
We showed in 6.1 that:

n
=
C
L
V a
2
_
(h
0
h) +V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
__
.
Hence,
h = h
0
+
V a
2
C
L

n
+V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
_
.
Therefore,
H
m
=
V a
2
C
L

n
.
This result is important because it demonstrates that there is a relationship between
the stick xed manoeuvre margin and the elevator angle to trim. There is, seemingly,
a discrepancy between the fact that h
m
moves with changing altitude and the above
expression. How can this discrepancy be explained?
6.4 Stick free manoeuvre stability
Stick xed analysis has enabled us to calculate the elevator angles required to trim
the aircraft in a steady pullout or bank, but tells us nothing about the stick forces
required for the manoeuvres (just as stick xed analysis told us nothing of the stick
forces for straight and level ight). Stick free manoeuvre stability analysis will allow
us to calculate these stick forces.
6.5 Analysis
In 5.1, we derived an expression that allowed us to calculate the hinge moments for
trim in straight and level ight:
C
M
= 0 = C
M
0
(h
0
h)C
L
V
_
a
1
a
_
1
d
d
_
C
L
+a
1
(
T

0
) +a
3
+
a
2
b
2
(C
H
b
0
)
_
.
The same process can be undertaken to nd the hinge moments for trim in a steady
pullout, C
H
+ C
H
. For the pullout, assuming that the tab is not used:
C
M
= 0 = C
M
0
(h
0
h)(1 +n)C
L

V
_
a
1
a
_
1
d
d
_
(1 +n)C
L
+a
1
(
T

0
) +a
3
+a
1
nC
L
2
1
+
a
2
b
2
(C
H
+C
H
b
0
)
_
,
44 CHAPTER 6. MANOEUVRE STABILITY
where C
L
is, again, the lift coefcient in straight and level ight.
Equating these two expressions and cancelling identical terms results in:
0 = (h
0
h)nC
L
V
_
a
1
a
_
1
d
d
_
nC
L
+a
1
nC
L
2
1
+
a
2
C
H
b
2
_
.
Hence,
V a
2
b
2
C
L
C
H
n
= (h
0
h) V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
_
.
The stick free manoeuvre point, h

m
, is dened as the c.g. position that gives C
H
/n = 0.
Therefore,
h

m
= h
0
+V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
_
.
The stick free manoeuvre margin, H

m
, is dened as:
H

m
= h

m
h,
which can easily be shown to be:
H

m
=
V a
2
b
2
C
L
C
H
n
.
The stick force per g is calculated from the hinge moment per g, in exactly the
same way as for straight and level ight:
P
e
n
= m
e
V
2
2
S

C
H
n
.
For handling safety the stick force required to pull high g should be appreciable to
avoid accidentally exceeding the structural limitations of the aircraft. A typical value
for a non-aerobatic aircraft is usually of the order of 20 N/g.
6.6 Tailless aircraft
The analysis for tailless aircraft is very similar to that for conventional aircraft. The
ight conditions, as for conventional aircraft, are shown in Figure 6.2.
For a tailless aircraft in steady trimmed ight we have already derived the equa-
tion (4.1):
C
M
= 0 = C
M
0
+
C
M
0

(h
0
h)C
L
.
6.6. TAILLESS AIRCRAFT 45
V
C
L1
, L = W
W = mg
V
C
L2
= (1 +n)C
L1
, L = (1 +n)W
W = m(1 +n)g
r
q = ng/V
1: Steady, level ight 2: Steady Pullout
Figure 6.2: Manoeuvre conditions for a tailless aircraft
When the aircraft is in a steady pullout with radial acceleration ng and with pitch rate
q we can write:
C
M
= 0 = C
M
0
+
C
M
0

( + ) (h
0
h)(1 +n)C
L
+
C
M
q
q.
Again, we can subtract one equation from the other to get:
=
1
C
M
0
/
_
(h
0
h)nC
L

C
M
q
q
_
,
where C
M
/q is an aerodynamic derivative. There are a large number of aerodynamic
derivatives that can be dened for any aircraft, and they enable us to calculate the
aerodynamic behaviour of the aircraft. We will encounter more aerodynamic deriva-
tives when we examine the dynamic stability of aircraft. There is a standard non-
dimensionalised form for each of these parameters. The non-dimensional form of
C
M
/q is given the symbol m
q
, and is dened as:
m
q
=
1
V Sc
2
0
M
q
.
Hence,
C
M
q
=
1
V
2
Sc
0
/2
M
q
=
2c
0
V
m
q
and we know that
q =
ng
V
.
Therefore,
=
1
C
M
0
/
_
(h
0
h)nC
L

2c
0
V
m
q
ng
V
_
.
46 CHAPTER 6. MANOEUVRE STABILITY
As for the conventional aircraft, we have an expression that includes the ight
velocity. Again, we can remove this by using the denition of the lift coefcient and
rearranging such that:
V
2
=
W
SC
L
/2
.
Making this substitution results in:
=
1
C
M
0
/
_
(h
0
h)nC
L
m
q
nC
L
gc
0
S
W
_
.
The longitudinal relative density,
1
, for a tailless aircraft is dened as:

1
=
W
gSc
0
.
Using this denition and rearranging results in:

n
=
1
C
M
0
/
_
(h
0
h)
m
q

1
_
C
L
.
This expression can be used to calculate the elevon deections required to undertake
manoeuvres.
6.7 Tailless aircraft manoeuvre point
As for the conventional aircraft, the manoeuvre point is dened by the c.g. location
that results in /n = 0. Therefore,
h
m
= h
0

m
q

1
.
The resulting aerodynamic forces due to a positive pitch rate are shown in Figure 6.3.
Forces oppose motion
Figure 6.3: Aerodynamic forces during pitching motion
These forces all oppose the motion of the aircraft. Hence, m
q
is always negative.
This means that the manoeuvre point for a tailless aircraft is always aft of the neutral
point for the aircraft (which is at h = h
0
). The damping in pitch has increased the
stability.
6.8. TAILLESS AIRCRAFT MANOEUVRE MARGINS 47
6.8 Tailless aircraft manoeuvre margins
The manoeuvre margin for a tailless aircraft, H
m
, is dened identically to that for a
conventional aircraft:
H
m
= h
m
h.
Hence
H
m
= (h
0
h)
m
q

1
= K
n

m
q

1
.
The elevon angle per g can therefore be written as:

n
=
H
m
C
L
C
M
0
/
.
The elevon angle per g is therefore directly proportional to the manoeuvre margin.
6.9 Relationships between static and manoeuvre
margins
Conventional aircraft
We have shown that the static margins, stick xed and stick free, for conventional
aircraft are:
K
n
= (h
0
h) +V
a
1
a
_
1
d
d
_
,
K

n
= (h
0
h) +V
a
1
a
_
1
d
d
_
.
Also, the manoeuvre margins for conventional aircraft are:
H
m
= (h
0
h) +V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
_
,
H

m
= (h
0
h) +V
_
a
1
a
_
1
d
d
_
+
a
1
2
1
_
.
Therefore,
H
m
= K
n
+
V a
1
2
1
,
H

m
= K

n
+
V a
1
2
1
.
The manoeuvre points of conventional aircraft are aft of the respective neutral points.
This is due to the stabilising inuence of additional lift at the tailplane due to the
pitch rate. Note that since
1
is a function of the air density the manoeuvre margin
decreases with increasing altitude.
48 CHAPTER 6. MANOEUVRE STABILITY
Tailless aircraft
We have shown that the static margin for tailless aircraft is:
K
n
= h
0
h
and that the manoeuvre margin is:
H
m
= (h
0
h)
m
q

1
.
Therefore,
H
m
= K
n

m
q

1
.
As for conventional aircraft, a tailless aircraft is more stable when manoeuvring
due to the stabilising effect of the pitch damping term m
q
(remember, m
q
is negative).
Again, the manoeuvre margin is reduced at high altitudes due to the presence of a
density term in
1
.
6.10 Modication of stick free neutral and manoeuvre
points
Two common ways of modifying the stick free neutral and manoeuvre points are
shown in Figure 6.4.
Bob weight
a: Spring b: Bob weight
Figure 6.4: Modication of neutral and manoeuvre points
A spring or bob-weight is dened as positive if it exerts a moment that would
cause a positive deection of the elevator.
These two additions have no effect on the stick-xed neutral or manoeuvre points
since the calculation of these locations does not require the consideration of hinge mo-
ments. However, it can be shown that a positive spring moves the stick free neutral
point aft but has no effect on the stick free manoeuvre point. In contrast, a positive
bob-weight moves both the stick free neutral point and the stick free manoeuvre point
of an aircraft aft. These effects are shown in Figure 6.5.
By combining positive and negative springs and bob-weights it is possible to
move the two stick free points independently of each other. Since the stick forces
6.10. MODIFICATION OF STICK FREE NEUTRAL AND MANOEUVRE POINTS49
N M
Spring
Bob-weight
N

Figure 6.5: Effects of positive springs and bob-weights


are directly proportional to the stick free static margin and the stick free manoeuvre
margin this enables the stick forces to be modied by a simple mechanical addition
to the system.
For example, an aircraft might have suitable stick free static stability but insuf-
cient manoeuvre margins. This results in a stable aircraft with good feel for the
pilot and suitable stick loads for trim, but the low stick force per g resulting from the
low manoeuvre margin might cause a risk of inadvertently overstressing the aircraft.
The addition of a negative spring together with a positive bob-weight would solve
this problem since the stick free static margin would be unchanged but the stick free
manoeuvre margin would increase. This is shown in Figure 6.6.
Positive bob-weight
Negative spring
N

Figure 6.6: Modication of stick force per g


Chapter 7
Compressibility effects
Everything that we have examined so far has assumed linear aerodynamics, incom-
pressible ow, and rigid aircraft. In reality, of course, none of these assumptions will
be valid at all ight conditions. At higher angles of attack the aerodynamics become
non-linear (i.e. C
L
/ not constant) and as the aircraft ies faster other effects be-
come important. In this section we will briey outline the major changes that occur
at high speeds, and the effect that this has on the control of the aircraft.
7.1 High speed effects
Changes from the low speed case arise primarily from Mach number (compressibil-
ity) and distortion (aeroelastic) effects. The dominant effects of increasing Mach num-
ber come from:
change of lift curve slope with Mach number;
movement of the aerodynamic centre rearwards, from quarter-chord at low
speed to mid-chord at supersonic Mach numbers.
These effects are shown in Figure 7.1. Combined with these are the effects of Mach
number on zero-lift pitching moment and zero-lift angle.
The effect of aeroelasticity is to reduce lift curve slopes with increasing V
2
/2,
dynamic pressure. The loads acting on an aircraft are proportional to the dynamic
pressure, if the lift and drag coefcients are constant. The deections are, similarly,
proportional to the forces. Hence, all aeroelastic effects depend on the dynamic pres-
sure. This results in changes in the aeroelastic response of the aircraft at different
altitudes, since the ambient air density varies with altitude. If we superimpose alti-
tude effects onto the variation of lift curve slope with Mach number for a typical (aft)
swept wing aircraft, we get a result as in Figure 7.3.
There are similar effects on the effectiveness of the tailplane and the elevator, as
shown in Figure 7.4. The downwash at the tail typically varies as shown in Figure 7.5,
decreasing to zero at high supersonic speeds, since any downwash is conned to the
volume of air inuenced by the wing.
51
52 CHAPTER 7. COMPRESSIBILITY EFFECTS
M
a
1.0 M
h
1.0
c/4
c/2
Figure 7.1: Compressibility effects on lift curve slope and aerodynamic centre
M
C
M
0
1.0 M

0
1.0
Figure 7.2: Compressibility effects on zero lift pitching moment and zero lift angle
The usual result of the combined effects is that stability reduces as the Mach num-
ber nears unity and then increases, sometimes rapidly, to a higher value at supersonic
speeds. The variation of C
M
for a typical aircraft is shown in Figure 7.6.
To counteract the nose down pitching moment that often occurs on swept wing
aircraft (subsonic jet transportsBoeing 707, 747, etc.) an up-elevator or stabilisation
input is provided by a Mach number sensing system. This is known as Mach trim. If
the nose down moment were allowed to take effect the stick force gradient would be
reversed, and there is also a danger that the maximum allowable speed of the aircraft
due to structural limits would be exceeded. The stick forces for such an aircraft are
shown in Figure 7.7.
7.1. HIGH SPEED EFFECTS 53
M
a
1.0
Rigid aircraft
Sea level
Decreasing altitude
Figure 7.3: Aeroelastic effects on lift curve slope
M

V a
1
1.0
Rigid aircraft
Sea level
Decreasing altitude
M

V a
2
1.0
Rigid aircraft
Sea level
Decreasing altitude
Figure 7.4: Aeroelastic effects on tailplane and elevator
54 CHAPTER 7. COMPRESSIBILITY EFFECTS
M
/
1.0
Figure 7.5: Variation of downwash with Mach number
M
C
M
1.0
Mach tuck
Figure 7.6: Variation of pitching moment with Mach number
M
Push
Pull
1.0
Uncorrected stick force
Mach trim input
Figure 7.7: Variation of stick forces with Mach number
Part II
Dynamic stability
55
Chapter 8
Dynamic behaviour of aircraft
In the rst part of the course, we examined the static stability of aircraft, which means
that we considered whether an aircraft tends to return to its equilibriumposition after
a perturbation, 1.1. We are nowgoing to analyze the dynamic stability of aircraft and
see how they respond over time to perturbations in ight.
8.1 Axes and notation
The axes and notation for the analysis of dynamic stability of an aircraft are given
in Figure 8.1 and follow a logical order. Once the x, y and z-axes are dened we then
have, for example, L, M and Nthe moments about the x-, y- and z-axes, respec-
tively.
The axis system uses body axes where the system is not locked in position in
space, but moves with the aircraft. The origin of the axis system is at the centre of
gravity of the aircraft, since all rotations take place about the c.g.
A rigid aircraft has six degrees of freedom. To simplify the equations used when
performing analysis of the dynamic modes of an aircraft, these degrees of freedomare
expressed as perturbation quantities in relation to steady straight ight (i.e. velocity
perturbations u, v and w and rotational velocities p =

, q =

and r =

).
8.2 Aerodynamic derivatives
We now have a co-ordinate system that allows us to dene any perturbation of the
aircraft from straight and level ight. To continue, we need to nd out what forces
are acting on the aircraft for a given perturbation
1
.
1
The analysis which follows is taken from MILNE-THOMSON, L. M., Theoretical aerodynamics,
MacMillan and Company, 1966.
57
58 CHAPTER 8. DYNAMIC BEHAVIOUR OF AIRCRAFT
z
x
y
Axis Perturbation Mean Perturbation Rotation Angular Moment Moment
force velocity velocity angle velocity of inertia
x X U u p A L
y Y V v q B M
z Z W w r C N
Figure 8.1: Notation for analysis of dynamic stability
If we assume that the effects of each perturbation are linear (true for small pertur-
bations), then:
M =
M
u
u +
M
v
v +
M
w
w +
M
p
p +
M
q
q +
M
r
r.
The partial derivatives in this expression are known as aerodynamic derivatives or
stability derivatives. We have already met the derivative M/q, often known as
pitch damping, in our analysis of control deections for tailless aircraft. Therefore,
if we know the aerodynamic derivatives and the perturbations, we can calculate all
of the forces and moments acting on the aircraft (six equations). If, in addition, we
know the mass of the aircraft and its inertia in roll, pitch and yaw (A, B, C) we can
calculate the acceleration of the aircraft, and hence its dynamic response.
The forces on the aircraft are the aerodynamic force F and mg:
F = Xi +Y j +Zk, (8.1)
mg = mg
1
i +mg
2
j +mg
3
k, (8.2)
8.2. AERODYNAMIC DERIVATIVES 59
where the components of g are needed because the reference frame is xed to the
aircraft and is not necessarily horizontal. The other quantities we need for the aircraft
are:
v = ui +vj +wk, velocity,
= pi +qj +rk, angular velocity,
h = h
1
i +h
2
j +h
3
k, angular momentum.
The equations of motion in translation and rotation are then:
d
dt
(mv) = m v + (mv) = F +mg, (8.3a)
dh
dt
=

h + h = L, (8.3b)
where the boxed terms are required because the frame of reference is rotating. The
applied moment L is:
L = Li +Mj +Nk.
Equations 8.3 are the general equations of motion for an aircraft and could, in princi-
ple, be used to calculate the motion given enough information about the aerodynam-
ics and mass distribution. We, however, want to know if the aircraft is dynamically
stable, so we need to make some approximations to see how the aircraft behaves
when perturbed from steady ight.
In steady ight, we write:
v = V, = 0, F +mg = 0,
and add the small perturbations so that:
V = V
1
+u,
V
1
= Ui,
u = ui +vj +wk,
= pi +qj +rk.
For a small rotation ,
= i +j +k,
=

i +

j +

k.
Similarly, the perturbation forces are:
F +F, m(g +g),
60 CHAPTER 8. DYNAMIC BEHAVIOUR OF AIRCRAFT
and it can be shown that
g + g = 0.
Inserting these assumptions into (8.3) yields the equations of motion for small pertur-
bations:
m u +m( V
1
+ g) = F, (8.4a)

h = L. (8.4b)
We can now simplify the system by making certain (reasonable) assumptions. First,
we assume that the forces and moments depend only on velocities and not on accel-
erations, with the exception of the dependence of pitching moment on w, the down-
wash acceleration. Then:
F = Xi +Y j +Zk,
L = Li +Mj +Nk,
and, for example,
X =
X
u
u +
X
v
v +
X
w
w +
X
p
p +
X
q
q +
X
r
r,
M =
M
u
u +
M
v
v +
M
w
w +
M
w
w +
M
p
p +
M
q
q +
M
r
r.
Secondly, we are assuming that the aircraft is symmetric so that a symmetric pertur-
bation can only cause a symmetric response. This means that a pitch disturbance, for
example, cannot cause a response in yaw or roll. Also, the symmetric response to an
asymmetric input has to be symmetric: if the aircraft rolls at a given rate, the pitch
response must be the same whether it rolls in a positive or negative sense. These two
requirements imply that:
Y
u
=
Y
w
=
Y
q
=
L
u
=
L
w
=
L
q
=
N
u
=
N
w
=
N
q
0,
X
p
=
X
q
=
X
r
=
Z
p
=
Z
q
=
Z
r
=
M
p
=
M
q
=
M
r
0.
Eliminating zero terms, we can write:
F = (
X
u
u +
X
w
w +
X
q

)i + (
Y
v
v +
Y
p

+
Y
r

)j
+ (
Z
u
u +
Z
w
w +
Z
q

)k,
L = (
L
p

+
L
r

+
L
v
v)i + (
M
q

+
M
u
u +
M
w
w +
M
w
w)j
+ (
N
p

+
N
r

+
N
v
v)k.
8.3. LONGITUDINAL SYMMETRIC MOTION 61
We need one more assumption about the aircraft, which is that there is no inertial
coupling between yaw and roll. This means that the only moments of inertia we
need consider are A, B and C, the moments of inertia about the coordinate axes.
Now, assuming disturbed horizontal ight and expanding the cross products in (8.4)
yields the equations of motion for each translational and rotational component:
m u =
X
u
u +
X
w
w +
X
q
q mg, (8.5a)
m( w Uq) =
Z
u
u +
Z
w
w +
Z
q
, (8.5b)
B q =
M
q
q +
M
u
u +
M
w
w +
M
w
w. (8.5c)
and
m( v +Ur) =
Y
v
v +
Y
p
p +
Y
r
r +mg, (8.6a)
A p =
L
p
p +
L
r
r +
L
v
v, (8.6b)
C r =
N
p
p +
N
r
r +
N
v
v. (8.6c)
The rst of these sets of equations covers symmetric motion, e.g. pitch oscillations,
while the second covers lateral motion, such as yaw and roll. An important point to
note is that these equations are uncoupled so that longitudinal motion does not affect
lateral and vice versa.
8.3 Longitudinal symmetric motion
The important information about the dynamic response of a system is the set of
modes in which it oscillates
2
. These can be found by the usual method of insert-
ing an assumed solution into the differential equations and nding combinations of
parameters which satisfy the system. The most convenient form of solution is:
u = u
0
e
t
, v = v
0
e
t
, =
0
e
t
.
Inserting these assumptions into (8.5a), for example, yields:
mu
0
e
t
=
X
u
u
0
e
t
+
X
w
w
0
e
t
+
X
q

0
e
t
mg
0
e
t
.
2
The following analysis, with different notation, is based on GRAHAM, W., Asymptotic analysis
of the classical aircraft stability equations, Aeronautical Journal, February 1999, pp 95103.
62 CHAPTER 8. DYNAMIC BEHAVIOUR OF AIRCRAFT
Now, we can divide through by exp t and, as always, non-dimensionalize the pa-
rameters, to give the non-dimensional equations of motion:
( x
u
)u

x
w
w

_
x
q

C
L
2
_

0
= 0, (8.7a)
z
u
u

+ ( z
w
)w

_
1 +
z
q

c
_

0
= 0, (8.7b)

_
m
w


c
+m
w
_
w

+
(b m
q
)

c
= 0. (8.7c)
The non-dimensional parameters are:
x
u
=
X
u
US
, x
w
=
X
w
US
, z
u
=
Z
u
US
, z
w
=
Z
w
US
,
x
q
=
X
q
USc
, z
q
=
Z
q
USc
, m
u
=
M
u
USc
, m
w
=
M
w
USc
,
m
q
=
M
q
USc
2
,
m
w
=
M
w
Sc
,
b =
B
mc
2
and
=
m
US
,
c
=
m
Sc
,
also given on the data sheet.
Chapter 9
Normal modes of aircraft
Phugoid
The rst approximate solution we consider is a low frequency oscillation. We state
without proof that there is a solution with and u

/ of order one and w

/
0
of order
1/
c
. This means that, in this case, the vertical motion is negligible or, equivalently,
the incidence is almost constant. We can rewrite (8.7) in matrix form, with the negli-
gible terms in each equation removed:
_
_
x
u
0 C
L
/2
z
u
0
0 m
w
(b m
q
)/
c
_
_
_
_
u

0
_
_
=
_
_
0
0
0
_
_
.
This equation can only have a non-trivial solution if the determinant of the matrix is
zero:

2
x
u

z
u
2
C
L
= 0.
Solving for gives:
=
(x
u
)
2
j
ph
_
1
_
x
u
2
ph
_
2
_
1/2
,
which species oscillatory motion with:

ph
=
_
z
u
C
L
2
_
1/2
, natural frequency, (9.1a)
c
ph
=
x
u
2
ph
, damping. (9.1b)
This solution denes the phugoid mode, which is a lightly damped long period os-
cillation. The incidence is almost constant and the aircraft varies altitude at constant
energy, trading potential for kinetic and back again, Figure 9.1.
63
64 CHAPTER 9. NORMAL MODES OF AIRCRAFT
h
max
, V
min
h
min
, V
max
Figure 9.1: Phugoid oscillation trajectory
An important point to note is that the damping is proportional to (z
u
), the rate
of change of vertical force with small changes in horizontal speed. Remembering
that the z axis points vertically down, we can see that z
u
< 0 and the damping is
positive. Although it is not proven on the basis of these results, a statically stable
aircraft always has a stable phugoid.
Short period oscillation
The second solution for longitudinal oscillation is for the case where is of order

1/2
c
, u
0
/
0
is of order
1/2
c
and w
0
/
0
is of order one. In this case, the approximation
to (8.7) is:
_

_
x
u
x
w

_
x
q

C
L
2
_
0 z
w

0
_
m
w

c
+m
w
_
(b m
q
)

c
_

_
_
_
u

0
_
_
=
_
_
0
0
0
_
_
.
Again, we nd the natural frequency by requiring that the determinant of the matrix
be zero:
( x
u
)
_

_
z
w
+
m
q
+m
w
b
_
+
z
w
m
q
m
w

c
b
_
= 0,
which, on solving the quadratic, gives a result for the non-dimensional natural fre-
quency and damping:

spo
=
_

c
(m
w
) +m
q
z
w
b
_
1/2
, natural frequency, (9.2a)
c
spo
=
1
2
spo
_
z
w
+
m
q
+m
w
b
_
, damping. (9.2b)
This is the short period oscillation and is a heavily damped mode with period typically
of a few seconds. The aircraft pitches rapidly about its centre of gravity which con-
tinues to y at almost constant speed in a straight line. The periodic time is typically
a few seconds, but must not be less than about 1.25s, otherwise there is a risk of Pilot
Induced Oscillation (PIO).
9.1. LATERAL MOTION 65
V
Figure 9.2: Short period oscillation
The frequency is proportional to K
1/2
n
, and increases with dynamic pressure, V
2
/2.
Therefore the aircraft will have the highest frequency SPO, and hence the shortest
time period, at high speed with the centre of gravity in the furthest forward position.
The SPO is always stable for a statically stable aircraft.
9.1 Lateral motion
In the case of lateral motion, we again need to insert the assumed form for the solu-
tion:
v = v
0
e
t
, =
0
e
t
, r = r
0
e
t
,
and non-dimensionalize quantities in (8.6), which we do in the same way as before
except that our reference length is now s, the wingspan:
( y
v
)v

_
y
p

s
+
C
L
2
_

0
+
_
1
y
r

s
_
r

= 0, (9.3a)
l
v
v

+ (a l
p
)

l
r

s
r

= 0, (9.3b)
n
v
v

n
p

0
+
c n
r

s
r

= 0. (9.3c)
The non-dimensional quantities are:
y
v
=
Y
v
US
,
y
p
=
Y
p
USs
, y
r
=
Y
r
USs
, l
v
=
L
v
USs
, n
v
=
N
v
USs
,
l
p
=
L
p
USs
2
, l
r
=
L
r
USs
2
, n
p
=
N
p
USs
2
, n
r
=
N
r
USs
2
,
a =
A
ms
2
, c =
C
ms
2
,
v

=
v
0
U
, r

=
mr
0
US
,
s
=
m
Ss
.
66 CHAPTER 9. NORMAL MODES OF AIRCRAFT
Dutch roll
The rst lateral mode we consider is Dutch roll which has oscillations of roughly equal
magnitude in pitch, yaw and roll. In this case, (9.3) reduce to:
_
_
0 1
l
v
a
2
/
s
0
n
v
0 c/
s
_
_
_
_
v

0
r

_
_
=
_
_
0
0
0
_
_
.
As before the determinant of the matrix must be zero for a non-trivial solution to
exist:

2
(c
2
+
s
n
v
) = 0,
and the frequency of the oscillation is, on the approximations we are using:

dr
=
_

s
n
v
c
_
1/2
. (9.4)
In Dutch roll, yawing oscillation (analogous to the longitudinal SPO) causes alternat-
ing sideslip. This in turn causes a rolling oscillation via L
v
v. The periodic time is
typically a few seconds, but as for the SPO it should not have a period of less than
1.25s due to PIO.
Dutch roll is not permitted to be divergent. Divergent Dutch roll can be xed by
a yaw damper on the rudder which damps the yawing oscillation, and hence the roll
response as well.
Spiral mode and roll subsidence
There are two further solutions to the dynamic equations which have small values of
. These are dominated by yaw and roll with weak sideslip and the corresponding
approximations to (9.3) are:
_
_
0 C
L
/2 1
l
v
(a l
p
)/
s
l
r
/
s
n
v
n
p
/
s
(c n
r
)/
s
_
_
_
_
v

0
r

_
_
=
_
_
0
0
0
_
_
.
The requirement for a non-trivial solution is then that:
an
v

2
+ [l
v
(n
p
cC
L
/2) l
p
n
v
] + (l
v
n
r
l
r
n
v
)C
L
/2 = 0.
The two roots of this equation can be approximated as:

rs
=
(l
p
)n
v
+ (l
v
)[cC
L
/2 + (n
p
)
an
v
, (9.5)
and

sm
=
C
L
2
l
v
n
r
l
r
n
v
(l
p
)n
v
+ (l
v
)[cC
L
/2 + (n
p
)]
. (9.6)
9.2. DIHEDRAL EFFECT AND WEATHERCOCK STABILITY 67
Note that both of these roots are real and so they do not describe oscillations. The
rst,
rs
, describes rolling subsidence which is a pure rolling motion that is generally
heavily damped, and is therefore usually stable. The damping is primarily from the
wings, where the incidence along the wing is changed due to the roll-rate, as shown
in Figure 9.3.

Roll rate p
Loading
Rolling moment L
p
p < 0
Figure 9.3: Rolling subsidence
This roll-rate results in a rolling moment L
p
p. Therefore, if L
p
is negative the
rolling subsidence mode is stable. This is generally the case. However, if L
p
becomes
negative, usually due to non-linearities in the lift curve slopes at high roll rates, auto-
rotational rolling can occur. This is what happens when an aircraft spins
1
The second root
sm
, which is much smaller than
rs
, corresponds to the spiral
mode of the aircraft. This is a combined yaw and roll motion which is allowed to be
unstable (i.e. negatively damped) as long as it does not double amplitude in less than
twenty seconds, so that it can be controlled out.
The dynamics of the spiral mode are that if the aircraft rolls slightly, it will start
to sideslip, and the n then tries to turn the aircraft into the relative wind due to a
yawing moment N
v
v. However, the rolling moment due to sideslip L
v
v tries to roll
the wings back level. Depending on which of the effects wins, the aircraft will be
spirally unstable or stable, as can be seen from the numerator of (9.6).
9.2 Dihedral effect and weathercock stability
The aerodynamic derivatives L
v
and N
v
dene whether an aircraft is stable or un-
stable in rolling subsidence and Dutch roll. L
v
and N
v
are known as the dihedral
effect and weathercock stability respectively. The effect of the two aerodynamic
derivatives on the lateral stability of the aircraft is shown in Figure 9.4.
Dihedral effect
L
v
is known as the dihedral effect since the majority of the rolling moment due to
sideslip comes from dihedral (on an aircraft with unswept wings), as shown in Fig-
1
Something similar can happen to microlights, in the so-called tumble, which is almost always fa-
tal: GRATTON, G, & NEWMAN, S., The tumble departure mode in weightshift-controlled microlight
aircraft, Proceedings of the Institution of Mechanical Engineers, Part G: Journal of Aerospace Engineering,
March 2003, 217(3), pp. 149166.
68 CHAPTER 9. NORMAL MODES OF AIRCRAFT
L
v
N
v
Increasing altitude
Unstable spiral mode
Unstable Dutch roll
All lateral modes stable
Figure 9.4: Stability of the lateral modes
ure 9.5. Positive dihedral combined with positive sideslip results in a negative rolling
moment (and hence negative L
v
).

Relative wind
Positive sideslip
L
v
< 0
Figure 9.5: Dihedral effect
Wing sweep has a large, negative, effect on L
v
due to reduced or increased effec-
tive sweep for positive sideslip. This is shown in Figure 9.6.
Wingfuselage interference effects give contributions to L
v
due to changes in wing
effective incidence near the root. These contributions are negative for high mounted
wings and positive for low mounted wings, as shown in Figure 9.7.
A reasonable level of L
v
may be achieved by using anhedral with swept and high
mounted wings (e.g. Harrier). Ground clearance issues may limit anhedral on low
wing aircraft, resulting in an unstable Dutch roll mode.
9.2. DIHEDRAL EFFECT AND WEATHERCOCK STABILITY 69
Reduced eective sweep:
increased lift
Increased eective sweep:
reduced lift
Positive sideslip
L
v
< 0
Figure 9.6: Wing sweep effects on L
v
L
v
< 0
L
v
> 0
High wing Low wing
Figure 9.7: Wing-fuselage interference effects on L
v
Weathercock stability
The aerodynamic derivative N
v
is known as weathercock stability since it is, effec-
tively, the ability of an aircraft to turn into the wind. It is produced mainly by the
sideways lift-force of the n in sideslip, and should always be negative. However, as
shown in 9.2, if N
v
is too large the aircraft may be spirally unstable.
The n contribution to N
v
generally reduces with increasing Mach number, since
the ns lift curve slope is reducing. Therefore an aircraft with a large n may be
spirally stable at high speeds but unstable at low speeds. This can be solved by using
paired ns close together. At low speeds their mutual interference reduces their
effectiveness, while at supersonic speeds this interference is progressively removed,
increasing their effectiveness to combat the decreasing lift curve slope. This is shown
in Figure 9.8.
For VSTOL aircraft (e.g. Harrier) engine air intake mass ow may give a negative
contribution to N
v
making the aircraft directionally unstable in the hover and at low
forward speeds. This is because the air undergoes a change in direction to go down
the intake, and hence a momentum change, giving a sideforce acting ahead of the
70 CHAPTER 9. NORMAL MODES OF AIRCRAFT
M
n a
1
1.0
Mach cone
Figure 9.8: Use of twin ns at high speed
aircraft c.g., as shown in Figure 9.9.
Intake ow
S
id
e
fo
r
c
e
N
v
< 0
Flight speed
N
v
Fin
In
t
a
k
e

o
w
Total
Directionally
unstable
Figure 9.9: Intake effects on N
v
Part III
Spacecraft dynamics and control
71
Chapter 10
Getting around: Orbits
Spacecraft are governed by different, simpler, equations than aircraft, because they
operate without friction and, most of the time, have no applied thrust. The control of
spacecraft is a problem in xing their orbital parameters, so we start by analyzing the
classic problem of two bodies orbiting each other.
10.1 The two-body problem
Figure 10.1 shows the two body problem: two masses m
1
and m
2
interact with a
gravitational force of magnitude Gm
1
m
2
/|r
1
r
2
|
2
. We note, in passing, that one
mass will usually be very much greater than the other (when the Earth orbits the sun
or a satellite orbits the Earth, for example), though the analysis does not depend on
this assumption.
r
1
r
2
r
2
r
1
m
1
m
2
Figure 10.1: The two-body problem
73
74 CHAPTER 10. GETTING AROUND: ORBITS
The equations of motion for the two masses, in an inertial frame, are:
m
1
r
1
=
Gm
1
m
2
|r
1
r
2
|
3
(r
1
r
2
), (10.1a)
m
2
r
2
=
Gm
1
m
2
|r
1
r
2
|
3
(r
2
r
1
). (10.1b)
We can extract one useful piece of information immediately by considering the
centre of mass of the system r
c
:
r
c
=
m
1
r
1
+m
2
r
2
m
1
+m
2
.
Adding equations 10.1 shows that r
c
0. In other words, the centre of mass of the
system moves at constant velocity.
To nd the relative motion of the masses, subtract m
2
times( 10.1a) fromm
1
times (10.1b)
and dene r = r
2
r
1
:
m
1
m
2
r =
Gm
1
m
2
(m
1
+m
2
)
r
3
r,
r =
r
r
3
, (10.2)
where the gravitational parameter = G(m
1
+ m
2
). For most purposes, m
1
m
2
and
Gm
1
. For Earth, = 3.98601 10
5
km
3
/s
2
.
x
y

r
Figure 10.2: Polar coordinate system
To solve for r, we rewrite the system using polar coordinates, Figure 10.2. There
are two unit vectors r and

, the radial and azimuthal vectors respectively:
r = [cos sin ], (10.3a)

= [sin cos ], (10.3b)


and differentiation will easily show that:

r =

,

r =

2
r, (10.4a)

r,

. (10.4b)
10.1. THE TWO-BODY PROBLEM 75
Since r = rr:
r = ( r r

2
)r + (2 r

+r

r
2
r,
and, extracting components,
r r

2
=

r
2
, (10.5a)
2 r

+r

= 0. (10.5b)
Equation 10.5b can be rearranged to show that:
d
dt
_
r
2

_
0.
This tells us that h = r
2

, the angular momentum, is constant. We can also derive an
energy conservation equation by taking the dot product of (10.2) with r:
dE
dt
0, (10.6)
E =
r. r
2


r
, (10.7)
the energy per unit mass.
What we really want to knowis the shape of the orbit, i.e. r() and the easiest way
to do this is to convert the derivatives with respect to t to derivatives with respect to
. We also make the transformation r 1/u. Then:
r =
1
u
,
r =
1
u
2
du
d

,
r =
1
u
2
d
2
u
d
2

2
.
Inserting these terms into (10.5a) and noting that h = r
2

, the differential eqution of
motion is:
d
2
u
d
2
+u =

h
2
, (10.8)
with solution:
u = Acos( ) +

h
2
. (10.9)
It can be shown, using (10.7), that
A =

h
2
_
1 + 2
Eh
2

2
_
1/2
,
76 CHAPTER 10. GETTING AROUND: ORBITS
so that the orbit is given by:
r =
h
2
/
1 +e cos( )
. (10.10)
The minimum value of r occurs when = . If we use this point as the origin of
, the equation for the orbit is:
r =
p
1 +e cos
, (10.11)
where p = h
2
/ and the origin is called the perigee (for Earth orbits). Equation 10.11
has the form of a conic section. In particular, for most spacecraft it is elliptical (or
circular).
Elliptical orbits
Equation 10.11 describes the trajectory of an orbiting body as a conic section. The
geometry of an elliptical orbit is shown in Figure 10.3 which shows the notation and
geometrical parameters for an elliptical orbit. The ellipse has two focii and the Earth
(say) is at one focus, which is the centre of the coordinate system. The size and shape
of the ellipse are dened by the semi-major and semi-minor axes a and b, respectively.
The eccentricity of the orbit is e and tells us how distorted the ellipse is:
b = a(1 e
2
)
1/2
,
so that when e = 0, the orbit is circular. Elliptical orbits have 0 < e < 1, parabolic
orbits have e = 1 and hyperbolic orbits have e > 1.
a
b
a
Focus Focus
r

p
Perigee Apogee
Figure 10.3: The geometry of an ellipse
10.2. ORBITAL MANEOUVRES 77
From Figure 10.3, we can also nd the radius to perigee a(1 e) and the radius to
apogee a(1+e). It can be shown, using the geometric properties of the elliptical orbit,
that the energy E is:
E =

2a
, (10.12)
and the orbital period T is:
T =
2

1/2
a
3/2
(Keplers third law), (10.13)
which tells us that once an orbital radius (or semi-major axis) is selected the period of
the orbit is xed.
Example: space debris
For a spacecraft orbiting the earth at an altitude of 200km, estimate the highest veloc-
ity at which it might be hit by space debris. Assume that both the satellite and the
debris are in a circular orbit.
Orbital radius a = (6400 + 200)km,
Orbital period T =
2

1/2
a
3/2
,
= 5336s.
Orbital velocity v = 2a/T,
= 7.7km/s.
In the worst case, a spacecraft might, in principle, be hit by debris travelling at 2v =
15.4km/s.
10.2 Orbital maneouvres
The rst thing we have to note that makes spacecraft different from aircraft comes
from (10.7). Rearranging that equation, the spacecraft velocity is:
v =
_
2E +
2
r
_
1/2
, (10.14)
so that the spacecraft velocity is xed by E, and r. Now is (more or less) constant
but we can vary E, r and v, though not independently(10.14) is always true. If we
change the spacecraft velocity at some radius r, then E will change and the spacecraft
will enter a different orbit. This fact can be exploited in spacecraft maneouvres.
78 CHAPTER 10. GETTING AROUND: ORBITS
Transfer
LEO
GEO
a
1
a
2
Figure 10.4: Walter Hohmann and his transfer from low earth to geostationary Earth
orbit.
Hohmann transfer
The Hohmann transfer is the minimum energy transfer between two circular orbits.
If we have a satelliteof any type, this also works for lunar or Mars missionsin a
low earth orbit of radius a
1
, we can shift it to a higher orbit of radius a
2
using the
scheme shown in Figure 10.4. The transfer orbit is elliptical with semi-major axis
a = (a
1
+a
2
)/2 and is tangential to the two circular orbits.
In order to change a spacecrafts orbit, we have to change its velocity v. To enter
the transfer orbit, we have to change from the circular orbit velocity v = (/a
1
)
1/2
to
the elliptical orbit velocity at that position:
v
2
=
_
2
a
1

1
a
_
.
The change in velocity v
1
is then:
v
1
=
_
2
a
1

2
a
1
+a
2
_
1/2

a
1
_
1/2
. (10.15)
To switch to the circular orbit at a
2
, the velocity change is
v
2
=
_

a
2
_
1/2

_
2
a
2

2
a
1
+a
2
_
1/2
. (10.16)
The fundamental measure of performance for a spacecraft is its v (Delta vee), its
capacity to change velocity, as this limits its range of maneouvre. Absolutely every-
thing that a spacecraft uses, including propellant for maneouvring, has to be lifted
from the Earths surface, so it is vital to use economical transfer orbits. The only
problem with a Hohmann transfer is that it is slowit minimizes the total v but
10.2. ORBITAL MANEOUVRES 79
maximizes the time. For transfer to GEO, this does not matter (5.3 hours) but can be
a problem for other missions (Mars, for example). An especially important v is the
velocity change required to escape the gravity of a planet. This is called the escape
velocity and is found by setting the total energy E to zero. From (10.7):
v
esc
=
_
2
r
_
1/2
, (10.17)
which for a body on the surface of the Earth is about 11km/s.
Orbital capture
Without going into the details of how it might be done, it is obviously possible to
have a spacecraft leave the orbit of one planet and approach the orbit of another. If all
we want is to y past the planet, no more need be said. Indeed, a common method
of speeding up spacecraft is to have them approach another planet and pick up grav-
itational energy to accelerate them in another direction. If we want the spacecraft to
enter the orbit of the planet, however, we need to slow it down. Remember that the
spacecraft is moving very quickly because it has reached the escape velocity for the
planet it has left so orbital capture is not an easy job.
The rst point to note, as a spacecraft approaches another planet, is that it is in
orbit about the sun. It is only as it comes close to the planet that it feels the effect of
the planets gravity. At some point, as for a Hohmann transfer, we need to change the
spacecraft velocity to match the velocity of a planetary orbit. These velocity changes
can be quite large (for Mars Global Surveyor, for example, v = 973m/s) so it is
obviously important to nd optimal methods which save propellant.
If we want to enter a circular orbit about the planet, the orbital velocity is v
1
=
(
p
/r
p
)
1/2
where
p
is the gravitational parameter for the planet and r
p
is the orbital
radius. If the spacecraft approaches the planet with speed v

, its speed in the planet


frame is:
v
1
=
_
v
2

+
2
p
r
p
_
1/2
, (10.18)
by conservation of energy. The velocity change required is then:
v =
_
v
2

+
2
p
r
p
_
1/2

p
r
p
_
1/2
. (10.19)
Now we have to think about which orbit to choose. Propellant costs mass costs
money so we would like to nd the radius which minimizes v. Making trajectory
adjustments during the trip is relatively cheap, because very small adjustments can
give large changes in the entry point, so we are free to pick the best r
p
. Differentiat-
80 CHAPTER 10. GETTING AROUND: ORBITS
ing (10.19) with respect to r
p
, we nd
r
p
=
2
p
v
2

, (10.20)
v
min
=
v

2
1/2
. (10.21)
This is not the absolute optimum because a highly elliptical orbit will have a smaller
v again. One method for entering a circular orbit, as used by Mars Global Surveyor,
is to pick a highly elliptical orbit (lowv) for the planetary capture and use aerobrak-
ing (no propellant used, though a bit slow) to lower the orbit.
Chapter 11
Getting things done: Spacecraft control
Spacecraft are hard things to control because, unlike aircraft, they have no drag act-
ing on them: they will go on doing whatever they were doing until a control input
is applied, subject only to gravitational effects. This makes life simple in one way
since basic Newtonian mechanics applies, without dissipation, but it does make the
designers life quite difcult, since there is no damping. The two main functions of a
control system on a spacecraft are attitude control and manouevre or orbit change.
11.1 Attitude control
A satellite is usually required to point stably in a certain direction. Typical examples
include:
earth observation can only be performed if a satellite can point a camera at a known
position on the earths surface;
communications satellites need to beam information at a particular region (Corona-
tion Street is not very interesting to penguins);
scientic satellites such as Hubble need to point in a direction of interest to as-
tronomers.
Additionally, many satellites of various classes need to align their solar panels in the
right direction to generate power. Depending on the reason for the alignment, and the
accuracy required, there are a number of methods available for aligning spacecraft.
The brute force technique of ring a thruster to shift the spacecraft back on to the
desired attitude is wasteful of fuel and makes the system bounce back and forth. In
practice, spacecraft are designed to be inherently stable, which is achieved by making
them spin.
A mechanics textbook will give the Euler equations for a spinning body. If the
body is axisymmetric (i.e. the moment of inertia is the same about two of its principal
81
82 CHAPTER 11. GETTING THINGS DONE: SPACECRAFT CONTROL
axes):
M
1
= A
d
1
dt
+ (C A)
2

3
, (11.1a)
M
2
= A
d
2
dt
+ (A C)
3

1
, (11.1b)
M
3
= C
d
3
dt
+ (A A)
1

2
, (11.1c)
where M
i
and
i
are the moment and angular velocity about the ith axis and A and
C are the moments of inertia. For a body in space, there are no applied moments so
M
i
0 and this means that d
3
/dt = 0 or
3
= and the body rotates at constant
frequency about its axis. Now, we can solve for the other two angular velocities:
d
1
dt
= +
A C
A

2
,
d
2
dt
=
A C
A

1
.
Differentiating the rst and substituting the second equation:
d
1
dt
2
+
2

1
= 0,
where
2
= (C A)/A. This is a simple harmonic oscillator and the solution is:

1
=
0
sin t, (11.2a)

2
=
0
cos t. (11.2b)
This solution says that the spinning body rotates about one axis at a frequency
and oscillates about the other two. The axis of rotation actually swings around and
describes a cone in space. If the spacecraft has been designed and set up properly, the
spin stabilizes its attitude and the coning motion is small.
Gravity gradient
A simple way to align a satellite is to use a gravity gradient device: if a spacecraft
is asymmetric, the variation in gravity over its extent is enough to generate a torque
which draws it back into the required alignment. This is like having a pendulum
which wants to swing into a vertical orientation.
Sun-synchronous orbits
By exploiting the fact that the earth is not perfectly spherical, a satellite can be placed
into an orbit which is synchronized with the sun. This means that the satellite will
spend half its orbit in sunlight and will never have to work in a twilight region.
For an observation satellite, this also means it will pass a given point on the earth at
11.2. MANOUEVRING 83
V-bar
V-bar approach
R-bar
R-bar approach
Figure 11.1: Two approaches for a docking spacecraft
the same local time every day, which makes it easier to interpret images because the
ground will always be illuminated from the same direction. Finally, this means that
the satellite can guarantee that its solar panels will be generating power: an important
consideration for some systems.
11.2 Manouevring
Manouevring is a special case of changing orbits, but generally on a smaller scale.
The same rules apply, but our intuition is even more confused than usual, because
things dont look right.
An example is the docking of one spacecraft with another (supply ights to a
space station; Shuttle to Hubble; Apollo lander with orbiter). If one spacecraft is
already on orbit, the other must approach that orbit to rendezvous with it. There are
two standard approaches.
The rst is to arrive along the orbit in front of the other spacecraft. Because this
means ying along the velocity vector of the main craft, it is called a v-bar approach
(as in v for velocity). The second approach is along the radius vector to the main craft,
called an r-bar approach. The v-bar approach is easier because ying along the
radius vector makes the docking craft move ahead of the main craft, if it approaches
from below, as the orbital velocity at the lower altitude is a bit higher. Likewise,
approaching from above, you tend to fall behind the vehicle you are rendezvousing
with.
While a v-bar approach is favoured for simplicity, an r-bar manouevre is often
favoured because it means that the thrusters used for braking never point at the main
craft. In order to dock on a v-bar trajectory, you need to re thrusters at the main craft,
84 CHAPTER 11. GETTING THINGS DONE: SPACECRAFT CONTROL
which can damage delicate components such as solar panels or instruments (e.g. on
Hubble). The manouevre is carried out by adjusting the docking vehicles speed to
change its orbit slightly and bring it onto the main orbit from above or below.
Part IV
Problems
85
Chapter 12
Problems
W=100kN
L
W
x =
0.3m
L
T
l =15m
M
0
=40kNm
W=100kN
L
W
x =
0.3m
L
T
l =15m
M
0
=40kNm
Figure 12.1: Aircraft with different centres of gravity
1. For the two situations shown in Figure 12.1, calculate the values of L
W
and L
T
required to give both a total lift equal to the aircraft weight and give zero net
moment about the aircraft c.g.
[L
W
= 99.3kN, L
T
= 0.7kN, L
W
= 95.3kN, L
T
= 4.7kN]
2. Draw the system of forces and moments acting on a conventional aeroplane in
steady straight and level ight.
Show that the pitching moment about the centre of gravity is given by
C
M
= C
M
0
(h
0
h)C
L
V C
L
T
.
For the sailplane whose details are given in Table 12.1, calculate the value of C
L
T
required for trim at 50kt EAS with a pilot weighing 0.75kN. The empty weight
equipped is 2.5kN, with c.g. on the mean chord 0.45c aft of the leading edge of
c. The pilot c.g. is assumed to be 0.8m ahead of the leading edge of c.
[C
L
T
= 0.552]
3. Distinguish between stability and trim. Show that for an aircraft to be both
stable and able to trim at positive lift coefcient the overall pitching moment
about the centre of gravity must be positive at C
L
= 0 in that conguration.
87
88 CHAPTER 12. PROBLEMS
S = 28m
2
S
T
= 1.4m
2
c = 1.15m
l = 5.35m h
0
= 0.25 C
M
0
= 0.11
Table 12.1: Sailplane data
4. From rst principles, show that the portion of the total lift coefcient (C
L
) pro-
vided by the wing, body and nacelles (WBN) group of a conventional aircraft is
given by:
C
L
WBN
= C
L
_
1 + (h
0
h)
c
l
_
C
M
0
c
l
.
If the aircraft stalls when C
L
WBN
reaches its maximum value, (C
L
WBN
)
max
say,
then obtain an expression relating the stalling speed to the c.g. position at any
one given weight.
Hence calculate the c.g. shift that would increase the stalling speed by 1% if
c = 5.6m, l = 15.5m and (h
0
h) = 0.05.
[h = 0.0566, hc = 0.317m]
5. Consider the two situations shown in the diagrams below.
h c
W
h
n
c
C
L
(C
M
0
)
h
p
c
h
n
c
C
L
(C
M
0
)
(C
M
p
)
Figure 12.2: Full-scale and model aircraft. C
M
p
is measured by the balance, which
restrains the model in pitch.
In (a) the full scale aircraft is in steady free ight with values C
L
, h, for the lift
coefcient, c.g. position and elevator angle respectively.
In (b), the model of the same aircraft is suspended from a wind tunnel balance
at the same C
L
and elevator setting as in (a). The balance measurement gives the
pitching moment coefcient C
MP
about the balance pivot axis which is located
at h
p
with respect to the same datum line as h.
a) Write down the moment equations for situations (a) and (b), and hence
derive the relationship between the balance reading C
MP
, equivalent to
the steady free ight conditions, and interrelating h
p
, h and C
L
.
b) An aircraft model is found to have a zero-lift pitching moment coefcient
of 0.027 for a particular elevator angle. The pitching moment is measured
89
about the wind-tunnel axis of rotation P and has a slope:
dC
Mp
d
= 0.15; lift curve slope a = 5.851.
Determine the position of the c.g. of the full-scale aircraft relative to P if a
stick-xed margin of 0.11 is required (c = 3.96m).
If the wing loading is 2.25kN/m
2
in steady level ight with the above ele-
vator angle, what is the airspeed if the air density is 1.030kg/m
3
.
[0.537m forward of P, 133.3m/s TAS.]
90 CHAPTER 12. PROBLEMS
1. The data shown below apply to an aircraft in steady level ight at 200kt EAS.
Calculate the elevator angle required for longitudinal trim. Also obtain the
stick-xed neutral point and static margin.
W = 30kN S = 23m
2
S
T
= 3.5m
2
c = 1.96m l = 5.5m
h
0
= 0.25 c.g. is 0.61m aft of datum
C
M
0
= -0.036
T
= -1.5

= 0.48
a = 4.58 a
1
= 3.15 a
2
= 1.55
[ = -1.658

, h
n
=0.4027, K
n
= 0.0915]
2. The centre of gravity range for an aircraft is found by considering that the
a) aft c.g. limit (h
aft
) is determined by the minimum stability condition (min-
imum K
n
);
b) forward c.g. limit (h
fwd
) is determined by the maximum elevator angle to
trim (while retaining enough elevator movement for manoeuvre).
By considering the static forces and moments on an aircraft in symmetric ight,
nd an expression for the static margin stick-xed, K
n
, and show that:
K
n
= V a
2
d
dC
L
= (h
0
h) +V
a
1
a
_
1

_
.
An aircraft has the following values of the aerodynamic coefcients:
h
0
=0.25, a=3.5, a
1
=3.0, a
2
=1.5, /=0.4.
Find the relationship between the c.g. position and the tail volume ratio:
a) for a static margin of 0.05 (h
aft
);
b) for the change in elevator angle to trimto 10

for a change of C
L
of 1.0 (h
fwd
).
Hence nd the minimum tail volume ratio such that with a c.g. shift of 0.15c the
change in elevator angle to trim is not more than 10

for a change of C
L
of 1.0
and the static margin is never less than 0.05.
[V = 0.764]
3. Atransport aircraft with conventional tail is to have zero elevator angle in cruis-
ing ight at 560km/h EAS (mass 100,000kg), with the c.g. in the mid position.
The landing approach, out of ground effect, is made with aps down at 210km/h
(mass 90,000kg), and the maximum elevator movement permitted for trimming
91
is = 10

. Using the data below, calculate the minimum tailplane area suit-
able for this aircraft, and the tailplane setting
T
relative to the aps-up wing
zero lift line.
Minimum K
n
= 0.05 c.g. range h=0.50 h
0
=0.075 S = 232m
2
c = 4.72m l=19.5m
a=5.7 a
1
= 2.7 a
2
= 2.1
C
M
0
= -0.14 =0.16.
The change in C
M
0
at landing ap setting C
M
0
= 0.10. Note that the wing
zero lift incidence angle changes by 10

when the aps are lowered to the land-


ing setting.
[S
T
= 68.5m
2
,
T
= 3.92

]
92 CHAPTER 12. PROBLEMS
1. The static margin, stick-xed may be obtained in practice from ight tests in
which the elevator angles to trim are found at certain speeds. In practice, the
aeroplane is trimmed at a series of speeds by adjusting the tab setting, and both
the elevator angle and tab angle are observed. Since the theory which relates
the stick-xed static margin to the elevator angles to trim implicitly assumes a
constant tab angle, show that a correction must be applied to elevator angles
obtained in this way such that

corrected
= +
a
3
a
2

where and are the observed elevator and tab angles to trim at a given speed.
Suggest how you would determine a
3
/a
2
in ight.
2. Atailless aircraft is controlled in pitch by six elevons. Each elevon is actuated by
an independent power control unit. These units are so designed that if a failure
occurs the affected elevon is able to move until its hinge moment is zero.
Assuming the failure of one such unit, calculate the elevon angles that will give
longitudinal trim of the aircraft whose details are given below:
Weight = 850kN Speed = 70m/s EAS
Wing area S = 358m
2
(h
0
h) = 0.15
C
M
0
= +0.02 C
M
0
/ = -0.45
a
1
= 4.0 a
2
= 0.95
b
1
= -0.7 b
2
= -1.05
Assume that each elevon contributes equally to a
2
and to C
M
0
/.
[
failed
= 9.61

,
operating
= 13.15

]
3. A conventional aircraft ying at low speed has a exible rear fuselage such that
the tailplane setting relative to the wing zero lift line is directly proportional to
the tail load. Prove that the reduction in stick xed static margin compared with
that of the rigid aircraft is given by:
K
n
= K
rigid
n
K
exible
n
,
= V
a
1
a
_
1

__
1
1
1 +
1
2
V
2
S
T
a
1
f
_
For the human-powered aircraft having the characteristics given below, nd
the fuselage exibility f (degrees deection per Newton) that reduces the stick-
xed static margin by 0.05 compared to the rigid case when ying at a speed of
9.2m/s at sea level.
S = 28m
2
S
T
= 1.4m
2
l = 5.34m c = 1.14m
a = 6.0 a
1
= 4.5 = 0.20.
[f = 0.1

/N]
93
4. The control column of a low-speed aeroplane is connected to the elevator by
an arrangement of cables which stretches when a stick force is applied. The
stiffness of the circuit is given by dH
E
/d = ENm/rad where H
E
is the hinge
moment and is the elevator deection, the stick being held xed.
Show that the stick-xed c.g. margin (as opposed to the elevator xed c.g.
margin) is given by:
K
n
= (h
0
h) +V
a
1
a
_
1

__
1
a
2
b
1
a
1
b
2
1
1
_
where
=
C
L
SE
b
2
S

W
.
It should be assumed that the aircraft is initially in trim with the tab angle ad-
justed to give zero stick force.
Show how this angle is related to the stick xed and stick free c.g. margins of a
rigid aeroplane. What practical use might you make of this information?
94 CHAPTER 12. PROBLEMS
1. What conditions dene the stick-xed and stick-free manoeuvre points of an
aircraft?
From rst principles, stating your assumptions, derive an expression for the
stick xed maneouvre point of a low speed aircraft of canard layout. Show
whether this is forward or aft of the corresponding neutral point and compare
your expression with that for a conventional aircraft.
2. Dene the maneouvre point stick-free for a conventional aircraft. How does it
differ from the corresponding neutral point?
Find the minimum stick force per g at sea level for the light aircraft whose
details are given below. Comment on your result and nd the c.g. position
required to give 22N/g. Suggest alternative means for increasing the existing
value.
W = 2.7kN S = 7.6m
2
l = 2.9m
h
0
= 0.238 c = 1.2m V = 0.34
= 0.385 a = 3.865 a
1
= 2.73
a
2
= 2.16 b
1
= 0.282 b
2
= 0.536
The permitted c.g. range is from 0.22c to 0.28c. The stick force per g is given by
Q =
P
e
n
= m
e
S

W
S
b
2
a
2
V
H

m
,
= 83.2H

m
N/g for this aircraft.
[Q = 5.8N/g, for Q = 22N/g, h = 0.0853]
3. The table below shows data for a tailless aircraft. When it performs a steady
pullout at A
N
= 2.5 (n = 1.5) at 250kt EAS at a height where the air density
= 1.150kg/m
3
, the change of elevator setting compared with steady level ight
under the same conditions is 3.20

.
Calculate m
q
if the static margin is known to be 0.05.
W = 160kN S = 50m
2
c
0
= 10m
C
M
0
/ = 0.5 K
n
= 0.05
[m
q
= 0.264]
4. An aircraft of conventional layout is controlled in pitch by an all-moving tailplane,
having no separate elevators (see Figure and table). Show that the tail angle per
g is given by

T
n
=
C
L
H
m
V a
1
95
where the symbols have their usual meanings.
Hence calculate the tail angle, tail load and pivot moment when the aircraft is
ying at 440kt EAS in an 8g pullout at a height where the relative density of the
air = 0.74. Comment on your results.
W=175kN V =440kt EAS S=33.2m
2
S
T
=19.1m
2
l=5.25m c = 2.39m a=3.8 a
1
=2.7
C
M
0
= +0.03 /=0.38 h
0
= 0.17 h = 0.50
= 0.74
L
WBN L
T
l
Pivot point
0.144m
[
T
= 4.72

,
T
= 4.85

, L
T
= 224.6kN, M
P
= 32.34kNm, C
L
T
= 0.3739]
96 CHAPTER 12. PROBLEMS
1. For a conventional aircraft show that if the tab setting remains unaltered, the
change of elevator hinge moment coefcient-to-trim

C
H
between two lift co-
efcients is given by

C
H
=
b
2
a
2
V
C
L
K

n
.
The aircraft described in the table below is making a zero stick force trimmed
landing approach at 155kt EAS. Calculate the value to which the speed may be
reduced while keeping the stick force within 150N without altering the trim tab
setting, indicating clearly whether this is push or a pull force.
Weight W =785kN c.g. at h = 0.26
Wing area S = 223m
2
smc c = 5.68m
Tail area S
T
= 46.5m
2
tail arm l = 15.66m
Elevator area S

= 11.2m
2
elevator mean chord c

= 0.908m
h
0
= 0.16 C
M
0
= -0.06 = 0.38
a = 4.5 a
1
= 2.75 a
2
= 1.16
b
0
= 0 b
1
= -0.133 b
2
= -0.16
The stickelevator gearing ratio m
e
= 1.0m/rad.
[118 kt, pull force]
2. Using the approach of 3.2, and the results of 5.2, derive a formula for dP
e
/dV ,
the gradient of stick force with ight speed. What does this tell you about the
handling qualities of an aircraft?
3. What are stick-xed and stick-free manoeuvre points and what is the signi-
cance of stick force per g.
Using the data of the previous question, calculate the change of elevator angle
required to pull 0.5g ying at 350kt EAS at an altitude where the relative density
= 0.374.
Explain, in simple physical terms, why this change of elevator angle would be
greater at a lower altitude when ying at the same lift coefcient.
[ = 1.005

]
4. The tailless aircraft shown in the Figure has been tted with a small retractable
foreplane. At low speeds this foreplane is extended and, operating in a semi-
stalled condition at constant setting, it generates a constant lift coefcient C
LF
=
1.2 (based on S
F
). Use of the foreplane enables the aircraft to take off at a higher
weight than the original aircraft without the foreplane. Calculate the increment
97
in take-off weight that may be achieved when using the foreplane, by consider-
ing the trimmed lift at 200kt EAS, if the incidence is restricted to 12

by ground
clearance problems, using the data in the table.
Calculate the elevon angles to trim of both versions of the aircraft. Comment on
your results.
[With foreplane: = 0.5

; L = 1842kN; without foreplane: = 5.8

; L =
1557kN]
l
F
c
0
hc
0
h
0
S = 438m
2
S
F
= 9.4m
2
C
M
0
= +0.002 C
M
0
/ = -0.25
a
1
= 3.0 a
2
= 0.80 h
0
= 0.61 c
0
=27.4m
l
F
= 13.26m hc
0
= 15.34m
5. The aircraft described in the Figure and table below is to have its capacity in-
creased by lengthening the cylindrical portion of the fuselage by 6m. The centre
section (including the wings), the nose and the tail portions are to remain unal-
tered.
It is assumed that the c.g. position will be adjusted to remain unchanged with
respect to the centre section unit and that, for the lengths considered, / is
constant.
Calculate how the additional fuselage length is to be inserted ahead of and be-
hind the centre section, if the low speed stick-xed static margin is to unaltered.
The movement of the aerodynamic centre of the aircraft less tail is assumed to
be affected only by the change of nose length l
N
and is given by
h
0
= 0.037
l
N
c
.
If a variant of the aircraft retains the original fuselage, but has its wing tips
extended, how could the longitudinal static stability be affected?
98 CHAPTER 12. PROBLEMS
S = 223m
2
c = 5.6m
S
T
= 46m
2
l=15.5m
h
0
= 0.25 h = 0.20
= 0.4 C
M
0
=-0.06
a = 4.5 a
1
= 2.75
[4.1m ahead of wing, 1.9m aft]
6. a) The 1903 Wright Flyer was a canard conguration of conventional layout,
summarized in the table below. Calculate the stick-xed neutral point,
assuming that the wing and canard have approximately equal lift curve
slope, and comment on your answer.
b) The 1903 Flyer was stabilized in pitch by the addition of ballast to shift
the centre of gravity forward. If 30% of the aircraft gross weight can be
carried as ballast, where should it be placed to move the centre of grav-
ity to the wing leading edge. What effect would this have on the aircraft
performance?
h
0
0 V = 0.134 C
M
0
= 0.141
h = 0.3c W 340kg
The 1903 Wright Flyer. The datum is the wing leading edge.
99
1. The table below contains ight test data for the X-15 spaceplane. Calculate the
static margin stick-xed and estimate the zero-lift pitching moment. Estimate
the dimensional and non-dimensional phugoid mode and SPO frequencies.
S = 18.58m
2
s = 6.82m c = 3.13m
h = 0.22
m = 7056kg B = 10700kgm
2
V = 331kt EAS
a = 3.5/rad C
M
/ = 0.8/rad
Z
u
= 332Ns/m M
w
= 40.7Ns
Z
w
= 14300Ns/m
M
q
= 158600Nms
[
ph
= 0.0946 (0.052rad/s);
spo
= 17.683 (9.723rad/s)]
2. NASA CR-2144, Aircraft Handling Qualities Data, contains stability informa-
tion for ten aircraft. For the Boeing 747:
a) calculate the static margin stick-xed;
b) estimate the zero-lift pitching moment;
c) estimate the phugoid, SPO and Dutch roll periods;
d) estimate the time constants for rolling subsidence and the spiral mode.
100 CHAPTER 12. PROBLEMS
1. a) Given that the velocity of a body in planetary orbit is:
v =
_
2E +
2
r
_
1/2
,
describe the following spacecraft maneouvres and calculate the velocity
change required for each:
i. Hohmann transfer;
ii. orbital capture;
iii. planetary escape.
The parameters E and are the orbital energy per unit mass and the grav-
itational parameter, respectively.
b) Alunar module in lowearth orbit travels to the moon on a Hohmann trans-
fer trajectory. Calculate the V required to enter the Hohmann transfer to
lunar orbit and, for comparison, the escape velocity of Earth.
c) Comparing the V required for transfer to lunar orbit and that to escape
Earths gravity, comment on the feasibility of a manned mission beyond
the moon.
2. a) A communications satellite for the high latitudes of the Northern hemi-
sphere is parked in a circular low earth orbit of altitude 174km. Calculate
the velocity change required to change its orbit to one with a perigee of
174km and an apogee of 39000km.
b) Sketch the shape of the new orbit, indicating the point at which the burn
takes place. Why is such an orbit used in practice?
Basic equations
101
103
Conventional aircraft Tailless aircraft
Steady ight:
C
M
= C
M
0
(h
0
h)C
L
V C
L
T
C
M
= C
M
0
+
C
M

(h
0
h)C
L
C
L
T
=
a
1
a
_
1

_
C
L
+a
1
(
T

0
) +a
2
+a
3

Steady pullout/bank:
C
L
T
=
a
1
a
_
1

_
C
L
+a
1
(
T

0
) +a
2
+a
3
+a
1

T
C
M
= C
M
0
+
C
M

(h
0
h)C
L
+
C
M
q
q
Static and manoeuvre margins:
K
n
= (h
0
h) +V
a
1
a
_
1

_
K
n
= h
0
h
H
m
= K
n
+
V a
1
2
1
H
m
= K
n

m
q

1
K

n
= (h
0
h) +V
a
1
a
_
1

_
H

m
= K

n
+
V a
1
2
1
General relationships

1
=
W
gSl
T

1
=
W
gSc
0
a
1
= a
1
_
1
a
2
b
1
a
1
b
2
_
m
q
=
1
V Sc
2
0
M
q
a
3
= a
3
_
1
a
2
b
3
a
3
b
2
_
C
H
= b
0
+b
1

T
+b
2
+b
3

V =
S
T
S
l
c
x
u
=
X
u
US
, x
q
=
X
q
USc
, m
q
=
M
q
USc
2

c
=
m
Sc
, =
m
US
.
Cranfield University Flying Laboratory Student Registration Form
All students participating in flights aboard the laboratory aircraft, operated by the National Flying
Laboratory Centre, must read and complete this form. Completion and submission of the form is
required to officially register students as members of a Cranfield University short course.
Please note:
1. Personal life insurance cover is usually restricted to scheduled or charter flights, and
does not extend to any other flying unless declared by the insured. Anyone concerned about
insurance are advised, in the first instance, to contact their own insurance company. The NFLC
will supply an interested insurance company with details of the CAA approved Air Operators
Certificate and exemption on request. In addition, special insurance cover is maintained for
students registered on a Cranfield course requiring flights in the laboratory aircraft, and details of
this cover may be obtained from Cranfield Universitys Commercial Accountant.
2. The NFLC reserves the right to refuse permission for any person to fly if, in the opinion
of the crew, the person is not adequately briefed for the flight, or if the carriage of the person would
constitute a hazard to the aircraft or its occupants.
3. Students must be medically fit to fly at the time of the course, if there is doubt advice
from a doctor should be sought.
4. Students weight and sex must be entered (for load planning purposes).
5. The course will be planned to run to a strict timetable but operational considerations
may necessitate last-minute changes; in order to retain the necessary flexibility students should not
make any other commitments (such as part-time working) during the course so that they are
available should rescheduling be required.
6. The NFLC reserves the right to refuse permission for any person to fly if, in the opinion
of the crew, the person is wearing inappropriate footwear or clothing that is prejudicial to their
safety in the event of an aircraft emergency.

Please print your details clearly
University:___________________________ Course:______________________
Name:_______________________________ Weight:___________kg
Address:_______________________________ Sex, M or F: ____________
__________________________
__________________________
__________________________
Tel: __________________________ E-mail:_______________________
Nationality:__________________________
I certify that I have read and understood the notes above, and the details I have supplied are
correct.
Signature:____________________________ Date:_______________
Student Registration V 4
November 2012 NFLC, Cranfield University

Vous aimerez peut-être aussi