You are on page 1of 340

Principles of Flight

and Aeroplane Performance

a P~:tf$c.c~ training manual

Volume 7 of the Commercial Pilot Series


covering the NZ CPL examination requirements in the subject

Stewart Boys
Walter Wagtendonk

p,;,e..;t~<J#
155 Waipapa Road
RD 6 Tauranga, New Zealand.
Phone (07) 548 1654 Fax (07) 548 1652
Email: info@pilotbooks.co.nz
Website:www.pilotbooks.co.nz
The Commercial Pilot Series
PRINCIPLES OF FLIGHT and Aeroplane Performance

First Edition April 1996


Revised (Eleventh) Edition May 2010
Published by:
Aviation Theory Centre (NZ) Ltd. t/a p.;l.,:tEc.cJ,
155 Waipapa Road,
Tauranga, Bay of Plenty,
New Zealand.
Telephone +64 7 548 1654
Fax +64 7 548 1652
e-mail: info@pilotbooks.co.nz
Website: www.pilotbooks.co.nz

Copright © 1996 Aviation Theory Centre (NZ) Ltd.

THE CONTENTS OF THIS MANUAL ARE PROTECTED BY COPYRIGHT


THROUGHOUT THE WORLD UNDER THE BERNE UNION AND THE
UNIVERSAL COPYRIGHT CONVENTION

All rights reserved. No part of this publication may be reproduced in any


manner whatsoever- electronic, photographic, photocopying, facsimile, or
stored in a retrieval system- without the prior written permission of the author.

ACKNOWLEDGEMENTS
Aviation Setvices Ltd., Lower Hutt
Andy Smith, Nelson Aviation College Ltd., Motueka

Cover Photograph: PSI Mustang taken by Glen Alderton during an air show at Wanaka. The effect of the
wingtip vortex from the right wing can be seen in the smoke at lower left.

Nothing in this text supersedes any regulatmy material or


operational documents issued by the Civil Aviation Authority
or the operators of aircraft.

ISBN 0-9583373-6-5

Printed by:

t2rQ"';nf
47 Second Avenue,
TAURANGA
Editorial Team

Walter Wagtendonk OBE Stewart Boys CBE AFC

Principal Director of Formerly Technical Director of


Aviation Themy Centre (NZ) Aviation The01y Centre (NZ)
Ltd., 'Wal' has had a long Ltd., Stewart retired from the
and successful career in RNZAF as an Air Commodore
flying training. Shortly after in 1991 on completion of 37
arriving in New Zealand years service. His Air Force
from Holland, he joined the career included many years
RNZAF and at the experience as a flying
completion of 8 years instructor and flight
seroice, retired as a Flight commander at the Pilot
Lieutenant and A2 flying Training Squadron; at the
Instructor. He became the Central Flying School; and on
CFI of the Nelson Aero Club and in 1978, together operational units. During his period at CFS,
with his wife Ann, started the Nelson Aviation Stewart was responsible for writing the RNZAF's
College which blossomed into one of New first flying instructors' handbook. Later, he
Zealand's most successful and internationally attended a one-year aero-systems (technology)
recognised training organisations. The college was course with the Royal Air Force college at Manby.
the first to operate under 'approval' to conduct One of the few in the Service to gain an AI QFl
both fixed-wing and helicopter training. His work category, Stewart also has experience as an A
was recognised in 1994 when he was awarded the categmy Civil flying instructor.
OBE for services to aviation.

John Wagtendonk

Involved in aviation from a


vety early age, John obtained
his pilot's licence when he
was 16. Becoming an Air
Traffic Controller in 1978, he
is now a highly experienced
radar controller. John has
previously lectured in
Aviation Law at the Nelson
Aviation College and
currently has some
responsibility for training new
radar controllers.

(iii)
Table of Contents
Chapter 1. Aeroscience
Units of Measurement ............................................................ 1-1
Scalar and Vector Quantities .................................................. 1-2
Newton's Laws of Motion ...................................................... 1-3
Motion on a Curved Path ........................................................ 1-7
Vectors ...... ....... .. ....... .. .... .. ........... ..... .. .. ............. .. .. ......... ..... .... 1-8
Moments and Couples ............................................................ 1-9
Equilibrium .............................................................................. 1-10
Centre of Gravity ..................................................................... 1-11
Work. Power. Energy .............................................................. 1-12
Use of Graphs .......................................................................... 1-13
Review 1 ................................................................................... 1-15
Chapter 2. The Atmosphere
Density ..................................................................................... 2-1
Pressure ................................................................................... 2-2
Temperature ............................................................................ 2-3
Density,Pressure and Temperature in the Atmosphere........ 2A
Density Altitude ....................................................................... 2-7
Humidity. Viscosity ................................................................ 2-7
Review 2 ................................................................................... 2-8
Chapter 3. Basic Aerodynamic Theory
Static Pressure ...... .................................................................... 3-1
Dynamic Pressure .................................................................... 3-1
Measurement of Airspeed ...................................................... 3-2
Aerodynamic Force ................................................................ 3-4
Aero foils ........... ........................................................................ 3-5
Angle of Attack ........................................................................ 3-6
Pressure Distribution .............................................................. 3-7
Venturi Effect ...... .... ... .. ........... .. .................. .... .... .... ................. 3-8
Airflow Around an Aerofoil ..................................................... 3-8
Centre of Pressure .................................................................. 3-11
Review3 ................................................................................... 3-14
Chapter 4. Lift
Factors Affecting Lift....................................................................... 4-1
The Coefficient of Lift ............................................................. 4-2
The Lift Formula ...................................................................... 4-2
Variation of CL with Angle of Attack ...................................... 4-3
Three Dimensional Flow Over a Wing ........................................ 4-6
Review 4 ................................................................................... 4-11

Chapter 5. Drag
Classification of Total Drag ..................................................... 5-2
Parasite Drag ................................................................................. 5-2
The Boundary Layer ............................................................... 5-2
Skin-Friction Drag ................................................................... 5-3
Factors Affecting Skin-Friction Drag ...................................... 5-4
Form Drag ................................................................................ 5-5
Factors Affecting Form Drag .................................................. 5-7
Interference Drag .................................................................... 5-9
Induced Drag ................................................................................. 5-9
Factors Affecting Induced Drag ............................................. 5-11
Measures for the Reduction of Induced Drag ....................... 5-12
Total Drag ...................................................................................... 5-13
The Coefficient of Drag ........................................................... 5-13
The Drag Curve ...... .. .... .... ......... .... ......... ....... ...... .... ..... ..... ...... 5-14
Lift/Drag Ratio ............................................................................... 5-15
An Alternative Classification of Drag ........................................... 5-16
Review 5 ..................................................................... 5-17

(iv)
Chapter 6. Lift Augmentation
Trailing-Edge Flaps . ...................................................................... 6-1
Effects of Trailing-edge flap .................................................... 6·2
Types of Trailing-Edge Flap .................................................... 6·4
Slats and Slots ............................................................................... 6·6
Leading-Edge Flaps ...................................................................... 6-7
Combined High-lift devices .......................................................... 6-7
Spoilers .......... ... ... .. ............ ..... .... ..... ... .. .. .. ..... ..... .. .. .. ... ...... ... .... ..... 6-8
Review 6 ................................................................................... 5.g

Chapter 7. Flight Controls


The Primary Flight Controls ......................................................... 7.J
Principle of Operation ............................................................ 7·2
How Control in Flight is Achieved ......................................... 7-3
Effect of Airspeed on the Controls ......................................... 7·7
The Effect of Slipstream ......................................................... 7·7
Unconventional Control Configurations ................................ 7~8
Trim Controls ................................................................................ 7.g
Balancing of Controls ................................................................... 7·I I
Aerodynamic Balance ............................................................ 7·I I
Mass Balancing ....................................................................... 7-14
Review 7 ................................................................................... 7-16

Chapter 8. Stalling and Spinning


Stalling ........................................................................................... 8·1
The Stall and the Lift Formula ................................................ 8·1
Symptoms of the Stall ............................................................. 8-2
The Stall ................................................................................... 8-4
Stall Recovel}' .......................................................................... 8-5
Factors Affecting Stalling Speed ............................................ 8·6
Wing Drop at the Stall ............................................................. 8·1 0
Use of Aileron Near, and During, the Stall ............................ 8-12
Recovel}' From the Wing-Drop stall ...................................... 8-13
Recovel}' at Onset ................................................................... 8-13
Spinning ......................................................................................... 8-14
Autorotation ............................................................................... 8-15
Spin Characteristics. Confirmation ........................................ 8-16
Spin Recove1y .......................................................................... 8-17
Review 8 ................................................................................... 8-18
Chapter 9. Straight and Level Flight
The Forces Acting ................................................................... g.2
Pitching Moments ................................................................... g.3
Variable Effects on the Couples ............................................. g.4
Increasing and Decreasing Speed in Level Flight ................ g.5
Performance in Straight and Level Flight .............................. g.8
The Power Required Curve .................................................... g.8
The Power Available Curve .................................................... g.] 0
Maximum and Minimum Speeds in Level Flight ................. g.] 0
The Effect of Weight and Altitude .......................................... g.]]
Review 9 ................................................................................... g.J3
Chapter 10. Climbing and Descending
Climbing ........................................................................................ I 0·1
The Forces Acting in the Climb ............................................. I 0-2
Climb Performance ................................................................ I 0-4
Factors Affecting Climb Performance ................................... I 0·8
Descending .................................................................................... I 0·1 0
Gliding ...................................................................................... I Q.JJ
Gliding Performance ............................................................... I 0-11
The Power-On Descent .......................................................... I 0-14
Review 10 ................................................................................. 10-15

(v)
Chapter 11. Turning
The Level Turn ........................................................................ 11-2
Load Factor ............................................................................. 11-3
Stalling in Turns ....................................................................... 11-5
Turning Performance .............................................................. 11 ~6
Steep Turns .............................................................................. 11-9
Maximum Rate and Minimum Radius turns ......................... 11-9
Effect of Power ........................................................................ 11-11
Effect of Wind .......................................................................... 11-12
Climbing Turns ... ......... .... ...... ................ ..... .. .. ............. .......... ........ 11-13
RateofCiimb ........................................................................... 11-13
Tendency to Overbank .................. ......................................... 11-13
Descending Turns ......................................................................... 11-14
Rate of Descent ....................................................................... 11-14
Tendency to Underbank ......................................................... 11-14
Manoeuvring in the Vertical Plane .............................................. 11-15
Manoeuvring Limitations ............................................................. 11-17
Speed Limitations ................................................................... 11-17
Load Factor Limitations ......................................................... 11-18
The V-n (orV-g) Diagram ....................................................... 11-18
Review!/................................................................................. 11-20
Chapter 12. Propellers
Terminology ............................................................................ 12-1
Basic Principles ....................................................................... 12-1
Factors Affecting Airflow Across the Blade Sections ...... 12-2
Forces Acting on a Blade Section .......................................... 12-3
The RPM/Airspeed Relationship ............................................ 12-4
Effective Blade Sections ......................................................... 12-5
Propeller Performance ........................................................... 12-6
Slip ............................................................................................ 12-7
Constant-Speed Propellers ........................................................... 12-8
The Constant Speed Unit ........................................................ 12-9
Operation of Constant-Speed Propellers .............................. 12-10
Other Modes of Operation ...................................................... 12-11
Propeller Twisting Moments .................................................. 12-13
Asymmetric Blade Effect ........................................................ 12-15
Propeller Solidity ..................................................................... 12-16
Review 12 ................................................................................. 12-17
Chapter 13. Stability
Static and Dynamic Stability ................................................... 13-1
Stability and Controllability ..................................................... 13-2
Longitudinal Stability .............................................................. 13-3
Factors Affecting the Degree of Longitudinal Stability ......... 13-5
Directional Stability ................................................................. 13-6
Lateral Stability ........................................................................ 13-7
Factors Affecting Lateral Stability .......................................... 13-8
Lateral and Directional Stability considered Together ......... 13-10
Stability and Control an the Ground ............................................ 13-11
Ground Roll Stability ............................................................... 13-11
Control on the Ground ............................................................ 13-12
Swing on Take-Off .................................................................. 13-14
Crosswind Take-Offs and Landings ...................................... 13-16
Ground Effect .. .. .. .. .. ..... .... .. ... .. .. .. ..... .. .. .. .... .. .. ... .. .. .. ... .. .... ... ... . 13-18
Review 13 ................................................................................. 13-20
Chapter 14. Asymmetric Flight
Yawing Moment ...................................................................... 14-2
Rolling Moment ....................................................................... 14-2
Other Factors ........................................................................... 14-2
Immediate Actions .................................................................. 14-4
Modes of Constant-Heading Asymmetric Flight ................... 14-5

(vi)
Minimum Asymmetric Control Speeds ................................. 14-8
14-9
Review 14 ..... .... ... .... .. ..... .... ... .. .. ... .. ... .. .. ..... .... .. .. .... ..... ..... ...... ..
Chapter 15. Range and Endurance
Range Flying - Theory ................................................................... 15-1
Airframe Considerations: Piston Engine Aircraft Range ...... 15-2
Engine Considerations ............................................................ 15-4
Factors Affecting SFC .............................................................. 15-4
Flying for Range: Practical Application ....................................... 15-5
Flying for Endurance ....... ... ..... .. .................................... ...... .......... 15-8
Engine Considerations ............................................................ 15-9
Practical Application ............................................................... 15-9
Review 15 ................................................................................. 15-11
Chapter 16. High Speed Flight
Flow Regimes ......................................................................... 16-2
Sound Waves .......................................................................... 16-3
The Speed of Sound ................................................................ 16-4
Mach Number ......................................................................... 16-4
Speed Ranges ..........................................................................16-6
Formation of Shockwaves ........................................................... 16-7
Pressure Waves From a Moving Source ............................... 16-7
Shockwaves ............................................................................ 16-9
The Bow Shockwave .............................................................. 16-10
Wing Shockwaves .................................................................. 16-11
Expansion Waves .................................................................... 16-12
The Nature of Supersonic Flow ............................................. 16-13
The Effects of Compressibility on Lift .......................................... 16-15
The Effects of Compressibility on Drag ...... ........................ .... ..... 16-18
Control at High Speed ................................................................... 16-19
Longitudinal Control ............................................................... 16-19
Lateral Control ......................................................................... 16-21
Directional Control .................................................................. 16-21
Design for High Speed Flight ........................................................ 16-22
Wing Thickness/Chord Ratio ................................................. 16-22
Supercritical Wing Section ..................................................... 16-23
Sweepback .............................................................................. 16-23
Disadvantages of Sweep back ................................................ 16-24
Area Rule ................................................................................. 16-27
Design for Supersonic Flight ......................................................... 16-28
Wing Sections ............................................................. ............ 16-28
Supersonic Planform Shapes ................................................. 16-29
Review 16 ................................................................................. 16-31
Chapter 17 Performance
Terms and Definitions ..................................................................... 17-1
Take-off Distance required (TODR) ......................................... 17-1
Take-off Distance available (TOOA) ......................................... 17-2
Accelerate-stop Distance ........................................................... 17 -2
Gradient of Climb ...................................................................... 17-3
Gross and Net Flight Paths ........................................................ 17-3
Landing Distance Available (LOA) ........................................... 17-4
Landing Distance Required (LOR) ........................................... 17-4
Dry Runway ................................................................................ 17-4
Wet Runway ............................................................................... 17-5
Contaminated Runway ............................................................. 17-5
Drift Down ................................................................................... 17 -5
Factors Affecting Take-off and Landing Performance ................ 17-5
Air Density ................................................................................... 17 -6
Weight ........................................................................................ 17-7
Wind ............................................................................................ 17-7
Runway Slope ............................................................................ 17-7
Runway Surface and Condition ............................................... 17-7

(vii)
Wet or Contaminated Runways ............................................... 17-7
Other Factors .............................................................................. 17 -8
Calculaling Pressure Altitude and Density Altitude ...................... 17-8
The International Standard Atmosphere ................................. 17-8
Pressure Altitude ........................................................................ 17-9
Using The Altimeter to Determine Pressure Altitude ............. 17-1 0
Density Altitude .......................................................................... 17 -II
Calculation of Temperature Deviation ..................................... 17-12
Calculation of Density Altitude .................................................. 17-13
Calculation of Take-off and Landing Distance Required ............ 17-16
Use of Aircraft Flight Manual Data to obtain TODR and LDR ..... 17-17
Take-off Graph ........................................................................... 17-17
LandingGraph ............................................................................ 17-19
General Notes on the use of Flight Manual Graphs ............... 17-20
Headwind and Crosswind Component Graph ..................... 17-21
Runway Slope and Surface Correction Factors ..................... 17 -22
Use of Older-style P Charts .............................................................. 17 -26
Take-off Graph ................................................................................. 17-26
Establishing Density Altitude .................................................... 17-26
Allowing for All-up Weight ........................................................ 17-28
Allowing for Runway Surface ................................................... 17-28
Allowing for Slope ...................................................................... 17-29
Allowing for Wind Velocity ....................................................... 17-29
Calculating Maximum All-up Weight ....................................... 17-30
Landing Graph ................................................................................. 17-34
Use of Later ?-charts ........................................................................17 -36
En-route Engine Inoperative Performance .................................... 17-39
Single-engine service Ceiling Graph ........................................ 17-39
Use of Tabulated Performance Data ............................................. 17-43
Review 17 ...................................................................................... 17-46

Appendix 1 Sample Examination .... .............................................. Appendix 1-1


Appendix 2 Answers to Review Questions
and Sample Examination ......................................... Appendix 2-1
Index lndex-1

(viii)
Introduction

This manual is aimed primarily at providing the knowledge required by


candidates for the New Zealand Commercial Pilot Licence in Principles of Flight
and Performance (Aeroplane). Chapters 1-15 cover Part I of the examination
(Principles of Flight) while Chapter I 7 covers Part II (Performance). Please note
that Chapter I 6, 'High Speed Flight' is not required reading for the CPL
examination. It is included not only to 'round out' the subjects covered by the
manual but also as being suitable for study toward the ATPL requirement for
Advanced Aerodynamics.

The manual is one of the 'Commercial Pilot Series' published by p~;(JS~<>t-.. This
series comprises:
Vol. 2 Navigation and Flight Planning (I)
Vol. 5 Air Law (I)
Vol. 6 Meteorology for Professional Pilots (2)
Vol. 7 Principles of Flight and Aeroplane Performance (2)
Vol. 8 General Aircraft Technical Knowledge.
Notes:
(!) Also part of the 'Recreational Pilot's Series'.
(2) Also suitable for ATPL studies.

Candidates for the New Zealand Commercial Pilot Licence (Helicopter)


examinations are referred to the book Principles of Helicopter Flight by W.J.
Wagtendonk which is published in the USA and is available by contacting
p~g,~ direct.

Also published by P~:tg,~ for candidates for the New Zealand Instrument
Rating are the companion volumes:
Vol 9. The Instrument Rating Manual; and
Vol I 0. Instrument Rating Law.

The information in this manual is presented in a direct, easy-to-follow manner


which is practical and thorough. Studying the manual and completing the
review questions will prepare you for the examination and the knowledge
gained will make you a better, safer and more confident pilot.

Review Questions
At the end of each chapter you will find a review which consists of a series of
simple but comprehensive questions. As you finish reading a chapter you
should take a break and complete the review. When you have written down
your answers, you can refer to Appendix 2 at the back of the manual to compare
your results with ours.

(ix)
Sample Examination
A sample examination is included at Appendix I. It is similar to the
examinations prepared by Aviation Services Ltd (ASL) to test candidates for the
CPL. Once you have finished reading the manual and completed the chapter
reviews, you should test yourself under simulated examination conditions. The
answers are given after the review answers in Appendix 2.

The chapter reviews and sample examination together provide a ready means of
'recap' prior to sitting the ASL examination.

Study Assistance
would be pleased to assist in answering any queries on study for the CAA
licences generally, or on any specific area covered by the manuals which we
publish. Please phone, fax, or send an e-mail message to any of the addresses
listed on the first page of this book.

Best wishes for successful, safe and enjoyable flying from all at P~:t"{;c.cJo..

(x)
Aeroscience

Introduction

A familiarity with certain basic mechanical and physical principles is essential to


a good understanding of the principles of flight and aircraft performance. This
chapter provides a summary of those mechanical and physical laws, concepts
and plinciples which are important to a comprehension of the chapters which
follow.

This manual assumes that readers will have a basic capability in mathematics-
particularly with handling and understanding simple algebraic equations. At the
level at which the manual is pitched, explanations of the principles of flight are
liberally sprinkled with such equations, as this is often the simplest and most
direct way to relate the factors involved and gain an understanding of how they
interact. While a working knowledge of these equations is unavoidable, most
will be relieved to find that in nearly all cases there is no requirement to actually
apply them to work out numerical answers to problems.

Readers who have studied physics at Secondary School should already be


familiar with the subject matter of this chapter. If this is the case, the chapter
can be treated as a refresher. Those who are newcomers to the subject are
advised that what follows is but a brief summary of the more important
principles involved. They may find that a separate study of a good secondary-
level text book on physics (available from most bookstores or libraries) will be
helpful, not only in fully understanding this manual, but also toward their
preparation for the Commercial Pilot Licence examinations in other subjects.

Units of Measurement
Throughout the world, the International System of units (or Sl system) is
increasingly being used for the measurement of physical quantities. The
fundamental SI units which we are concerned with in this manual are:

• Length - metre (m).

• Mass - kilogram (kg).

• Time -second (s).

• Temperature- degrees Celsius COC).

(The remaining fundamental units in the SI system-but which are of no


concern to us in this manual are; electric current-the ampere; luminous
intensity-the candela; amount of substance-the mole.)

NOTE: Strictly speaking, the degree Kelvin (K) is the fundamental unit for
temperature and the oc is a derived unit. Although the starting points of the
Kelvin and Celsius scales are different, I oc change in temperature is equivalent
to one degree K change in temperature, and for practical purposes it is usually
much more convenient to use the Celsius scale, where ooc and I oooc
respectively represent the freezing and boiling points of water.

Principles of Flight Aeroscience 1-1


Derived units are those which can be formed by the combination of the
fundamental units. There are many derived units, but some which are pertinent
to this manual are:

• Force - newton (N). A newton is the force required to accelerate a I kg


mass at lm per second per second. You will sometimes see the term
kilograms force (kg/f) being used. I kg/f = 9·81 N.
• Pressure -pascal (Pa). I Pa = I newton per square metre (Nm2 ).
• Power - watt (W). Power is force x distance per unit time and watts are
equivalent to newton metres/second (Nm/s). Although the power of
modern car engines is rated in kilowatts (kW or l,OOOW), most light aircraft
engines are still rated in horsepower. For interest, I horsepower = 746 W
(or roughly% kW).

Non-S! Units used in Aviation

In aviation some units from the older 'Imperial' or 'English' system are still
almost universally used, and are likely to be for some time. These are:
• altitude- measured in feet (ft). I ft = 0·3048 m.
• navigational distance - measured in nautical miles (nm). I nm = I ,852 m.
• Speed (airspeed, groundspeed, wind speed)- measured in knots (kt).
I kt = 0·514m/sor 1·85km/hr.

Scalar and Vector Quantities


A scalar quantity is one which has only magnitude (or size). When we refer to
a scalar quantity we are concerned only with the amount of the quantity and it is
either unimportant or not possible to specify the direction of the quantity. An
example of a scalar quantity is temperature.

A vector quantity is one which has both magnitude and direction. When
referring to vector quantities, a direction must be specified, othe1wise the
quantity becomes scalar in nature. An example is windspeed. If speed alone is
stated, windspeed is a scalar quantity. If both speed and direction are stated, it
becomes a vector quantity which, in aviation, is referred to as the wind velocity
(W/V) vector. Vector quantities can be represented in a graph or diagram by an
arrow (or vector) with its length representing magnitude and the arrowhead
representing the direction of the quantity. We will be discussing vectors in more
detail later in this chapter.
'
length _..-----"',
'
indicates magnitude head
Fig. 1-1. of the quantity
',~ indicates
' direction of
the quantity

tail

Describing Motion
The principles of flight are based to a large extent on the study of the movement
of aircraft through the air. There is therefore a need for a clear understanding of
what is meant by the different ways of describing motion.

1-2 Aeroscience The Commercial Pilot Series


Speed

Speed is well understood by most people. It is distance travelled in unit time,


e.g. metres/second, kilometres/hour, or-of importance in aviation-in nautical
miles/hour (or knots). In the way in which we use the term, speed is a scalar
quantity.
distance travelled
speed=
time taken

Velocity

Velocity is the vector equivalent of speed. It is distance travelled in a given


direction in unit time. The term velocity is used (rather than speed) when the
direction of travel is important. We will be using velocity vectors in many of the
diagrams in the following chapters.

distance travelled in a given direction


velocity=
time taken

Acceleration

Acceleration in a straight line will also be familiar to most readers. If speed is


increased, the acceleration is said to be positive. If speed is decreased, it is
negative-and usually referred to as deceleration.

Acceleration also occurs when there is solely a change in direction but not in
speed. Consider an aircraft which is turning with a constant speed registered on
the airspeed indicator. Although its speed through the air is constant, the aircraft
is accelerating because its direction is changing. The direction of the
acceleration is toward the centre of the curved path which exists at any given
moment.

It follows from the foregoing that acceleration occurs when there is both a
change in speed and direction. In summary, acceleration occurs whenever;
• there is a change in speed; or
• there is a change in direction; or
• there is a change in both speed and direction.

A more concise way of phrasing the foregoing is to state that:


• acceleration occurs whenever there is a change in velocity.

Newton's Laws of Motion


Sir Isaac Newton (1642 - 1727) was the first to describe three simple rules
governing the motion of objects in his work The Principa which was written in
Latin and published in 1687. We will be discussing each of these three laws in
detail as we progress through this chapter.

Newton's First Law

Newton's First Law can be written in English as:

"Every object continues in its state of rest or of uniform motion in a straight


line, unless acted upon by an external force."

Principles of Flight Aeroscience 1-3


Mass and Inertia

The tendency of an object to continue moving at a certain velocity (or to remain


at rest with zero velocity) is called inertia. It is a property of, and depends solely
on the mass of the object. A common definition of mass is that it is 'the amount
of matter in an object'. In the SI system, mass is measured in kilograms.

Mass and inertia cannot be separated-objects with a lot of mass have greater
inertia than those with less mass. Consider for example, a railway wagon
standing on a straight piece of level track. Because of its considerable mass, a
large shunting force must be applied to overcome its inertia and to get it moving.
Once it is moving at a constant velocity it will tend to retain that velocity and the
shunting force can be reduced to that which is just sufficient to overcome rolling
resistance- the hictional forces and air resistance. By comparison, a lower-mass
vehicle like a supermarket trolley requires less force to get it moving because it
has less inertia.

small force I !ow mass I


required to
overcome inertia

Fig. 1-2. Large masses require a large force to overcome their inertia.

Force

Newton's First Law also indicates what a force is, and what it does. A force can
be described in various ways - as a push or a pull, an attraction or repulsion, an
influence, or a pressure on an object. Whatever the description, a force is
identified by what it does- it tends to move the object out of its state of rest or of
uniform motion in a straight line. Or, put another way, if a force exists, it is
always trying to accelerate the object which it is acting upon. Whether a force is
successful or not in causing an acceleration depends on whether it is balanced
by other opposing forces.

Consider again our example of the stationa1y railway wagon. If the shunting
force at one end of the wagon is opposed by an equal and opposite force at the
other end, the wagon will remain stationa1y. In this situation, the forces acting
on the wagon are in equilibrium and no acceleration is caused. Before the
wagon can begin to move (or accelerate) one or other of these balanced forces
must be removed or reduced so that a net force is available to cause the change
in velocity. In Newton's words, an 'external' force must exist.

. - - - - - - - - - - - - from the force applied:


inertia and this amount is needed to overcome
rolling resistance inertia and rolling resistance
this net amount remains
to provide acceleration

Fig. 1-3. A net force must be available to cause an acceleration.

1-4 Aeroscience The Commercial Pilot Series


Newton's Second Law

"The external force acting on a body is proportional to the product of its


mass and the acceleration produced by the force" That is, F oc rna.

If a consistent system of units is used, such as the SI system, Newton's Second


Law can be written as:
Force= mass x acceleration. (F = ma).
This equation neatly sums up the connection between force, mass and
acceleration. If a net (or unopposed) force is applied to a mass it will cause a
certain acceleration. If the mass is increased, the force required to achieve the
same acceleration must also be increased. Alternatively, if the force is increased
but the mass remains the same, the acceleration will be greater. Under the SI
system, where the force is in newtons (N), and the mass in kilograms (kg), the
resulting acceleration will be in metres per second per second (m/s 2 ).

Weight

Weight is not the same thing as mass. As we have seen, mass is the 'amount of
substance in an object' whereas weight is a force-it is the force produced
when an object (or mass) is acted upon by gravitational attraction.

All objects positioned within the earth's gravitational field are attracted toward
the centre of the earth by a force which we call weight. The magnitude of this
force of gravity (or weight) depends on the mass of an object and its distance
from the centre of the earth. No matter where it is taken in the universe, the
mass of an object does not vary-unless something is physically done to change
it. A !kg block of cheese is a !kg block of cheese whether it is on the earth or on
the moon. However, as the distance from the centre of the earth increases the
force of gravity (and therefore weight) decreases. At an altitude of about 2,000
km the strength of the earth's gravitational field is approximately halved and, at
that point, while the mass of our I kg block of cheese would remain unaltered, it
would weigh approximately 0·5 kg/f. On the surface of the moon, where the
strength of the moon's gravitational field force is about one-sixth of that on the
earth's surface, a !kg mass weighs about 0·16 kg/f.

The strength of the earth's gravitational field varies slightly around the surface of
the globe and decreases very slightly as normal flying altitudes are gained. For
all practical purposes however, it is taken to be constant at all places on or near
the surface of the earth. That 'constant' attraction of gravity is such as to cause
(or t1y to cause) an acceleration of 9·81 m/s 2 on all masses. The force (i.e.
weight) which is required to produce that constant acceleration therefore varies
depending on the mass of an object. From Newton's Second Law (F = rna), we
can say:
W=mg
(where W =weight, m =mass, g =the acceleration caused by gravity)
In the Sl system, an object's weight on or near the earth's surface will be:
Weight (inN) =mass (in kg) x 9·81 (m/s 2 )
It can be seen from this equation that a I kg mass has a weight of 9·81 newtons.
To simplify matters in normal life, this weight (or force) of 9·81 newtons is called
a kilogram force, meaning the amount of force required to accelerate a I kg
mass at the normal gravitational rate of I 'g'. If we divide both sides ofthe above
equation by 9·81, or alternatively say that g-(an acceleration of 9·81 m/s 2)-is
equivalent to I 'g' we can then say:
Weight in kilograms force (kg/f) = mass in kilograms (kg).

Principles of Flight Aeroscience 1-5


To simplify matters further (or perhaps to confuse the issue) we often refer to
weight in kg (as opposed to kg/0 and hence the unit 'kilogram' is commonly
used to measure both mass and weight. This is a reasonable assumption at or
near the surface of the earth, where a I kg mass will weigh I kg/f-or very close
to it. Sometimes however, we must be careful to distinguish between the two
and it is as well to keep in mind that there is an essential difference between
kilograms weight (a force) and kilograms (mass).

To summarize:
• Weight is the force which gravity exerts on an object. On or near the earth's
surface an object will weigh:
- in newtons- 9·81 times its mass in kilograms;
- in kg/f-the same as its mass in kilograms.
• g is the acceleration due to gravity. For all masses on or near the earth's
surface, it is taken to be constant at 9·81 m/s'.
• 'g' (with inverted commas) is a term which can be applied to any
acceleration. I 'g' ; the normal acceleration of gravity; 9·81 m/s 2 •

Note the difference between the terms g and 'g' and be careful with your use of
them. When we use the term g, (without inverted commas) we mean an
acceleration of 9·81 m/s 2 . When we use the term 'g' we mean an acceleration
which is n times greater than the normal acceleration due to gravity. For
example 2 'g' means an acceleration which is twice as great as gravity (or 2 x
9·81 m/s 2). A person subjected to an acceleration of 2 'g' will feel twice as heavy
as his or her normal weight. 3 'g' is 3 times the normal acceleration of gravity-
and so on.

Momentum

The momentum of an object is the product of its mass and velocity, i.e.
momentum = mv

The term momentum refers to the difficulty of stopping a moving object, or of


changing its direction of travel. A large mass with low velocity can have the
same momentum as a small mass with high velocity, and each will be just as
difficult to stop or to turn away from a straight-line course. Increased mass and/
or increased velocity means an increase in momentum.

Momentum is often confused with inertia. Note that from the above definition a
stationary object with no velocity has no momentum (although it retains its
inertia). As we have seen previously, inertia is solely a property of mass, and two
objects with the same mass have the same inertia. If however, those objects are
moving and one has a higher velocity it will have a greater resistance to being
stopped, slowed down or being turned because it has greater momentum (and
not solely because of its inertia).
Newton's Third Law

"For every action, there is an equal and opposite reaction."

If a force is applied to an object, an equal reaction force will be experienced in


the opposite direction. If you were, for example, to stand and push against a
wall with your hands, the wall would push back against you with the same force
which you were using. Imagine doing the same thing again, but this time with a
pair of roller skates on. The reaction to the push you apply in one direction will

1-6 Aeroscience The Commercial Pilot Series


force you to move in the opposite direction.-and, of course, the harder you
push the greater the reaction you would experience in the opposite direction.
Another example of Newton's Third Law of 'action and reaction' is seen in the
operation of the jet engine. This type of engine works on the principle of moving
a stream of gases at a high-velocity in one direction to achieve a useful reaction
force (thrust) in the other.

In vector diagrams showing forces in later chapters we will, for simplicity, not
normally show the reaction forces. It is as well to keep in mind however, that all
forces are paired with a reaction, but it is simply that-a reaction. Once the
primary force is removed or altered, so is the reaction to it. And, although it
should be obvious, reactive forces cannot exist alone.

Motion on a Curved Path


A significant proportion of almost every flight is taken up with turning or
otherwise manoeuvring the aircraft on a cmved path. If any object (including an
aircraft) is to follow a curved path, a force must be applied to accelerate it
toward the centre of the curve. As has already been discussed, an object is
subjected to an acceleration if its velocity-either its speed and/or direction-is
changed. Any motion on a cmved path, where direction is continually being
changed, therefore involves acceleration regardless of whether speed is also
changing.

Fig. I-4 shows an object in motion on a curved path. The direction of its velocity
vector (v) at any instant is tangential to the curve. The turning force and the
resulting acceleration are directed along the radius (r) of the curve toward the
centre. This radial force is called centripetal force (CPF). Radial acceleration
is given by v2/r and, substituting in the equation F = rna, centripetal force may be
calculated by:
CPF = m x v2/r
It is usually more convenient to express this equation in terms of weight, rather
than mass. We know that W = m x g, and therefore that m = Wig. Again
substituting, this time for mass in the above equation, the expression for
centripetal force then becomes: Wv '
CPF= gr

velocity
force required toward
centre of turn (v)
Wv'
gr ...
I
I
I
I

rate of turn
or angular velocity
depends on
-vr

Fig. 1-4. For an object to follow a curved path, a force must be applied toward the centre
of the curve.

Principles of Flight Aeroscience 1-7


This equation for centripetal force is used when considering the factors involved
in turning an aircraft or manoeuvring it in the vertical plane such as during a pull-
out from a dive. In these situations, g-the normal acceleration due to gravity-
can be considered to be a constant. It can be seen that to turn the aircraft or to
pull out of a dive at a given radius, as weight and or velocity are increased, so
must the centripetal force which must be generated.

The rate of tum (or the rate at which the direction of the aircraft is changing) is
also sometimes a consideration. The rate of turn depends on the ratio v/r which
is a part of the equation for CPF. At a given aircraft weight, it can be seen that as
velocity is increased, centripetal force must also be increased if the same rate of
turn is to be maintained.

We will be revisiting this equation for CPF when we consider the factors
involved in turning and manoeuvring in more detail in a later chapter.

Vectors
Vector diagrams are frequently employed in explanations of the Principles of
Flight, particularly of the forces acting on an aircraft. The following few simple
rules for handling vectors will help in interpreting those diagrams.

Vector Addition and Subtraction

Two (or more) vectors acting in the same direction can be combined by adding
or subtracting them as appropriate to give a single resultant (Fig. 1-Sa).

A +B A -B
Fig. 1-Sa.
resultant resultant
Two (or more) vectors acting in different directions can be combined into a
single resultant by drawing them with the tail of one vector connected to the
head of another. The resultant is the vector which can be drawn from the tail of
the first, to the head of the last (Fig. 1-Sb).

B B
A A
Fig. 1-Sb.
-
----u-;t;;:nt
res·'
-- c
resultant

Where two vectors can be drawn as acting from the same point, the resultant
may be found by drawing a parallelogram (Fig. 1-Sc). The resultant is formed by
the diagonal which originates from the intersection of the component vectors.

A __ ::p
Fig. 1-Sc.
--
--
- - - - nt ,"
resu.\ta ""
'
'
B B

1-8 Aeroscience The Commercial Pilot Series


Resolving Single Vectors

Any single vector can be resolved into two components which are at right
angles to each other, as shown at Fig 1-Sd. Note that the components can be
drawn in any direction to suit, the only requirement being that the vector must
be included within the right angle.
·f--
''

2
I '
I '' '
component A-! It '
I
I I
component A-\ . . . ~.
Fig. 1-5d. y;
1 ......1 omponent B
component B

Use of Trigonometric Functions

Trigonometric (trig) functions are sometimes used to describe the relationship


between vectors and their components. The t1ig functions sine (sin), cosine
(cos), and tangent (tan) are simply numbers representing the ratios between the
length of the different sides of a right-angle triangle, as shown at Fig. 1-Se.

Fig. 1-5e. ratios of sides depend on the size


of the angle a and are given by:
opposile (o)
0 a
sin a=
h
cosa=-
h
tan a= a
0

If for example, we wish to know the relationship between the component (o)
and the vector (h) in the triangle illustrated above, it is given by o = h sin a.
Similarly,the relationship of (a) to (h) is given by a = h cos a, and (o) to (a) by
the equation o = a tan a. The precise figures can be determined by looking up
the trig tables for the angle (a) involved. Of note however is that the cosine of
small angles (up to about 15°) is very close to 1, in which case when such small
angles are involved, the length of side (a) can for practical purposes be taken to
be the same as side (h).

Moments and Couples


A moment is a turning force about a pivot or turning force
point. Other terms used for a turning force are torque pivot point 1
1
A---. distance 3m
and leverage. The strength of a moment is given by
\J)r--------------.~~:
the product of the force and the perpendicular 1__I
I
distance (called the arm) between its line of action I
I
and the pivot point. In the Sl system, the moment of a

I

force is measured in newton metres (N-m). You will


moment6 Nm 2N
also see moments expressed in kilogram-millimetres
(kg-mm)-particularly in aviation where the weight
and balance of an aircraft is concerned. Fig. 1-6. The moment of a force.

Principles of Flight Aeroscience 1-9


A couple consists of two equal but opposite 2N
parallel forces When the forces of a couple
act around a point mid-way between the force
two, a turning force, usually referred to as distance 2m
moment
torque, will be produced. The torque (or torque)
produced is the product of one of the forces i 4Nm
pivot
and the perpendicular distance between the
two- Again, in the Sl system, torque is force
measured in N-m. 2N
Fig. 1-7. The moment of a couple.

When applied to an object, both moments and couples will cause rotation about
the pivot, provided that they are not counteracted by moments or couples which
act equally in the opposite direction. If the forces involved in a moment or
couple can be made to change their direction as the object turns, as is achieved
for example by the connecting rods/crankshaft in a piston engine, the rotmy
motion can be made to continue.

Equilibrium
An object in a state of rest or moving at a constant velocity is said to be in a state
of equilibrium. In these circumstances the sum of all the forces acting upon it is
zero-no net force will exist to cause it to accelerate.

Fig. 1-8 (a) shows three forces acting on an object in different directions. At first
sight, they may appear to be unbalanced but, as indicated at Fig. 1-8 (b), the
resultant of forces A and B is equal and opposite to force C. The sum of the
forces is therefore zero and the object is in equilibrium. Note that the state of
equilibrium does not necessarily mean that the object is at rest. The object
represented at Fig. 1-8 could well be moving at a steady speed in any direction.
All that the statement of being in equilibrium tells us is that the object is not
accelerating.

~-------------
1 ~
I ~~
B I ">)!....... B
~
~
~

A A
c c

Fig. 1-8. (a) (b)

The foregoing describes the conditions for translational equilibrium. It is also


important to consider whether an object may be in rotational equilibrium. For,
although the object in question may be staying in one place or travelling with a
constant velocity, it is possible for it to be rotating at the same time.

An object is in rotational equilibrium when it has no angular acceleration and the


sum of all the moments acting on it is also zero. It will either be rotating at a
constant rate, or not rotating at all. For this to occur, all of the 'clockwise'
moments must be balanced by the 'anti-clockwise' moments-or, in other
words, if a torque exists in one direction, it must be balanced by a torque in the
opposite direction.

1-10 Aeroscience The Commercial Pilot Series


Fig 1-9 shows a beam which is balanced on a pivot point. For the beam to be in
balance (or in rotational equilibrium), clockwise moments must equal anti-
clockwise moments. We see that this is the case in the diagram-the weight of
2 newtons acting at a distance of I metre from the pivot point produces an anti-
clockwise moment of 2N-m which is the same as the clockwise moment
produced by the weight of I newton acting at a distance of 2 metres from the
pivot. (For simplicity, we have assumed the weight of the beam to be
negligible!)

i - 1 m --.J---2m ------+1
anti-clockwise clockwise
moment moment
2Nm 2Nm

Fig. 1-9. For balance, clockwise moments must equal anti-clockwise moments.
engine torque
A second example of rotational equilibrium is given at Fig. torque~ from drag
1-1 0 which depicts an aircraft propeller which has a
constant rate of rotation (constant revolutions per minute-
or rpm). For the rpm to remain constant, the torque
produced by the engine and transmitted through the hub of
the propeller must be balanced by the torque produced by
the air resistance (or drag) on the propeller. If these
torques are not in balance, the rpm will increase or
decrease depending on which of the two is the stronger.
Fig. 1-10. For constant rpm, the
torques acting on the object must be
in balance.

Centre of Gravity (CG)


The weight of an object can be tal<en to act through its centre of mass-more
often called the centre of gravity (CG). If an object is suspended from its CG
(or a point directly above it), it will balance and not rotate. If the suspension
point (or some other pivot point) is displaced horizontally from the CG, a
moment will exist and the object will not balance.

object suspended from CG

with horizontal
displacement

/
!/ and point
between CG
of suspension
no turning X 1 a moment will exist
moment I
weight

Fig. 1-11. The weight of an object can be taken to act throught its CG.

Principles of Flight Aeroscience 1-11


Work
Work is done when a force moves an object through a distance, i.e.

work = force x distance moved in the direction of the force.

If F is the force in newtons and d is the distance in metres, the work done in Nm
or joules (symbol J) is:

W =fxd

One joule of work is done when a force of I newton moves an object through a
distance of I metre.

Power
Power is the rate at which work is done, i.e.
work done fx d
power= =
time taken t

In the Sl system, power is measured in watts (symbol W). One watt = one joule
per second. The watt is a fairly small unit so, when we refer to the power of an
engine, we normally use units numbering thousands of watts, or kilowatts (kW).

Engine power is a very important determinant in the performance of aircraft. All


aircraft engines regardless of type 'do their work', in effect, by applying force to
a mass of air and accelerating it rearward. As a result, a fmward thrust force is
generated (in accordance with Newton's Third Law) which propels the aircraft
in a fmward direction. The greater the power which can be produced by the
engine and effectively applied to moving a mass of air rearward, the greater the
thrust which will be generated as a result.

We will be considering engine power in more detail in later chapters.

Energy
Energy is the capacity to do work. Not surprisingly, it is measured in the same
units as work-joules.

Energy comes in many forms-chemical, mechanical, electrical and so on-the


list is seemingly endless. When work is done, energy must be converted from
one form to another. In an aircraft engine for example, when the fuel is burned,
the chemical energy stored in it is converted to heat, light and sound. A large
proportion of the heat energy is then converted into mechanical energy to
provide the thrust force as described in the previous paragraph. In this manual,
we are concerned with two kinds of energy-kinetic energy and potential
energy.

Kinetic Energy

Kinetic (or dynamic) energy is the energy due to motion. Every moving mass
has kinetic energy and therefore the capacity to do work. The amount of kinetic
energy which it has depends on its mass and its velocity relative to the object
which it is able to 'work on', and is measured by the equation:

kinetic energy = %mv'

1-12 Aeroscience The Commercial Pilot Series


In a world which is full of fast moving objects, everyday examples of kinetic
energy abound. Evety time there is a collision between those objects, work is
done-and it is not always useful! The kinetic energy of a falling mass is utilised
in a pile driver-the heavier the weight and the greater the velocity at impact the
further the pile will be driven in (more work will be done). Note the squared
relationship between velocity and kinetic energy. If the velocity of a moving
object is doubled, its kinetic energy and its ability to do useful work (or damage)
will be quadrupled.

Potential Energy

Potential energy is the energy of position. Water stored at altitude (in a high
position) in a dam has gravitational potential energy. Because of its position, it
can be made to flow under its own weight, gain kinetic energy and do work-for
example, by being directed through a turbine and used to generate electrical
power. An aircraft at altitude also has gravitational potential energy which can
be readily converted into kinetic energy and allow it to maintain flying speed in a
descent with little or no power from the engine.

Use of Graphs
Graphs are employed extensively in this manual to demonstrate the relationship
of one variable against another. The following is a brief introduction to
interpreting and using graphs for those who may not have had that experience.

The type of graph most commonly employed indicates the value of one variable
(usually the dependent or uncontrolled variable) on a 'vertical' axis called they-
axis, and the value of the independent (or controlled) variable on a 'horizontal'
axis called the x-axis. The relationship between the two variables is then drawn
in the form of a straight line or a curve on the graph. The gradient of the line or
shape of the cutve is obtained from an equation or by plotting the results of an
experiment (or empirically). In most cases, we will not be showing actual
values on either axis-but merely indicating that values increase upward on the
y-axis and to the right on the x-axis.

Fig. 1-12 gives an example of a simple straight- (y)


line graph, in this case of distance travelled
against (or versus) time. The line on the graph
is for a constant 50 kph and, to determine how E
far one would travel at that speed after a certain c.
"0
time, it is simply a matter of selecting the time .!!! /
on the x-axis, tracing vertically upward until the ~ 6 /
line is met, and then across to the y-axis to read _g ------- ---------+
off the distance travelled. As in this case, Q) 4 /
() /
straight-line graphs originating from zero at the c:
intersection of the axes indicate a direct l9
.!'! 2
relationship exists between the variables. The "0

steepness (or gradient) of the line gives an


indication of how rapidly the relationship
between the variables is changing. Also shown 10 20 30 40
at Fig. 1-12, is a dashed line representing a time in minutes
speed of I 00 kph. As you would expect, the
distance travelled after a given time is greater Fig. 1-12. Graph of time vs distance.
than at 50 kph, and the gradient of this line on
the graph is steeper.

Principles of Flight Aeroscience 1-13


Most of the graphs in this manual will show curves instead of straight lines,
indicating that the relationship between the variables is not a direct one. Often,
these curves will have been established empirically from experiments like wind-
tunnel tests. The rules for 'reading' these graphs are the same as has just been
described-to establish a relationship between a variable on the x-axis with one
on the y-axis simply move up to strike the curve, then across. A relationship can
equally be established in the opposite direction by moving from they-axis across
to the curve, and then down to the x-axis.

An example of this type of graph is given at Fig. 1-13, which represents fuel
consumption versus speed for a certain car when it is being driven on a level
road. The curve will be typical of most cars which have been designed to have
best fuel efficiency at moderate speeds. Some points which have general
application in interpreting this type of CUIVe are:

• The lowest point of the cUive gives the driving speed which will achieve the
lowest fuel consumption. This is the speed which would enable the car to
remain driving on the road for the longest time before the fuel runs out.

• Note that the curve is very 'flat' at the bottom. Within this area, there is a
range of speeds over which there will be little difference in fuel
consumption.

• At high and low speeds, the gradient of the curve becomes markedly
steeper, indicating that the fuel consumption is changing more rapidly. Note
that on either side of the minimum consumption (or endurance) speed,
there are two speeds which will achieve the same consumption.

• The speed which will give the greatest range-or greatest distance driven for
the least amount of fuel used-occurs where a line drawn from the origin is
tangential to the curve. At any other point on the cUive, the ratio between
speed gained and fuel used (represented by the length of arrow x compared
to the length of arrow y) will not be as good.

fuel
consumption

A B speed

A= speed for best endurance


B = speed for best range

Fig. 1-13. Sample fuel consumption vs speed graph.

1-14 Aeroscience The Commercial Pilot Series


Review 1

I. The SI unit for force is the ...................... , abbreviated to .......... ..


Sometimes, the term kilogram/force (kg/f) is used. One kg/f = .............. N

2. The SI unit for power is the ................ , abbreviated to ......... . Most light
aircraft engines are however, still rated in ............................. .

3. The units used in aviation are; for altitude-...................... ; for navigational


distance-.................. .. ........... ; and for speed-.................... .

4. A scalar quantity has only ............................ . A vector quantity has


both .............................. and .......................... .

5. Velocity is speed in a given ............................. .

6. Acceleration occurs whenever there is a change in speed and/or direction;


i.e. whenever there is a change in ......................... .

7. Mass is measured in ........................... A common definition of mass is that it


is the .......................... of ...................... in an object.

8. The tendency of an object to continue moving at a certain velocity, or


remain at rest, is called ........................... .

9. Large masses have high ........................ , smaller masses have less ................ ..

10. A force is something which has the potential to cause


an ................................... .

II. Newton's First Law states 'Every object continues in its state of ............... or
of uniform ................... in a straight line unless acted upon by an
external ................ .'

12. Newton's Second Law can be written as: Force = ............ x ................... .

13. Weight is the ................... produced when an object (or mass) is acted upon
by ............................... attraction.

14. g is the normal acceleration of gravity. It causes an acceleration


of ................ m/s 2 on all masses on or near to the surface of the earth.

15. 'g' is another term given to any acceleration. A person subjected to 3 'g' will
feel ............... times heavier than his or her normal weight.

16. The resistance of a moving object to being turned, slowed down or stopped
is called its ........................... , measured by the product of its .......... ..
X ..................... .

I 7. Newton's Third Law states that for every ................... there is an equal and
opposite ......................... .

18. For an object to follow a cu1ved path, a force must be applied .................. the
centre of the curve. This force is called ................................... force.

19. The formula for measuring the strength of this force toward the centre of the
curve is given by CPF = -- .

Principles of Flight Aeroscience 1-15


20. Sketch in the resultant of these two forces.
/
21. Resolve this single vector into two components, and draw them in. /
~a
22. What is the cosine of the angle A in this diagram?

c
force
pivot point
23. What is the moment of the force in this distance 500mm
diagram-in kg-mm?

24. What is the moment of the couple in the 4kg


following diagram?
10N
force
distance 3m

I
pivot
force

10N

25. If the forces acting on an object are in equilibrium, this means that it is:
a. at rest;
b. either at rest or moving with a constant velocity;
c. not accelerating;
d. accelerating.

26. Are the forces acting on the object in the following diagram in equilibrium?
If not, sketch in a vector representing the net force acting on the object.

27. Work is done when a force moves an object through a ........................ .

28. Power is the ........... at which work is done; i.e.


Power = = -----

29. Kinetic energy (KE) is the energy due to ....................... It is measured by the
formula KE = ..................... .

30. An aircraft at altitude has gravitational ............................. energy, which can


be converted to ............................ energy and used to maintain
flying .................... in a descent.

1-16 Aeroscience The Commercial Pilot Series


The Atmosphere
Introduction

To become a capable pilot, you must have an understanding of the aerodynamic


forces which act on your aircraft when it moves through the air. A first step
toward this understanding of aerodynamics is a good appreciation of the
physical properties of the medium in which we fly-the atmosphere. In this
chapter, we are concerned with the inter-related atmospheric properties of
density, pressure and temperature. Other properties such as humidity and
viscosity will also be covered briefly.

Density
The air in the earth's atmosphere is comprised of a mixture of gases, principally
nitrogen (78%) and oxygen (21%). The proportion of the individual gases which
constitute air is not important aerodynamically-all have equal value for the way
in which they add to the density of the air. However, oxygen indirectly has
added significance in aviation for, of all the gases in air, it alone sustains the
power produced by the engine, and the life of the pilot. Both are vital to the
success of powered and manned flight!

Density is the term used to represent the number of molecules and how closely
they are packed into any given volume of a substance. The density of a given
solid or liquid remains constant for all practical purposes. In contrast, the
density of any gas, air included, is variable. In the atmosphere near to the earth's
surface, there are many, many millions of molecules of air in (say) just one cubic
metre. At altitude (as we will discuss in more detail shortly), there are fewer
molecules of air in any given volume and it is therefore less dense.

dense air less dense air

(many millions {fewer molecules per


of molecules per cubic metre)
cubic metre)

at sea level - about 1.225 kg mass on top of Mt. Cook- about 0.85 kg mass

Fig. 2-1. Air density.

Density is a property which is difficult if not impossible to measure directly.


However, the molecules of any substance have a known, finite mass, and in
physics the usual way of measuring density is by calculating the total mass
contained in a given volume. The density of air at sea level is calculated as
being about 1·225 kg/cubic metre. By comparison, the density of fresh water is
1000 ]{g/ cubic metre.

The density of the surrounding air is the most important determinant of


how well an aircraft and its engine can perform. As an aircraft climbs, the
density of the air in which it is flying decreases and the wings become
increasingly less able to generate the lift required. At the same time, the power
output of the engine reduces (from sea level if the engine is normally

Principles of Flight The Atmosphere 2-1


aspirated-from some higher altitude if it is supercharged). Ultimately, an
altitude will be reached where the density of the air becomes too low to sustain
airborne flight.

Before flight, we often need to be able to predict how well our aircraft and its
engine(s) will perform. To do that we need to know the prevailing air density.
However, as already discussed, density is practically impossible to measure
directly. In addition, the scientific method of measuring and comparing different
densities in terms of mass/cubic volume is unsuited to practical application in
aviation. As aircrew, we need to have a much more immediate method of
measuring air density. Fortunately, there are two related properties of the air
which are easy to measure, and which, when considered together, provide a
good indication of the prevailing air density. They are pressure and
temperature.

Pressure
In a given parcel of air there are countless millions of molecules. Each of these
molecules is in continuous motion-vibrating and moving about at high speed in
a random fashion. The situation can be compared to that of a container full of
tiny fast-moving tennis balls, continuously bouncing back and forth. When they
come close enough to a solid surface to collide with it, each molecule exerts a
tiny pressure on that surface. The sum of the forces exerted by many millions of
molecules constantly colliding with and rebounding from a given surface can
add up to a total force which is quite considerable.

Millions of air in still air,


molecules in pressure acts
rapid and equally in all
random motion directions

t
results in many
millions of
collisions per
_ _ _,.
..... 1 /+ ',\: which adds
up to a
considerable
second \
' '-
___ ,......,
/ I
I \
' '- ___ ,.,... I
/
I force called
pressure

Fig. 2-2. Air pressure.

The total force acting on a given area is called pressure, which, in the Sl system
of units, is measured in newtons/sq. metre or Pascals. (One Pascal = I newton/
sq. metre or approximately 0·102 kg/f/sq. metre). The Pascal is a fairly small
unit, so, for convenience in aviation spheres, atmospheric pressure is normally
measured in hectoPascals (hPa), i.e. hundreds of Pascals.

2-2 The Atmosphere The Commercial Pilot Series


The amount of pressure exerted by air on a solid surface depends in the first
instance on its density. The more dense the air, the more molecules there are
to collide with the surface and to generate pressure as a result. In two parcels
of air with the same temperature, the one which has greatest density, has the
higher pressure. Hence, pressure is a good indicator of density, but to more
accurately determine the density of air we must also take its temperature into
account.

dense air less dense air


more molecules , fewer molecules
higher pressure lower pressure

Fig. 2-3. A high air pressure indicates a high density.

Temperature
Temperature is a measure of how much energy the molecules of a substance
have-of how fast they vibrate. If heat is transferred to air, its molecules will
become more 'excited', its temperature will rise and, with more kinetic energy,
the force with which they collide with a solid surface will increase. Hence, if the
heated air is confined in a container and is unable to expand, its pressure will
rise even though its density (the number of molecules contained) will have
remained constant. Conversely, if heat is lost from the air in the container, its
pressure will fall.

Fig. 2-4a. The relationship


between temperature and
pressure.

0
Air heated in a container, unable to expand
pressure rises,
density remains constant
If however, the heated air is not contained, its molecules-given greater kinetic
energy-will move further apart and it will expand. Fewer molecules will be
present in the original volume and its density will decrease. Under these
conditions, the 'parcel' of heated air will continue to expand until its pressure is
equal to that of the surrounding air. Conversely, if a parcel of uncontained air is
cooled it will contract and its density will increase.

Principles of Flight The Atmosphere 2-3


Fig. 2-4b. The relationship
between temperature and density.

0
Uncontained air expands when heated
density decreases

To summarise thus far, the density of air is related to both its pressure and
temperature such that:
o In different masses of air with the same temperature, that with the highest
pressure has the highest density.
o In different masses of air with the same pressure, that with the lowest
temperature has the highest density.

Density, Pressure and Temperature in the Atmosphere


The atmosphere is a blanket of air surrounding the earth which is held in place
by gravity. Although there is no strictly defined outer limit, it is generally
accepted that the atmosphere extends to about 500 miles from the earth's
surface. The weight of the air in the outermost reaches bears down on the air
nearer the surface and compresses it. As a result the air nearest the surface is
the most dense.

The atmosphere is recognised as having several layers where different


conditions exist. In this discussion we will be confining ourselves to the lower
part of the atmosphere, i.e.-below an altitude of about 60,000 ft,-where all
commercial flying takes place. This part of the atmosphere is comprised of the
troposphere and (usually) the lower part of the stratosphere.

In the troposphere-the first layer of the atmosphere-pressure, density, and


(normally) temperature, all decrease with altitude. Note that:
• About 75% of the total mass of the atmosphere is contained in the
troposphere. That is, below an average altitude of some 36,000 ft. By
comparison with their sea-level values, pressure and density are generally
reduced respectively to:
-7596 % by 8- 10,000 ft
-50% 1/2 by 18- 22,000 ft
-25o/o /4 by 34- 41 ,QQQ ft
1

• As already mentioned, in the troposphere, temperature also normally


decreases with altitude. Occasionally however, a layer of air may be
encountered where the temperature remains constant (in an isothermal
layer) or increases with altitude (in an inversion layer). These layers can be
several thousand feet in depth.

At an altitude called the tropopause, the temperature stops decreasing and


becomes isothermal in a layer which extends well up into the stratosphere. The

2-4 The Atmosphere The Commercial Pilot Series


tropopause forms the 'boundary' between the troposphere and the stratosphere. It
is not always clearly defined but it occurs at an altitude of about 30,000 ft over the
poles increasing to a maximum of about 60,000 ft over the equator. Above the
tropopause, pressure and density continue to decrease with altitude.
/'\
~-\
to about 50 km
1\
I \
' I~
the I I
up to 60,000 ft I .
stratosphere (equator) pressure and density decrease
temperature remains constant
6
tropopause -
I '
at about
- 3_± ::__4.!.., QQ Q_ft_

v
about 30,000 ft
(poles)
pressure and density
reduced to % S.L. value.

at about
pressure,
temperature (usually), _1§::...2~.QQQ_f!_ -
pressure and density
and density all decrease reduced to %S.L. value

at about
8- 10,000 ft
--------
pressure and density
reduced to 3.4 S.L. value

Fig. 2-5. Features of the atmosphere.

(For a more detailed explanation of the atmosphere, see our companion book
"Meteorology for Professional Pilots" which is a part of this series).

Owing to different amounts of solar radiation received at the surface, atmospheric


conditions around the earth are not uniform and are continually changing. The
pressure, temperature and density conditions of equatorial, polar and mid-latitude
air masses are quite different and they change with the seasons and as different
weather systems move around the globe. To provide a yardstick against which the
continually varying atmospheric conditions can be compared, an International
Standard Atmosphere (!SA) has been devised. The ISA also provides the means
for calibrating and setting altimeters, and of predicting aircraft performance.

The !SA is a hypothetical set of atmospheric conditions which represents an


average of the conditions experienced world-wide. Among other things, !SA
conditions assume:
• a sea level pressure of 1013-2 hPa, with pressure decreasing with altitude at
rapid rate initially (about 1 hPa every 30 ft to 5000ft), and at a lower rate
thereafter.
• a sea level temperature of + 15'C, decreasing at a uniform rate of 1·98'C/1 OOOft
to -56·4' Cat the tropopause (set in the lSA at 36,090 ft).
• a sea level density of 1·225 kg/m 3 , decreasing with altitude (like pressure) at a
rapid rate initially, and at a lower rate thereafter.

Principles of Flight The Atmosphere 2-5


_,....
,r'', _.,.,'',,_
....
........ ~
::. ... ......:
-:::. ...
'' '' ''
''' ''' '''
reduces :' :' reduces:' '~ constant:'

3
'' '' ''
~ogo--'-"'--=""0 --~-- tropopause--__:_ '
~--~""'- ~' -

Fig. 2-6.
The International Standard
Atmosphere.
slower slower steady
reduction reduction reduction
(1 ·98°C/
1OOOft)

rapid
reduction

The actual conditions existing at any one time and at any one place on the globe
will very rarely exactly match those of the !SA. The !SA is nevertheless an
important benchmark. Whenever we fly, we should be aware of how much the
existing meteorological conditions vary from the !SA 'standard', for this will give
an indication of the performance we can expect from our aircraft.

Remembering that it is the density of the air which matters we must therefore
take account of the way in which it changes as the temperature and pressure
vary with different weather conditions. Atmospheric density and hence,
performance, will vary under the influence of pressure and temperature as
follows:
• A high temperature will indicate low density and poorer performance .
• A low temperature will indicate high density and better performance .
• A high pressure will indicate high density and better performance .
• A low pressure will indicate low density and poorer performance .
Putting temperature and pressure effects together:
• High pressure + low temperature = best density and best performance
conditions.
• Low pressure + high temperature = worst density and worst
performance conditions. ·
As can be seen from the above, the temperature and pressure effects on density
can, in some circumstances, counteract one another. For example, a high
pressure tends to increase air density but if, at the same time, the temperature of
the air is high, this will tend to lower its density. Just what the end result will be,
depends on which effect is the stronger-and for that you will normally need to
have performance charts, or be able to work out what the prevailing density
altitude is.

2-6 The Atmosphere The Commercial Pilot Series


Density Altitude
Density altitude is the altitude in the ISA which has the same air density as the
actual altitude. If actual conditions are the same as ISA, (a rare event), then
density altitude will be the same as actual altitude. When actual atmospheric
pressure and temperature conditions differ from ISA, (as they normally do), then
density altitude will be some higher or lower figure than actual altitude, by an
amount depending on how much the prevailing pressure and/or temperature
conditions differ from ISA.

Density altitude is a theoretical altitude which is used as a yardstick for


measuring performance. Aircraft performance charts (or graphs), are based on
ISA (theoretical) conditions, and if we wish to know how well an aircraft will
perform under actual conditions, we must know the prevailing (or forecast)
temperature and pressure at the altitude we are concerned with, and either;

enter these in the charts to obtain density altitude: or


• use the navigation computer to work it out: or
• do a manual calculation.

For most aircraft operations, the calculation of density altitude is normally


incorporated in the appropriate performance charts. The methods of calculating
density altitude are covered later in chapter 17 Performance.

Humidity
The hypotheticallSA is based on d1y air-i.e. air which contains no water vapour
(water in the form of a gas). However, the real atmosphere almost always
contains some water vapour, with relative humidity usually ranging between 20 -
80%. When water vapour mixes with air, some of the molecules of the other
gases are displaced. Water vapour molecules have less mass than the
molecules of the 'd1y' air they have displaced, hence, at the same temperature
and pressure, moist air is less dense than dry air. In normal conditions, the
amount of water vapour in the air does not make a significant difference to its
density. However, as warm air can contain much more water vapour than cold
air, at high temperatures and high humidity, the reduction in density may
become significant.

Viscosity
Viscosity is the tendency for the molecules of a fluid (a gas or a liquid) to 'stick
together' and to cling to any solid surface they come in contact with. Viscosity is
patently obvious in thick fluids like treacle or heavy oil; less obvious in thinner
fluids like water. It may be hard to imagine air, thin as it is, having any
viscosity-but it does. Viscosity in air plays an important part in aerodynamics
and that is the reason for mentioning it here. When air moves across a perfectly
flat plane surface, the layer of molecules in immediate contact with that surface
will be brought to a standstill (or nearly so) because of its viscosity (or
'stickiness). Successive layers of air above that very bottom layer will be slowed
down by decreasing amounts, until the point is reached where the effect of
viscosity is no longer felt. This layer of retarded air is called the boundaty layer-
it is generally not very thick, usually not much more than a millimetre or so,
ranging up to 20mm depending on conditions. We will be discussing the effects
of viscosity, or skin friction, in more detail later.

Principles of Flight The Atmosphere 2-7


Review 2

I. The density (or mass) of air at sea level is about ................ kg/cubic metre. By
contrast, the density of fresh water is ................. kg/cubic metre.

2. Air density, which is vital to aerodynamic and engine performance, is difficult


to measure directly. Fortunately, there are two other properties of air which
give a good indication of air density and are easy to measure. They
are ....................... and .............................. .

3. In aviation, atmospheric pressure is normally measured in .............................. ,


abbreviated to ................ .

4. If atmospheric temperature is normal, but the pressure is high, you would


expect the air density to be (higher/lower) than normal.

5. If atmospheric pressure is about normal, but the temperature is high, you


would expect the air density to be (higher/lower) than normal.

6. The International Standard Atmosphere (!SA) sea-level:

a. pressure is .................... hPa.;

b. temperature is ............. 'C.

7. !SA conditions assume:

a. a pressure lapse rate of ........ hPa for every ......... ft gain in altitude from
sea-level up to about 5,000 ft;

b. a uniform temperature lapse rate of... ..... 'C (or approximately....... 'C)
up to the tropopause which is assumed to be at about... ................ feet.

8. Atmospheric air density reduces at:

a. a steady rate up to the tropopause;

b. at a rapid rate initially, and then at a reducing rate as altitude is gained;

c at a rate which is solely dependent on pressure.

9. By comparison with sea-level values, pressure and density in the atmosphere


are generally reduced:

a. to ............ % by 8- 10,000 ft;

b. to ............ % by 18-22,000 ft;

c. to ............ % by 34-41,000 ft.

I 0. In actual atmospheric conditions, the altitude of the tropopause varies


between about ................ .ft over the poles to a maximum of
about .................... ft over the equator.

2-8 The Atmosphere The Commercial Pilot Series


II. By comparison with !SA conditions, which are used to rate aerodynamic and
engine performance:

a. a high temperature will indicate (high/low) density, and therefore


(better/poorer) performance;

b. a low temperature will indicate (high/low) density, and therefore


(better/poorer) performance;

c. a high pressure will indicate (high/low) density, and therefore


(betteiipoorer) performance;

d. a low pressure will indicate (high/low) density, and therefore


(better/poorer) performance.

12. The best density (and therefore performance) conditions are those with a
(high/low) pressure and (high/low) temperature.

13. The worst density (and therefore performance) conditions are those with a
(high/low) pressure and (high/low) temperature.

14. Density Altitude is the altitude in the ............ which has the same density as
the actual altitude.

15. Two locations have the same atmospheric temperature and pressure. At
one, however, the air is very dry, while at the other, humidity is high. Aircraft
performance will be better at the (dry/humid) location.

Principles of Flight The Atmosphere 2-9


2-10 The Atmosphere The Commercial Pilot Series
Basic Aerodynamic Theory
Static Pressure
The prevailing pressure at any point in the atmosphere is called static pressure.
This is the pressure which results from the weight of all the air above that point
bearing down. As its name implies, static pressure does not involve movement.
!f a solid body is immersed in the atmosphere and has no movement relative to
the air around it, static pressure will be experienced equally on all of its surfaces.

A term often used in aerodynamics for the prevailing atmospheric pressure is


freestream static pressure with the symbol p 00 used to denote it. The word
'freestream' indicates the air conditions which exist well ahead of a body
moving through the air, as yet unaffected by its passage. As we saw in Chapter
2, the freestream static pressure decreases with altitude.

Fig. 3-1. When there is no


movement between a body
and the air around it, pressure
is experienced equally on all
surfaces.

Dynamic Pressure
When there is movement between a solid body and the air surrounding it, the
distribution of pressure around the body will no longer be even. The surface
pressures experienced in those areas facing into the airstream are increased
above the freestream static value, while the pressures to the side and to the rear
of the body are generally reduced. These differences in the pressure
experienced are related to the kinetic (or dynamic) energy the air has by virtue
of its movement, and the extra pressure (called dynamic pressure) which it is
capable of exerting as a result.

Fig. 3-2. When there is air


movement between a body and
the surrounding air, the movement
distribution of pressure around
the body is uneven.

Principles of Flight Basic Aerodynamic Theory 3-1


Any solid body which is moving has kinetic energy. The amount of energy which
it has is calculated by:
kinetic energy = 1/2 m V2
Air also has kinetic energy when it is moving-normally referred to in
aerodynamics as dynamic energy. The mass of air is measured by its density-
i.e. its mass per unit volume. The symbol for density is the Greek letter p,
pronounced RHO. If we substitute density for mass in the above equation, we
can calculate the amount of dynamic energy in a moving mass of air as:
dynamic energy= 1/2 pV'
Where p= freestream air density
and V = velocity of the airstream

If this moving mass of air is stopped by a solid body and brought completely to
rest, the dynamic energy which it contains is converted to pressure energy. The
pressure which alises for this reason is called dynamic pressure and it is exerted
on the body in addition to the ambient, or prevailing static pressure. It is also
measured by:
dynamic pressure = /z pV'
1

The term 1/z pV' thus stands for 'the additional pressure imposed when air of
a certain density moving at a given velocity is brought completely to rest'. It
is also used in a more general way to describe the amount of dynamic (or
kinetic) energy contained in a moving airstream.

Vety little of the air moving past an aircraft in flight is 'brought completely to rest'.
As will be discussed shortly, the actual pressure distribution around wings (or
other parts) of an aircraft also depends on other factors including size, shape,
and orientation in the airstream. The term for dynamic pressure, (V, pV'), is
nevertheless very important-all aerodynamic forces are proportional to it. You
will find that the term (spoken as half-rho-vee-squared) is used a lot in
aerodynamics. In some references, you will see it represented in equations by the
symbolq.

Measurement of Airspeed
Dynamic pressure is utilised in the measurement of airspeed. A small amount of
the air moving past the aircraft is brought to rest in a fmward-facing tube called
the pitot tube. The pressure which is present inside the tube is called total (or
pilot) pressure and it complises the dynamic pressure caused by blinging the
moving air to rest, plus the freestream static pressure, i.e:

total (or pitot) pressure = freestream static pressure + dynamic pressure

= Pw+ V, pV'
Static pressure is tapped from another point on the aircraft where, from testing, it
has been determined that the freestream static pressure is least affected by the
passage of the aircraft through the air.

As shown in Fig. 3-3, these two different pressures (total and static) are fed
separately to each side of a diaphragm in the airspeed indicator. By a suitable
system of gearing, any movement of the diaphragm is indicated by the position
of a pointer on the face of the instrument. Hence, within the instrument, total
pressure is compared with static pressure such that a reading for dynamic
pressure (calibrated as airspeed) is obtained, i.e:

3-2 Basic Aerodynamic Theory The Commercial Pilot Series


total pressure (dynamic + static)- static pressure = dynamic pressure
which can be written as ( Poo + \1, pV
2) - Poo = Yz pV2

dynamic
Fig. 3-3. +
static
Airspeed indicator:
principle of operation.

Hence an airspeed indicator (ASI) is simply a dynamic pressure gauge which is


calibrated to read in airspeed (knots). (A more detailed coverage of the
functioning of airspeed indicators is given in our General Aircraft Technical
Knowledge manual which is a part of this series.)

The Relationship Between Airspeeds

Airspeed indicators are calibrated at lSA (or 'standard') sea-level density


conditions. It is only under these conditions that the airspeed indicator will
accurately indicate the actual speed-that is, the true airspeed {TAS)---of the
aircraft through the air. When the ambient (freestream) air density differs from
standard sea-level conditions, the indicated airspeed will be different from the
actual airspeed. This difference (sometimes called density error) and other
errors which arise in the airspeed indication system has given rise to a number
of different terms for airspeed, as follows:
• Indicated Airspeed (lAS). The reading on the airspeed indicator (ASI).
There may be some difference between the indications registered by
individual AS!s, but these instrument errors are usually so small that for all
practical purposes, they can be ignored.
• Calibrated Airspeed (CAS). lAS which has been corrected for pressure {or
position) error. Pressure error arises mainly from incorrectly sensing total
and static pressure when the aircraft is in different flight attitudes. When
pressure error correction (PEC) cards are displayed in the cockpit, they
include the instrument error conection for the particular AS!. In practice, the
PEC at normal cruise speeds is also small and can under most
circumstances be ignored.
• Equivalent Airspeed (EAS). Most AS!s are designed to measure dynamic
pressure (V,p\12) on the assumption that air is incompressible. Air is
compressible, and as speed is increased the pilot tube will increasingly
register a higher pressure than it should because the air becomes
compressed. When CAS is corrected for this compressibility error it
becomes known as EAS.
• True Airspeed (TAS). TAS may be obtained by dividing the EAS by the
square root of the relative density-i.e. the prevailing density as compared
with the standard sea-level density. If the prevailing density at sea level
happened to be standard-the same as in the !SA-the relative density
would be I and EAS would be equal to TAS. Flying at 40,000 ft under
standard conditions, where the relative density is one-quarter of the sea-level
value, the EAS would be half theTAS (the square root of •;, is 1/z). In practice,
TAS is usually calculated by using a navigational computer.

Principles of Flight Basic Aerodynamic Theory 3-3


For our purposes in this manual, of the foregoing airspeeds which we have
listed, there are two which are important-TAS and lAS.

• TAS is important in aerodynamics because it gives a measure of the actual


speed of a body relative to the freestream flow. The velocity factor, V, in the
dynamic pressure function 1/2pV2 is the true airspeed of the aircraft.

• lAS is important for two reasons. First, because it is the airspeed which is
most readily available to the pilot. Secondly, because the aerodynamic
forces which the pilot has at his disposal to fly and manoeuvre the aircraft
depend on it. At any given TAS, it is the indicated airspeed which gives a
measure of air density and thus how well the aircraft will perform. If at that
same TAS, air density is higher, this will be reflected by a higher lAS and
better performance. Hence we say that lAS is the 'aerodynamic' airspeed.

NOTE: Strictly speaking, EAS is the exact aerodynamic airspeed. However, the
error in assuming that lAS and EAS are the same-in other words that air is
incompressible-is very small. In the low-subsonic speed range (i.e. at speeds
up to about 40% of the speed of sound), this error amounts to less than 5% and is
usually much smaller.

Aerodynamic Force
An aerodynamic force is generated if a solid body is placed in a moving
airstream. This force originates mainly from the pressure distribution around the
body and, to a lesser extent, the friction between the body and the air. You can
experience this force for yourself if, for example, you hold up a large rigid sheet
in a strong wind. If the flat surface of the sheet is held more or less at 90' to the
wind, dynamic pressure is at its maximum and a considerable aerodynamic
force will be generated in the 'downwind' direction. If however, the sheet is
held with its edge facing the wind, the effect of the wind is considerably
reduced. There will still be some aerodynamic force on the sheet, caused by
dynamic pressure on the thin edge and skin friction between the air and the
broader surfaces, but it will be so slight as to be barely noticeable.

If you were then to tilt the into-wind edge of the sheet slightly upwards you
would create an angle of attack. A strong aerodynamic force would quite
suddenly reappear, but this time acting in an upward direction, almost at a right
angle to the wind. In a strong wind, this aerodynamic lifting force can be large
enough to overcome the force of gravity and the sheet is liable to take off.

wind direction ==i>

Fig.3-4. The aerodynamic force on a flat sheet being held in a wind.

It is this lifting capacity of large, relatively flat bodies inclined at a small angle to
an airflow, which enables airborne flight.

3-4 Basic Aerodynamic Theory The Commercial Pilot Series


Aerofoils
The wings and tail section (empennage) of an aircraft, which are designed to
produce useful aerodynamic forces, are called aerofoils. In the early
developmental stages of manned flight the aerofoils which were used were not
very different to the thin sheet we used in the example above. These earliest
aerofoils were of very light construction and were typically curved, with a convex
upper surface and a concave lower surface. The term used to describe this
overall curvature of aerofoil surfaces is camber.

It was discovered during the first World War that thicker cross-sectional
shapes-called sections-gave better lifting characteristics and, therefore, better
manoeuvrability. Another advantage was that greater strength could be built in
to a deeper cross-sectional shape. From about 1916 these thicker sections, with
a flat, or slightly convex lower surface began to be used and, although the
shapes of the general purpose aero foils in use today are similar in section, many
other aerofoil shapes have been developed to meet different requirements.
Broadly speaking, there are three classes of aerofoil section in use today-high-
lift, general purpose, and high-speed-as shown in Fig. 3-5.

high lift pre- 1916

general purpose high speed

Fig.3-5. Various classes of aerofoil section.

The wing sections used on training aircraft are usually of the general purpose
type. These are characterised by a rounded leading edge; a moderate amount
of curvature on the upper surface; and less curvature on the lower surface. The
sections used for the empennage (tailplane and fin) will generally be more
symmetrical, i.e. with equal curvature on both surfaces.

Aerofoil Nomenclature

The nomenclature used to describe aerofoil sections is:


• Leading edge: The edge facing into the airstream.
• Trailing edge: The edge at the 'downstream' side.
• Chord line: The straight line joining the leading and trailing edges.
• Chord: The distance between the leading and trailing edges, measured
along the chord line.
• Thickness: The depth of the aerofoil. On most aerofoils, the point of
maximum thickness is forvvard toward the leading edge, (usually at about 30
- 40% chord-i.e. of the distance back from the leading edge).
• Thickness/chord (tic) ratio: The maximum thickness of the aerofoil
expressed as a percentage of the chord. High lift aerofoils have a t/c ratio in
the order of 15-17%; general purpose aerofoils I 0-12%; high speed aerofoils
7%.

Principles of Flight Basic Aerodynamic Theory 3-5


• Camber: The curvature of a line drawn equidistant between the upper and
lower surfaces (the line of mean camber). Most aerofoils will therefore have
some camber, only symmetrical aerofoils are not cambered.

thickness mean camber line


\ ~
leading
edge
'0"'~~~U~~f~~~~~~-t
~ 4 t trailing
edge
chord line
-----chord-----------

Fig.3-6. Aerofoil nomenclature.

Angle of Attack
The angle of attack is the angle between the chord line and the relative air
flow (RAF). The RAF is that airflow which is remote from the aircraft and is
unaffected by its passage through it-i.e. the freestream airflow. (You should
note this latter point carefully. The airflow ahead of the wing usually changes
direction as the aircraft approaches, particularly at high angles of attack). The
relative airflow can be represented diagrammatically by a vector indicating
velocity and direction. The RAF vector is of the same magnitude but opposite
direction to the aircraft's flight path vector at any given moment.

relative air ftow angle of attack

Fig.3-7. Angle of attack.

NOTES:

1. The Greek letter alpha (a) is often used to denote angle of attack.

2. The angle between the RAF and the chord line as defined above is
sometimes also referred to as the geometric angle of attack.

3. In discussing aerodynamic theory, the subject is conventionally treated as if


the aerofoil is stationary and the air is moving-as it would be, for example, in a
wind-tunnel experiment. In reality, it is the aircraft which moves and the air
which is stationa1y (putting aside any turbulence which may be present in the
airmass). What matters is the relative motion between the solid body and the
air. Whether it is the aerofoil or the air which is moving makes no difference to
the resulting forces.

3-6 Basic Aerodynamic Theory The Commercial Pilot Series


Pressure Distribution
An aerofoil, placed in a moving airstream at a small angle of attack, parts the air
at the leading edge. The air then flows smoothly above and below the upper and
lower surfaces before being joined again at the trailing edge. Static pressure is
decreased in the region above the wing, and altered in the region below the wing.
The net effect of these pressure changes is to produce a force, mostly in the
upward direction but also inclined slightly toward the rear. This aerodynamic
force is called the total reaction (TR).

Most people, if asked to explain the origin of that force, would probably say it was
the result of a wedge of high pressure air being built up under the aerofoil. That
would however, be far from being a complete answer. Over the normal range of
operating angles of attack, the greatest contribution to the TR (and therefore
to lift) comes from a reduction in the static pressure over the upper surface.
Why is that so? For a more detailed explanation, we turn to Daniel Bernoulli
(1700- 82), a Swiss scientist who first described the principle involved.

Bernoulli's Theorem

Bernoulli's theorem states "In the streamline flow of an ideal fluid, the sum of the
energy of position, plus the energy of motion, plus the pressure energy, will
remain constant". We know, of course, the 'energy of motion' to be dynamic
energy. The 'pressure energy' is the amount of static pressure present and, when
applied in aerodynamics to the flow of air around an aerofoil, we can ignore the
'energy of position' (potential energy) because the changes caused to the height
of the air are so small. We can therefore reduce Bernoulli's theorem (or
principle) to:
Dynamic energy + static pressure = a constant; i.e.
Y2 pV2 + Pro = a constant

Air is compressible and has viscosity. For these reasons it is not an ideal fluid.
Bernoulli's principle can however be applied with a good degree of accuracy in
streamline airflows with a velocity of up to about 250 knots. In higher velocity
airflows the effects of compressibility and viscosity have to be increasingly taken
into account until, at speeds approaching the speed of sound, the Bernoulli
principle is no longer appropriate. (These effects are discussed in a later
chapter.)

Hence, we can apply Bernoulli's principle to the streamline airflows around an


aerofoil at the lower subsonic speeds and state with confidence that:
• wherever the speed of the airflow is increased, the air gains dynamic
energy and its static pressure is accordingly reduced, and conversely;
• wherever the speed of the airflow is decreased, the air loses dynamic
energy and its static pressure is increased-

Streamline Flow

If succeeding particles of air in an airstream follow the same steady and


predictable path, then this path can be represented by a streamline. There will be
no flow across the streamlines, only along them. Streamline flow can be
maintained provided the air particles flowing around or through a body can
change direction gradually and smoothly, and shapes designed to achieve this are
said to be streamlined. If the airflow is required to change its direction too
abruptly then the flow will 'break down' and become turbulent and unpredictable.
Bernoulli's principle applies only to streamline flow.

Principles of Flight Basic Aerodynamic Theory 3-7


Venturi Effect
A good example of the Bernoulli principle in
operation can be seen in venturi effect. A
venturi is a convergent-divergent duct. The
cross-sectional area of a venturi decreases
smoothly until a 'throat' is reached about one-
quarter of the way back from the entrance. It
then increases smoothly and more gradually Fig.3-8. A venturi tube.
toward the exit.

Placed in a steady stream of air, a properly designed venturi enables the volume
of air flowing into it over any given time span, to accelerate smoothly and pass
through the restriction of the throat in the same amount of time. Once it has
passed the throat, the air then decelerates to pass out through the exit at the
same speed as it entered. Wherever the air has a higher speed than the free
stream flow, it has gained dynamic energy and, accordingly, static pressure
is reduced. The greatest reduction in static pressure is experienced at the
throat of the venturi, where the increase in the speed of the flow is highest.

increasing speed decreasing speed


decreasing pressure increasing pressure
D - - - - - - ~----------- - , ' - - - - - +
D------
__
- - - - - - -.-.-11!-~-~~-~--:;---
..;..~---
.....
......
_ _ . J - 1 - - - ....

Fig.3-9.
Airflow through a venturi. ====.!--===~===:
---------1--1---·
====2-==-==~===!
D-------- -----·-~-~-~-;-~:;_;~~-~~
»--------~----------------+
lowest static pressure

Note that streamlines drawn for the flow through a ventmi indicate what is
happening to the static pressure. Where they converge, this indicates a lowering
of static pressure; where they diverge, this indicates static pressure is increasing
again. Where they are closest together, the reduction of static pressure will be
greatest. Converging streamlines do not indicate the air is being compressed. If
sufficient time is given for the flow to speed up and slow down, air resists being
compressed. At flow velocities in the low-subsonic range, there is 'sufficient
time' and the amount of compression which does occur is insignificant.

Airllow Around an Aerofoil


A streamline airflow around an aerofoil behaves in the same way as the flow
through a venturi.

reduced static pressure


~------------------1------------------~

Fig.3-10.
Airflow over an aerofoil
shape. ~11~1~~~~li"<>-;z~')"l"'>c,'.l_))))ii~
~-------------------------------------~
»- ___________ Jr~e~tc.ea01_SJC!.t~_pre..s§Y.r~- ________ _...
~--------------------- ---------------~

3-8 Basic Aerodynamic Theory The Commercial Pilot Series


Air moving over the aerofoil shape depicted in Fig. 3-10 must accelerate to pass
over the top surface. It therefore gains dynamic energy, and the static pressure in
that region is decreased. In effect, the aerofoil is acting as a 'half venturi', with the
air being forced to flow through a 'throat' comprised of the upper surface of the
aerofoil and more remote air above it. At the same time, air passing below the
aerofoil is not deviated from its path-there is no change in velocity and the static
pressure remains the same as for the free stream flow. With the pressure
distribution which exists under these circumstances, (lower above the aerofoil
than below it), a small force is generated (TR) which tends to move the aerofoil
toward the lower pressure area.

(This example se1ves to illustrate the fact that lift can be generated at zero angle of
attack, provided the curvature of the upper surface of the aerofoil is greater than
that of the lower surface. In the real situation however, the amount of lift normally
available at zero angle of attack is small, unless the aircraft is travelling at very high
speed!)

The following diagrams illustrate the effect of low-subsonic airflow around an


aerofoil section at various angles of attack. The section depicted is representative
of a general purpose (GP) aerofoil with streamlines drawn from those observed
when smoke streamers were generated in wind-tunnel tests.

The GP aerofoil sections used on general aviation aircraft are many and varied. In
general terms however, they will typically have a rounded leading edge; some
curvature of the lower surface; and a moderate amount of camber. The normal
operating angles of attack are usually between about 2" and 15". The following four
diagrams represent the airflows which may be expected at low, moderate, and
high angles of attack within this range, and at a little beyond the stalling angle. In
each case, the direction of the freestream relative airflow is shown so that the
direction of the streamlines can be compared with it. The areas in which the
pressure is lower than freestream static are depicted with a lighter shading and a
(-) symbol. Areas of higher pressure have a darker shading and a ( +) symbol.
These areas are sometimes referred to collectively as the 'pressure envelope'.

Fig. 3-11a.

a =4"

At low angles of attack (Fig. 3-11a), as the streamlines indicate, there is relatively
little disturbance of the airflow past the aerofoil. Ahead of, and slightly below the
leading edge, there is an area of higher pressure. Within this area, there will be a
point on the leading edge called the stagnation point where the flow is brought
completely to rest. There is another smaller area of slightly raised pressure around
the trailing edge. In accordance with Bernoulli's principle, there are areas of lower
pressure above and below the aerofoil with the upper area being more extensive.

Principles of Flight Basic Aerodynamic Theory 3-9


Fig. 3-11b.

Fig 3-11c.

As the angle of attack is increased (Figs. 3-11 b and 3-11 c) the airflow must
increasingly deviate from its path and accelerate to follow the contour of the
upper surface-particularly over the fmward part. As a result, the upper area of
low pressure moves forward. By the time an angle of attack of about I 0" is
reached, the area of lower pressure under the aerofoil has disappeared. At
higher angles, the area of high pressure forward of the leading edge spreads
toward the rear until it eventually covers the whole of the lower surface.

Fig. 3-11d.

a= 18"

3-10 Basic Aerodynamic Theory The Commercial Pilot Series


Beyond an angle of attack of about 15°, the change in direction around the leading
edge and folWard upper surface becomes too abrupt and the airflow can no longer
conform. The airflow separates from most of the upper surface and the turbulent
wake behind the aerofoil becomes greatly enlarged. The angle at which this occurs
is called the critical (or stalling) angle. When it is exceeded, the low pressure
envelope over the upper surface virtually collapses and becomes unpredictable as
indicated in Fig. 3-11 d. Pressures below the aerofoil continue to increase with
angle of attack as more of the lower surface is presented toward the oncoming
airflow.

NOTE: The magnitude of the pressure changes occurring within the pressure
envelope should not be exaggerated. In the low-subsonic speed range with which
we are mainly concerned, the changes of pressure around the aerofoil are for the
most part not more than one or two percent of the freestream static value.

Upwash

The streamlines indicate that the airflow turns upward ahead of the aerofoil. This
phenomenon is called upwash and it becomes more pronounced as angle of attack
increases and there is, as a result, a greater pressure differential between the upper
and lower surfaces of the aerofoil. Upwash is generated by small pressure
disturbances transmitted ahead of the aerofoil at the speed of sound, which cause
the air particles to move toward the area of lowest pressure. In effect, the air is able
to sense the approach of an object through it and begin moving to take the path of
least resistance.

Downwash

As the airflow passes the aerofoil, it is turned downward with respect to the
freestream direction. This movement of the affected air is called downwash and it
extends for some distance behind the aerofoil. Downwash is an inevitable
consequence of lift production-a mass of air must be moved in a given direction to
produce a lift force in the opposite direction. When an aerofoil is producing lift,
downwash must therefore exceed upwash. As the preceding diagrams indicate,
within the normal operating range, downwash increases with angle of attack. We
will be considering the effects of downwash in more detail in later chapters.

Centre of Pressure (CP)


The pressures existing at various points around an aerofoil can be measured and
compared with the freestream static pressure. The magnitude of the measured
differences in pressure from freestream static may then be represented by vectors
drawn normally (at a right angle) from the aerofoil surface at those points. Vector
directions away from the surface indicate lower pressure, while those toward the
surface indicate higher pressure. Such a vector diagram is shown at Fig. 3-12.

resultant
(not to same scale)

centre of pressure
Fig. 3-12. Vector diagram for an aerofoi/.

Principles of Flight Basic Aerodynamic Theory 3-11


The individual vectors can be combined into a single resultant which indicates
the magnitude and direction of the aerodynamic force on the aerofoil under
different conditions. This resultant is of course the Total Reaction (TR) and it
acts through a point within the aerofoil called the centre of pressure (CP). The
changes and magnitude and direction of the TR with angle of attack in a
constant-speed airflow are shown at Fig. 3-13. Note that:
• the TR force increases with angle of attack and becomes more tilted toward
the rear.
• as the stalling angle of attack is passed, the TR force suddenly reduces and
becomes more markedly tilted toward the rear.
TR
Fig. 3-13. Variation in TR with
angle of attack.
TR

20' (stalled)
Movement of the CP with Angle of Attack

As can also be seen in Fig. 3-13, with a cambered (i.e. non-symmetrical) aerofoil,
the centre of pressure (CP) moves gradually forward with angle of attack. At low
angles of attack, the CP is located at a point some 30 - 40% chord ( i.e. 30 - 40 %
of the distance back from the leading edge). As the stalling angle is reached, the
CP will have moved forward to be located as far forward as 15 - 20% chord. As
the stalling angle is passed, the CP moves rapidly rearward. A graph of typical
movement of the CP for a cambered aerofoil is given in Fig. 3-14.

20

15

angle location
of 10 ofCP
Fig. 3-14. attack
Movement of CP
with angle of attack.

0 10 20 30 40 50 100

lt;l I I I I I IJ_i
percent chord
NOTE: The figures quoted in the preceding paragraph and the graph above must
be regarded as a guide only, as the actual location and movement of the CP with
angle of attack depends to a large extent on the amount of camber and the
specific shape of the aerofoil. The movement described is however generally
true for a cambered aerofoil. With symmetrical aerofoils-such as may be used
for the tail section of an aircraft-there is virtually no movement of the CP over
the normal operating range of angle of attack and airspeed.

3-12 Basic Aerodynamic Theory The Commercial Pilot Series


It is nmmal and more convenient to divide the TR into two component forces
and consider them separately. Those two components are lift and drag. In
flight, the lift force is used to oppose the weight of the aircraft and to provide the
means of manoeuvring, while the drag force is the air resistance which opposes
the aircraft's motion. By definition:
• Uft is the component of the total reaction (TR) at a right angle
(perpendicular) to the relative air flow.
• Drag is the component of TR parallel to the relative airflow and opposing
motion.
TR
1
LIFT I
Fig. 3-15.
The total aerodynamic reaction centre
I
is resolved into two forces -
lift and drag.
of pressure ,
I

relative airflow

Note that the total reaction force can never be at right angles to the relative
airflow. If that were the case, it would mean that lift was being generated
without any drag-the aerodynamic equivalent to perpetual motion. There must
always be some drag if lift is being generated, hence the TR force is always tilted
back at an angle to the relative airflow-albeit at a small angle when the aerofoil
is operating efficiently.

Principles of Flight Basic Aerodynamic Theory 3-13


Review3
I. The prevailing pressure at any point in the atmosphere is called ................. .
pressure.

2. When moving air is brought completely to rest, an additional pressure


called ........................ pressure is imposed.

3. The atmospheric conditions existing well ahead of a moving body and as yet
unaffected by its passage, are denoted by the term .............................. .

4. The term V,pV2 stands for ............................. pressure, where p


= ............................................... andY= .............................. .

5. Pitot (or total) pressure = ......................... Pressure + ....................... pressure.

6. The airspeed indicator is simply a pressure gauge which


measures .................... pressure, but is calibrated to read in ......................... .

7. Equivalent airspeed (EAS) is the exact measure of dynamic pressure ('/, p


V2). In flight below about 250 kts, for practical purposes EAS can be taken to
be the same as ............. , given that instrument, position and compressibility
errors are usually small in that speed range.

8. All aerodynamic forces are proportional to EAS and therefore to ................. in


low-subsonic speed flight

9. The wings and tail section of an aircraft which are designed to produce
useful aerodynamic forces, are called .................................. .

10. Draw a typical cambered GP aerofoil section and annotate the trailing and
leading edges, chord line, line of mean camber, and point of maximum
thickness.

11. The relative air flow (RAF) vector (is/ is not) equal and opposite to the
aircraft's flight path vector at any given moment.

12. Sketch a cambered aerofoil and show the relationship between the chord
line, RAF and the angle of attack.

13. When applied to an airflow around an aerofoil shape, Bernoulli's Theorem


can be reduced to dynamic energy + static pressure = a ........................... .

14. In accordance with Bernoulli's Theorem:

a. wherever the speed of an airflow is increased, static pressure will be


(increased/decreased);

b. wherever the speed of an airflow is decreased, static pressure will be


(increased/decreased).

15. A steady flow of air around a body, where successive particles of air follow
the same smooth path, is called ................................. flow.

16. A disturbed, unpredictable flow with eddying is called ............................ flow.

3-14 Basic Aerodynamic Theory The Commercial Pilot Series


17. Sketch the different areas of pressure which will exist around a typical
cambered GP aerofoil, placed in an airflow:

a. at a low angle of attack (about 4°); and

b. at a moderately high angle of attack (about 12°).

18. The point on the leading edge at which the flow around an aerofoil is
brought completely to rest is called the ................................ point.

19. At an angle of attack of about 15°, the streamline flow around an aerofoil
breaks down and separates from most of the upper surface. This angle is
called the ......................... or ....................... angle.

20. To produce lift, downwash must ....................... upwash.

21. The resultant of all the aerodynamic forces acting on an aerofoil is called
the .................. .. .............................. (TR). It acts through a point within the
aerofoil called the ..................... of ........................... .

22. On a cambered aerofoil, the centre of pressure (CP) moves (forward/


rearward) as angle of attack is increased until, passing the stalling angle, it
moves more rapidly (forward/rearward).

23. Over the normal operating range of speeds and angle of attack, the CP of a
symmetrical aerofoil has (little if any/a lot of) movement.

24. The component of the TR:

a. which is perpendicular to the RAF is called .................... ;

b. which is parallel to the RAF is called ....................... .

Principles of Flight Basic Aerodynamic Theory 3-15


3-16 Basic Aerodynamic Theory The Commercial Pilot Series
lift
Introduction

If an aircraft is to be capable of flying it must produce sufficient lift from its


aerodynamic surfaces to both counteract its weight and enable it to manoeuvre.
As we defined in the previous chapter, lift is the component of total aerodynamic
reaction which is perpendicular to the relative airflow. Normal aircraft design is
such that by far the greatest proportion of lift is produced by the wings. Our
focus here will therefore be on wing lift and we will not be considering the small
amounts of lift which may be produced in some circumstances by the fuselage
or the tailplane.

In this chapter we look mainly at the factors affecting lift generated by a wing in
a low subsonic airflow-i.e. at airspeeds of up to about 250 knots. This is the
flight regime of most General Aviation aircraft, and in which any errors in
assuming air to be incompressible are negligible. We therefore do not have to
be concerned with the effects of compressibility, and for practical purposes, we
can say that lAS equates with dynamic pressure.

Factors Affecting Lift


General
From many years of aerodynamic testing and flight trials, it can be shown that
the magnitude of the total aerodynamic reaction (and therefore of the lift)
generated by an airflow around a wing depends upon:

• freestrearn air density (pool;

• freestream velocity (Vool:


• size of the wing (S)-in aerodynamics, the planform area is used;
• shape of the wing-both in section and in planform;
• condition of the surface-whether rough or smooth, etc; and
• angle of attack.

This list may seem at first sight to be a little complicated but in practice it can be
combined and reduced to three factors, as follows:
• The freestream density and velocity are incorporated in the expression for
dynamic pressure (1/2 pV2) which, as we saw in the previous chapter, is for
all practical purposes the measurement of lAS.
• The effect of wing area (S) is straightforward. Lift is produced as a result of
the pressure differential above and below the wing. The greater the area
that a given pressure differential can act upon, the greater the lift force that
will be produced.
• The remaining variables are combined into a single factor called the
coefficient of lift (CJ.

Principles of Flight Lift 4-1


The Coefficient of Lift
The coefficient of lift (CJ is simply a number-a multipliet~which depends on
the shape and condition of the wing and varies as angle of attack is changed.
The CLprovides a measure of the lifting capability of a given wing (or aerofoil) at
different angles of attack. It also indicates how this lifting capability changes if
the aerofoil shape changes (for example, by extending the flaps) and can be
used to assess how effective different wing designs are in producing lift. If two
wings with the same planform area are placed side by side in an airstream at the
same lAS and angle of attack, the one with the higher CL (generally the one with
the greatest camber) will produce more lift. The value of the CL for any given
aerofoil shape can only be determined through aerodynamic testing.

The Lift Formula


When brought together, the above three factors (dynamic pressure, wing area,
CL), provide what is called the lift formula:
lift = CL 1
/2 pV' s
[If a coherent system of units is used (such as the Sl system), this formula can be
used to work out the actual value of the lift force under different conditions. If
this is done, it must be remembered that the freestream values for density and
velocity-p 00 and V00 (TAS) must be used. Do not be confused by the different
functions here. For although 1/2 pV2 stands for !AS-basically true airspeed (V)
modified by the variation in density (p) from the 'standard' value-when the
symbol Vis used alone, it stands for TAS ].
What the lift formula means to us as pilots can be summarised as follows:

~ 9 qJ
lift generated by the wings =

coefficient dynamic wing


of lift pressure area

for given shape


I
equates
I
usually
depends on to constant

t ~
angle of
lAS
attack

For the majority of our flying therefore, where the wing area and aerodynamic
section remain constant (i.e. when we are not changing them by lowering flaps
etc), the lift from the wings depends only on the angle of attack and lAS.
This interrelationship between lift, angle of attack and airspeed is the most
important and fundamental in flying. If lAS is kept constant lift depends solely on
angle of attack. If, on the other hand a constant angle of attack is maintained,
the amount of lift generated will depend on lAS. Note however that the
relationship between speed and lift is a 'squared' one. If the speed (TAS-'V') is
doubled, lift will be increased four-fold, all other factors including altitude and
!. ' angle of attack remaining constant.
1 '
4-2 Lift The Commercial Pilot Series
II
Variation of CL With Angle of Attack
The most immediate and direct way of controlling the distribution of pressure
around the wing (and thus the lift), is through the angle of attack. The way in
which the CL varies with angle of attack (a) is therefore important.

Each aerofoil shape has its own particular value of CL at any given angle of
attack. When these values are plotted on a graph of CL versus a, a coefficient of
lift curve results. A typical CL cwve for a GP aerofoil is shown at Fig. 4-1. (This
cwve is similar to the CL curve for the wing of a training aircraft; with each type
of aircraft having its own particular 'inbuilt' CL cwve. The values for CL are
shown so that the reader gets 'a feel' for the sort of numbers involved).

in-flight angle of attack range

Fig. 4-1
A typical coefficient of lift curve
for a GP-type aerofoil.

The CL cwve provides a valuable insight into how a particular aerofoil will
perform in practice. Points of note from the curve in Fig. 4-1 are:

• Since the aerofoil is cambered, at oo angle of attack the aerofoil will produce
a small amount of positive lift depending on the speed of the airflow. This is
reflected in the CL having a small value at oo angle of attack.

• A cambered aerofoil must be placed in an airstream at a small negative


angle of attack if no lift is to be produced. At this angle-called the zero lift
angle-the reduction in pressure over the upper suface is balanced by the
reduction in pressure below the lower surface and no lift is produced.
Accordingly, the value of CL is zero at the zero lift angle of attack, i.e. about
minus 3° for the GP-type aerofoil. An aircraft is rarely flown at the zero-lift
angle of attack, which will only occur momentarily during some aerobatic
manoeuvres, or if the aircraft is placed in a vertical climb or a vertical dive.

• From the zero-lift angle to a moderately high angle of attack (in this case,
about 10°), the CL increases more or less in direct proportion to the angle of
attack. This linear relationship from low to moderate angles of attack occurs
with most aerofoil shapes.

• As the angle of attack increases beyond this moderate angle, the rate of
increase in CL begins to drop away until the critical (or stalling) angle is
reached. Beyond that point, any further increase in angle of attack results in

Principles of Flight Lift 4-3


a significant reduction of C,. Note that the maximum value of coefficient of
lift (C, max.) is attained at the stalling angle of attack which, in this case,
occurs at about 16". The value of C,max. and the shape of the peak of the C,
curve are important indicators. As discussed shortly, these factors can tell us
a lot about the performance of a particular wing.

• As indicated in Fig. 4-1, the normal operating angle of attack range is from a
little over 0" to an angle approaching the stalling angle (15-16"). In level
flight, low angles of attack are associated with high speed-the lift required
is generated mainly by the speed of the aircraft and a high value of C, is
unnecessary. Conversely, at slow speeds, the velocity factor in the lift
equation is very much reduced and a high Cc (high angle of attack) is
required to generate the same amount of lift.

Note: The coefficient of lift cwve is sometimes loosely referred to as the 'lift
curve'. It is not a graph of the lift force which, for straight and level flight, would
be represented by the same straight-line value when plotted against angle of
attack. Always remember that the actual value of the lift produced by an aerofoil
at any given time is a product of its coefficient of lift (which depends on angle of
attack) and the lAS at which it is being operated. Hence, a low C, (low angle of
attack) combined with a high lAS can produce the same lift as a high C, (high
angle of attack) combined with a low lAS.
The Effect of a High CL max.

All other factors being equal, a wing with a high maximum value of coefficient of
lift (high C,max ) has a greater lifting capacity and is able to produce more lift over
all of its normal operating angles of attack, than with one with a lower C,max.
The aircraft with the higher Ccmax-the 'higher lift' wing-will be able to fly at
slower speeds without stalling, and has more manoeuvrability. Lift augmentation
devices, such as flaps, are designed to increase the CL of a given wing,
particularly when being operated at at high angles of attack. On the other hand,
as we explain shortly, untoward surface roughness and such things as ice or
damage can reduce the CL.
The Shape of the CL Curve

As illustrated in Fig. 4-2, the peak of the C, curve reflects the mechanism of flow
breakaway from the upper surface of the wing as the angle of attack approaches,
and then passes, the stalling angle. At low angles of attack, the flow remains
attached to the surface almost all the way back to the trailing edge before
breaking away at the separation point to form a thin wake of turbulent flow. At a
moderately high angle of attack, the separation point begins to move forward and
a thicker wake is formed. This early stage in the breakdown of streamlined flow
is reflected by the reduced rate of increase in c,.

Beyond the peak of the cwve at CLmax., the separation point moves forward
much more rapidlly and the airflow breaks away from most of the upper surface
of the wing to form a large turbulent wake. The pressure 'envelope' over the
upper surface collapses and, accordingly, the CL-and the lift from the wing-
decline rapidly. The wing is now said to be stalled. Note that some lift is still
produced when the wing is in a stalled condition but it decays rapidly with
increasing angle of attack.

A curve with sharp peak at C, max indicates that the flow breakaway process
occurs rapidly and the wing will stall relatively suddenly. One with a more
rounded and flatter peak indicates a more gradual flow breakaway and a 'softer' stall.

4-4 Lift The Commercial Pilot Series


Fig. 4-2. separation point

Approaching stalling Beyond the stalling angle,


angle, separation point airflow breaks away from
moves forward and most of upper surface in
wake thickens. a large turbulent wake.
Rate of increase CL decreases rapidly.
in CL falls away.

Q)l
g>l
"'I
g'l
~I
Ull
I
I
I
I
I

The Effect of Camber

The effect of camber is illustrated in Fig. 4-3, which


gives representative CL curves for symmetrical, GP,
and high-lift aerofoils. The main effect of increased
camber is that the CL is increased over all normal
operating angles of attack. Note that the zero-lift angle
of attack for the symmetrical aerofoil is 0°. As camber
is increased, the angle at which the wing will stall may
also change, depending on thickness/chord (tic) ratio,
the point of maximum thickness, and a number of

other factors.
Fig. 4-3. The effect of camber on CL.

The Effect of Surface Roughness

The CLmax attained by a wing in operation is very


sensitive to the roughness of the f01ward part of upper
surface, from the leading edge to about 20 - 30%
chord. This part of the wing is normally constructed 1·2-14----/-'---,.,....1 I -standard
with a smooth finish and on some aircraft will be I~ roughness
highly polished. The reason is that flow breakaway is I 1'-...
I I contaminated
encouraged by any roughness of this part of the wing, II
which brings an early onset of the stall, and a II
significant reduction in the CLmax. Fig. 4-4 illustrates II
this effect and shows curves for the same wing with II
three different surface conditions-ve1y smooth, with II
'standard roughness', and contaminated-as is next o·
explained.
Fig. 4-4.
The effect of surface condition on CL.

Principles of Flight Lift 4-5


The recommended take-off and landing speeds for an aircraft have a safety
margin which includes a 'standard roughness' factor for the wing surface. This
takes account of any roughness arising during manufacture and normal
operational wear and tear. It takes no account of any extra contamination of the
surfaces from such things as ice, snow, frost, bird droppings, dirt, and even
insect remains or dust. These can cause a wing to stall at a lower angle of attack
than normal with a significant loss of lift.

Contamination, particularly of the leading edges and fmward upper part of


the wing surface can be dangerous.

Before flight, ensure that these surfaces are clean and free from contamination.

Three Dimensional Flow Over a Wing


So far, we have considered the airflow over a wing only in two dimensions, i.e.
vertically in relation to the chord, and in the direction of the chord from leading
to trailing edges (called the chordwise flow). There is now a need to consider
the airflow in a third direction-along the wing from root to tip or vice versa-
called the spanwise flow.

vertical
chordwise component component

~-::::::::::===;;~
c--~---==::::.. - ----- spanwise
component
(exaggerated)

Fig. 4-5. The direction of the aitflow over a wing has three components:
vertical, chordwise and spanwise.
Fig. 4-6 depicts the pressure envelope of a wing when viewed from behind. As
we know already, lift is produced when the pressure above the wing is lower
than that below. In addition, and particularly if the wing is tapered in section or
in planform, there is also often a variation in the spanwise distribution of
pressure, such that the areas of greatest pressure difference are toward the
wingroots. The total pressure distribution pattern induces an outward spanwise
flow under the wing (from higher pressure to atmospheric) and an inward flow
over the upper surface (from atmospheric to lower pressure).
As a result of this spanwise flow, vortices are formed and are shed from the

Fig. 4-6. The pressure distribution


lower pressure
above and below the wing induces a
spanwise flow. ,.....,----~---- ........
/
/
- - ' - ' '
atmospheric l / - __.. ~ ..- ...,._... ' atmospheric
pressure \ +; pressure
"'~- +-*~/
/ -- s
direction of induced flow - - -~-----
h -

higher pressure

wingtips, and from all along the trailing edge of the wings. (Note-a vortex is a
rapid whirling or spinning motion in a mass of fluid).

4-6 Lift The Commercial Pilot Series


As illustrated in Fig. 4-7:

• The wingtip vortices are the major effect and are caused by the air 'spilling'
from high pressure (through atmospheric) to low pressure around the
wingtips. The core of each vortex spins at high speed, dragging more air
from its surroundings with it and growing as it extends back from the
wingtip. Wingtip vortices can be comparatively large and can last for some
time before finally dissipating well behind the aircraft.

• The trailing edge vortices are the result of the airflow meeting at the
trailing edge at slightly different angles. They are less pronounced and less
stable than the wingtip vortices and generally become absorbed in the
turbulent and unpredictable flow of the wake from the trailing edge. It is
known however, that the more pronounced trailing edge vortices, which are
triggered by such things as small proturbences, tend to 'roll up' toward the
wingtip vortices and add to their effect.
airflow over airflow over
lower surface upper surface

trailing edge
wingtip vortices - - - •
vortices

Fig. 4-7. Wingtip and trailing edge vortices.

The formation of vortices is least at high speed and low angles of attack. Under
these conditions, the chordwise flow has greater momentum and the pressure
gradient has little effect in turning the flow in a spanwise direction. Conversely,
at low speed and high angles of attack, the pressure gradient is more effective in
turning the flow in a spanwise direction and stronger vortices are formed. The
overall effect of the wingtip vortices is to produce a downwash behind the wing
as shown at Fig. 4-8. It should be noted that each of the vortices also produces a
compensating upflow* but as this is outside the wingspan and the area being
swept by the wing, no advantage in terms of lift or drag can be gained by the
aircraft. (*Note: We will call it an 'upflow' to distinguish it from the upwash
which is normally present ahead of the wing and which does have an effect on
the airflow swept by the wing).

downwash downwash

Fig. 4-8. The wingtip vortices produce a downwash behind the wing.

Principles of Flight Lift 4-7


The 'induced' downwash affects the overall average angle of the airflow over the
wing, tilting it downward at the rear as shown at Fig. 4-9. This airflow-which is
what the aerofoil 'sees' and reacts to-is called the effective relative airflow. It
can be seen that the geometric angle of attack (between the remote RAF and
the chord line) is reduced by the downwash angle to what is called the effective
angle of attack. The difference between the effective and the geometric angles
of attack only becomes significant at the higher angles of attack/slower speeds
and has an effect on lift (as is discussed shortly) and on drag (which will be
covered in the next chapter).

high speed/low a /

remoteRAF ~
little -~,.!==~;;;::;;;;;~~---=~=:;;:··"'-;:;g~-N;,.,
~- ..
__,;;_::;:_.,
__,..,_,.,.
downwash effective RAF

Fig. 4-9.
Increased downwash reduces effective angle
the effective angle of attack. of attack

geometric angle
downwash angle of attack

The Effect of Aspect Ratio

The CL of a wing is affected not only by its cross-sectional shape (the aerofoil
section used), but also by its shape in planform. Many different planform shapes
are used with aircraft-straight wing, tapered, swept-wing, delta, and so on. All
have different lift and stalling characteristics which depend on their planform
shape and an important factor in this is the aspect ratio of the wing. In this
manual, we will limit consideration of planform shape in the main to straight
wings, which may include a degree of taper toward the tips.

Aspect ratio is the ratio of the wingspan to the chord of a wing. To give a good
basis for comparison between different planform shapes, aspect ratio is usually
measured by span2 divided by wing area (S). The wing area used is gross wing
area (i.e. it includes that area 'cut out' by the fuselage) as shown in Fig. 4-10.

k------ span -------J>l


span
aspect ratio = chord

Fig. 4-10. 2
Aspect ratio. measured by: _ _span
-'-,:-_ __
gross wing area

4-8 Lift The Commercial Pilot Series


Aspect ratio (AR) has a major effect on the formation of vortices, and therefore
on the amount of induced downwash. The higher the AR, the lower the amount
of downwash, all other factors remaining equal. With a high AR wing, the
chordwise flow has little time to develop as the air crosses the wing.
Proportionally less air therefore 'spills' over the wingtip and the angle at which
the air meets at the trailing edge is small. In addition, the area affected by the
downwash is but a relatively small proportion of the total area behind the wing.
Conversely, with a wing which has a relatively short wingspan and a long
chord-i.e. a low AR-a much larger proportion of the total flow is spilt around
the wingtip; the angle at which the airflow meets is greater; and the area behind
the wing affected by downwash is much greater.

For a wing of given area and section, the effect of decreased AR is to:
• decrease the CLmax ;and
• increase the (geometric) stalling angle.

Fig. 4-11 (which is for two wings with a different aspect ratio but with the same
section and wing area, both at the critical angle) illustrates how increased
downwash on the low AR wing tilts the TR further to the rear and reduces the lift
vector. As both wings have the same section, they will stall at about the same
effective angle of attack-the angle which the wing 'sees'. However, the
geometric stalling angle of attack-which is what the pilot sees through nose
attitude-will be noticeably higher in the aircraft with the low AR wing.

TR L
Fig. 4-11.
A--------------r-~~~~~~~
' .li. L
The effect of AR on lift. '
:' reduced lift vector ''
:
'' ''
''' ''
' '
RAF
RAF
effective RAF

· eRP..f
HIGH ASPECT RATIO ell:ec\1'-1 LOW ASPECT RATIO

You can see this difference between effective and geometric angles of attack if,
for example, you obse1ve the nose attitude of a high AR-winged aircraft (like a
glider) just prior to touchdown. It will be relatively low. By contrast, a low AR-
winged aircraft, like the Concorde, has an extremely high nose attitude-well
over 30°-just before touchdown, necessitating the use of the 'droop snoot' so
that the pilots can still see the runway! Admittedly. the wing sections and
planform are vastly different between Concorde and a glider, but a principal
reason for the different angle of attack at high CL is the difference in AR.

Unless stated otherwise, when we refer to angle of attack throughout this


manual, we mean the geometric angle of attack. Remember that the geometric
angle of attack is the angle between the chord line and the remote relative
airflow. It is related to the nose attitude which the pilot sees. In many diagrams
(like the one following), changes in angle of attack will be indicated. These will
be changes to the geometric angle of attack and it is worth noting that as a
general rule most wings, regardless of planform or section will reach their CLmax
at about 15-16° effective angle of attack.

Principles of Flight Lift 4-9


Fig. 4-12 shows CL curves for a typical high AR wing (like a glider wing); one of
moderate AR as may be found on a typical training aircraft; and a low AR wing.
The effect of AR in reducing the CL max can be clearly seen. All other factors
being equal, a high AR wing is more efficient in producing lift. The increase in
geometric stalling angle as AR reduces can also be seen, but remember that the
effective angle of attack for each of these wings will be much the same.

Fig. 4-12.
The effect of aspect ratio on CL.

mediumAR

same wing section and


wing area

0'

As all three of the wings illustrated have the same wing section, the respective
peaks of the CL curves are similar in shape. The shape of the curve in this area is
determined by a number of factors including the roundness of the 'nose' (or
leading edge), the t/c ratio and the point of maximum thickness. These factors
are discussed in more detail in a later chapter.

4-10 Lift The Commercial Pilot Series


Review4
I. Write down the lift formula and state what each of the factors stands for.

2. In flying, the wing area is usually constant. If the flap setting remains
unchanged, the amount of lift generated depends on two factors. These
are ............................................ and ................. .

3. The coefficient of lift (CLl is simply a number which describes the


lifting .................................. of an aerofoil shape at different angles of attack.

4. Sketch a typical CL cu!Ve for a GP aerofoil. Annotate the approximate angles


of attack for CL = 0, and CL max.

5. In level flight:
a. high speed is associated with a (high/low) angle of attack, and
therefore (high/low) CL.
b. low speed is associated with a (high/low) angle of attack, and
therefore (high/low) CL.

6. All other factors remaining equal, if the CL max. of a given wing can be
increased, the aircraft will:
a. be able to fly level with a (higher/lower) angle of attack at the same
speed;
b. stall at a (higher/lower) speed;
c. have (increased/decreased) manoeuvrability.

7. A CL cu!Ve with a sharp peak at CL max. indicates that flow separation and the
stall occur relatively (slowly/suddenly).

8. A highly cambered aerofoil has (greater/less) lifting capability than one with
less camber.

9. Contamination of wing surfaces with such things as ice, snow, frost, bird
droppings etc., can cause the wing to stall at a ........................ angle of attack
than normal and result in a significant ....................... of lift.

I 0. The generation of wingtip and trailing edge vortices is greatest at ................. .


speeds and ........................... angles of attack.

I I. Wingtip and trailing edge vortices combine to induce an additional


downwash behind the wing. This downwash reduces the (geometric/
effective) angle of attack.

12. At the same effective angle of attack, increased downwash tilts the total
reaction (TR) further back which (increases/decreases) the lift vector.

13. Aspect ratio (AR) is the ratio of ...................... to ....................... .

14. High AR wings have (greater/less) induced downwash than those with low
AR.

15. If two wings have the same aerofoil section but different AR, the high AR
wing has a (higher/lower) CL over the normal operating range of angle of
attack and will stall at a (higher/lower) geometric angle of attack.

Principles of Flight Lift 4-11


4-12 Lift The Commercial Pilot Series
Drag
Introduction

Drag is the aeronautical term for the air resistance experienced by an aircraft in
flight. In Chapter 3-when discussing aerofoils-we defined drag as that
component of aerodynamic reaction which acts parallel to the relative airflow
and opposes the motion of the aircraft through the air.

In flight, each and every surface of an aircraft will produce an aerodynamic


force. Of all of these surfaces, it is the wings which produce by far the greatest
proportion of the total lift. For practical purposes therefore, we could afford to
disregard any small contributions made to total lift by the tailplane, fuselage or
any other parts of the aircraft. With drag it is different. A much greater
proportion of total drag is comes from the fuselage and surfaces other than the
wings which are exposed to the airflow. Hence, when discussing drag, we must
take the whole of the aircraft into account.

The main purpose of the power-plant produced thrust is to move the aircraft
through the air. To maintain any given speed the amount of thrust produced
must be sufficient to overcome the total drag at that speed. Drag is the enemy of
efficient flight. The lower the drag, the less the thrust required to counteract it-
and the higher the maximum level flight speed which can be attained with a
given engine. The advantages of a lower thrust requirement are obvious-
smaller, possibly fewer engines, lower fuel flows, less strain on the engines and
associated structures, and lower operating costs.

drag

-....
thrust d~&'/
/
~
~
~
...
thrust

both aircraft 100 KIAS

Fig. 5-1. Lower drag requires lower thrust to counteract it.

However, not all about drag is bad. The ability to deliberately increase it-for
example by lowering the undercarriage, deploying airbrakes or spoilers on high-
performance aircraft-confers a distinct advantage in enabling the pilot to slow
the aircraft quickly, and reduce the length of the landing run. Almost all aircraft
are equipped with flaps to provide for an increase of lift. A consequence of
lowering the flaps is an increase in drag which gives the advantage of steeper
and safer approach angles and shorter landing distances. At other times,
increased drag from lowered flaps enables higher power to be used at low
speeds, giving better control and throttle response. (These effects are discussed
in more detail later).

Principles of Flight Drag 5-1


Classification of Total Drag
The total drag on an aircraft is the sum of all those components of aerodynamic
force which act parallel and opposite to the direction of flight.

It is usual and convenient to group the various sources of drag under different
headings so that they may be more easily studied and understood. Drag is
classified in slightly different ways by different authorities. For convenience, in
this manual we will consider drag under two main groups:
• The drag force directly associated with the production of lift, known as
induced drag, which arises from the generation of wingtip and trailing-edge
vortices.
• The drag forces not directly associated with lift production, known as
parasite drag. Within this group are:
-profile drag (form drag and skin friction) and
-interference drag.

I TOTAL DRAG I
I
I
induced drag parasite drag

profile drag interference drag


Fig. 5-2. The various types of drag.

skin friction form drag

NOTE: We outline an alternative way of grouping the above types of drag


toward the end of this chapter.

Parasite Drag
All of the elements of parasite drag-skin friction, form drag and interference
drag-arise because air is a viscous medium. This physical quality of viscosity is
easily seen and felt in a thick fluid like treacle, but is difficult to imagine in a thin
and invisible fluid like air. Although the viscosity of air is much lower, it is there,
and it causes the particles of air to 'stick' to one another and to any surface they
come into contact with for the same reason that the particles of treacle 'stick'
together and adhere to a spoon.

The effects of viscosity are felt through a relatively thin layer of air adjacent to the
surface of a moving body, called the boundary layer. The nature of the airflow in
the bounda1y layer has a significant effect on the lift and drag characteristics of
an aircraft, hence it deserves study in a little more detail.

The Boundary Layer


When air moves past a solid body, its viscosity causes the particles next to the
surface to adhere to the surface of the body, and the airflow in the immediate
vicinity to be slowed down. This layer of retarded flow which is sandwiched
between the local freestream flow and the surface of the body is known as the
boundary layer. Beginning at the outer edge of the boundary layer, the velocity
of the airflow is progressively decreased until it it brought to a halt at the surface.
The bounda1y layer in the airflow over a wing is usually relatively thin-no more
than a maximum of about 2 or 3 em in depth.

5-2 Drag The Commercial Pilot Series


As illustrated at Fig. 5-3, the flow within the bounda1y layer exists in two forms:

• Laminar flow. The initial part of the flow over most smooth surfaces is
laminar in nature. That is, each successive sheet of air slides smoothly over
the one nearer the surface and there is a relatively uniform increase in
velocity from zero at the surface to the freestream value at the outer edge.
Laminar flow boundary layers are very thin-in the order of 2 mm in depth.

• Turbulent flow. After progressing for a certain distance over a surface


(even if that surface is flat and aligned with the airflow) the flow in the
boundary layer normally becomes turbulent and the layer becomes much
thicker. This turbulent bounda1y layer flow is characterized by high-
frequency eddies and swirls and there is considerable inter-mixing of the
flows at successive levels. The point at which the flow changes from
laminar to turbulent is known as the transition point-although in reality
this process takes place over a finite distance. As faster-moving air (with
greater kinetic energy) from the outer part of the boundmy layer mixes with
the air nearer the surface, the lower flow is 're-energized'. This enables the
air nearer the surface to retain more of the freestream velocity (i.e. not to be
slowed up as much) resulting in a changed velocity profile as shown in the
diagram. Turbulent-flow bounda1y layers are about 10 times thicker than
their laminar counterparts-on average in the order 2 em thick.

not to scale I I turbulent flow I


transition point

I laminar flow I
V freestream \ T
2cm
approx

2mm approx

velocity profile velocity profile


1
Fig. 5-3. The two types of boundary layer flow.

Skin-friction Drag
Skin-friction drag is the result of shear stress between successive levels of air
within the boundary layer. Shear stress is the force required to separate the air
particles at one level from those at the next and move them along at a faster
rate. In the laminar-flow boundary layer, where one 'sheet' of air slides
smoothly over its neighbour and the rate of change of velocity between
successive sheets is gradual, the drag produced is relatively low. In a turbulent-
flow boundary layer, the intermixing between air particles from different levels
prevents this smooth sliding effect and, as the rate of change of the velocity of
the flow near the surface is less gradual, there is greater shear stress and the
resulting skin-friction drag is much higher.

It follows from the foregoing that to keep skin-friction drag to a minimum, it is


better to have a laminar bounda1y layer over as much of the aircraft surfaces as
possible.

Principles of Flight Drag 5·3


Factors Affecting Skin-friction Drag
Speed

An increase in speed means that the rate of change of velocity across the flow in
the boundary layer is increased which increases the shear stress. Skin-friction
drag increases in proportion to the square of lAS.

Shape

Laminar-flow boundary layers are sensitive to adverse pressure gradients, which


occur where the flow is toward an area of higher static pressure. If the airflow
around an aerofoil at a low angle of attack is considered (Fig. 5-4), it can be seen
that the lowest static pressure is located on each surface at about the point of
maximum thickness. Laminar boundary layer flow can usually be maintained
from the leading edge to this point-i.e. across the area with a favourable
pressure gradient. As the airflow progresses beyond the point of lowest pressure
toward the trailing edge, the pressure gradient becomes increasingly adverse,
and the boundary layer responds by transitioning to turbulent flow.
low pressure peak
and transition point
favourable adverse
pressure pressure
gradient _ _ gradient

laminar flow "" / / / - - ~~' , / turbulent flow


boundary layer>-~::~:> c:::--~_,/ boundary layer
r-- - ,.- -- --

~::~-:::::::::::::::-:-:~:~:~::':,::~~~
,~ :0,.
,--:;:
c-
'--
'
I:::_-::::--'2 ~- -
---~>
tc------------=-----v
- ----_-___
~ r-:;_"7
+ wake

low pressure peak boundary layer depth


Fig. 5-4. The effect of pressure exaggerated
and transition point
gradient on boundary layer flow.

NOTE: Static pressures are transmitted without modification through the


boundary layer to the surface of an aerofoil.

Hence, laminar boundary layer flow is encouraged by long slender aerodynamic


shapes which have the point of maximum thickness located well back This is
one of the reasons why high-speed aero foils have this point located at about 50%
chord. This delays transition and enables a greater proportion of the wing to be
covered with a laminar flow boundary layer-resulting in less skin-friction drag
at high speed.

Surface Condition

Laminar flow is also very sensitive to surface irregularities, and any roughness-
to a degree which can be felt by the hand for example-is sufficient to cause the
flow to become turbulent, even if the pressure gradient is favourable. For this
reason, the forward surfaces of most aircraft are generally constructed with a
smooth surface finish and, on high speed aircraft particularly, measures such as
flush riveting and protection against damage are taken to keep the surface free
from irregularity.

5-4 Drag The Commercial Pilot Series


Although, when in flight, a turbulent boundary layer will exist on most aircraft
surfaces from the 'point of maximum thickness' toward the rear, a smooth surface
finish in this area is also important in keeping the bounda1y layer from becoming
too thick and in helping to delay separation, as will be described shortly.

Size

If the size of a body of a given shape in a given airflow is increased, there is an


increase in drag (and lift, if it is an aerofoil), which is out of proportion to the
increase in size. For example, if the size of a given shape is doubled, the increase
in drag (and lift) under the same airflow conditions will be more than doubled.
This is called scale effect, and it is measured by Reynolds Number. (Reynolds
Number relates the size of a body to the density, velocity and viscosity of the
airflow. Although it is an important parameter in aerodynamics, it is a complex
subject and a full discussion is beyond the scope of this manual.) Thus, to keep
drag low it is better that not only should aerodynamic shapes be slender, but also
as small as possible.

Surface Area

The magnitude of skin-friction drag depends on the 'wetted area' of the aircraft-
i.e. the total surface area exposed to the airflow.

Angle of Attack

At high angles of attack, the low pressure peak over the upper surface of a wing
moves forward and the transition point moves forward in sympathy with the
change in pressure gradient. If the angle of attack is increased at any given speed,
there is a slight increase in skin-friction drag as the transition point moves forward
and a greater proportion of the surface becomes covered with a turbulent
boundary layer.

Form Drag
Whenever a solid body moves through air, the pressure on the forward-facing
surfaces will always be higher (even if only slightly) than on the rearward-facing
surfaces. That component of the force generated by the pressure difference which
is parallel to the airstream is form drag. This fore/aft pressure difference (form
drag) has its origin in the separation of streamline flow around the body and the
formation of a turbulent wake. The pressures inside this wake are always lower
than those forward of the body.

The pressure dependency of form drag can be readily appreciated in Fig. S-5,
which shows a flat plate placed in two different attitudes in an airstream. With the
plate set at right angles, a large turbulent wake is formed. The difference of
pressure ahead and behind the plate-and form drag-are at a maximum. On the
other hand, with the plate laid parallel with the flow, most of the drag will be from
skin fliction. There will still be a small turbulent wake to the rear and, although
minimal, there will be a small difference in pressure front and rear and therefore
some form drag present.

-+-drag
----....,~drag
Fig. 5-5. Mostly form drag and mostly skin-friction drag.

Principles of Flight Drag 5-5


The point at which the streamline flow around an aerodynamic surface breaks
down to form a turbulent wake is important in determining how much form drag
will be generated. As we saw from the previous chapter that point is called the
separation point. The separation point should not be confused with the
transition point which is related only to the flow inside the boundary layer. The
nature of the turbulent flow in the wake is also different from the turbulent flow
in the bounda1y layer-the eddying is slower and on a much larger scale. To
distinguish between the two, the turbulent flow in the wake is often referred to
as separated flow.

The point of separation of the streamline flow around an aerofoil shape is


determined by the conditions in the boundary layer. As it flows from the point of
maximum thickness of the aerofoil toward the trailing edge, the turbulent
bounda1y layer continues to thicken gradually and, because of friction, the flow
nearest the surface slows down. At some point, this lower flow is slowed so
much that it stops and may begin to reverse as a result of the adverse pressure
gradient. When that occurs, the flow as a whole is no longer able to conform
with the shape of the body and separation occurs. Fig. 5·6 shows how the
velocity profile changes in the boundmy layer, leading up to separation.

turbulent
boundary
Fig. 5·6. Separation occurs layer
when the lower flow in the flow
boundary layer slows to a stop
and begins to reverse.

aerofoil

Fig. 5-7 shows the relationship between the bounda1y layer and form drag on an
aerofoil placed in an airstream at constant speed but at different angles of attack.

At a low angle of attack, the pressure distribution and adverse gradient above the
aerofoil are such that the boundmy layer is able to maintain its energy almost all
the way back to the trailing edge before separation occurs. The wake formed is
small and form drag is low (Fig. 5-7 a).

As the angle of attack is increased, the peak of low pressure moves forward and
the pressure gradient becomes increasingly adverse-the separation point
moves forward, the wake thickens and form drag increases (Fig. 5-7b).

As the critical angle is approached, this process is accelerated and the


separation point begins to move rapidly forward until separation has occurred
over most of the upper surface and the aerofoil stalls with a sudden Joss of lift
accompanied by a large increase in form drag (Fig. 5-7 c). At this angle of attack,
the aerofoil has begun to perform more like the flat plate placed at right angles
to the airflow shown in Fig. 5-5.

5-6 Drag The Commercial Pilot Series


transition
point
separation
Fig. 5-la. ' point

r==~~i!!m;4;x,
Low angle of attack; separation point
well back; small wake, low form drag.

transition
point

Fig.5-7b. separation
As angle of attack increases, boundary layer point
thickens; separation point moves forward; wake
thickens; form drag increases.

separation
point

Fig. 5-lc.
Passing the stalling angle, separation point
moves rapidly forward; large turbulent wake
forms; form drag increases dramatically.

Factors Affecting Form Drag


Form drag can be a large part of total drag and good design should aim to
reduce it as much as possible. The main factors affecting form drag are shape
(streamlining), size, angle of attack and airspeed.

Streamlining

Streamlining is aimed at reducing the effect of adverse pressure gradients-by


making the curvature of surfaces more gradual, particularly toward the rear.
This delays separation; which in turn reduces the size of the turbulent wake and
reduces the pressure difference between forward and rearward surfaces. The
dramatic reduction in form drag which can be achieved by streamlining is
illustrated at Fig. 5-8, which is for four different shaped bodies with the same
cross sectional area.

DRAG

DRAG

Fig. 5-8. Streamlining,


especially behind the
shape, reduces form drag
substantially.

=~~:::::------.._

:~DRAG
----== I
o~s~~.~,~s'~k--------~s~o~~.------------~,~oo"'%

Principles of Flight Drag 5-7


Other measures for reducing form drag include such things as retracting the
undercarriage within the wings or fuselage so that it is not left exposed to cause
drag when not in use. Or, if that is not possible, by adding fairings to present a
more streamlined shape.

Fig. 5-9. The addition of fairings reduces form drag

Streamlining of shapes is made less effective if the aircraft surfaces have


irregularities such as ice or damage-anything which intenupts the smooth
streamlined flow will precipitate separation and increase form drag.

-drag ice accretion drag

Fig. 5-10. Ice, damage, or any irregularity on the airframe will increase drag.

The effectiveness of streamlining a body of a given cross-sectional area for


subsonic airflows is determined by fineness ratio. The fineness ratio is the ratio
between the length of a body to its depth (or breadth), as shown in Fig. 5-11.
Least form drag is achieved with a fineness ratio of between 3 and 4 with the
maximum depth placed at about one-third of the distance back from the leading
edge. A fair amount of latitude may be taken with these dimensions without
much increase in drag, but if the fineness ratio becomes too high, the surface
area and skin-friction drag are unnecessarily increased.

Fig. 5-11. Fineness ratio is given by Length (L)


divided by Depth (D).
L -------------~

Size

The magnitude of form drag is proportional to the size of the surface on which
the front and rear pressure difference acts. To keep form and skin-friction drag
(i.e. profile drag) to a minimum, it is preferable that cross-sectional areas when
viewed from the forward aspect, be kept as small as possible in the design of
aircraft.

5-8 Drag The Commercial Pilot Series


Angle of Attack

As we have seen, if the angle of attack of an aerofoil is increased at constant


speed, the separation point moves forward and form drag increases. It should
be noted that this forward movement of the separation point with increased
angle of attack is not confined to the wings. The airflow around other surfaces
such as the fuselage will be similarly affected and they will also produce more
drag. Another effect of increased angle of attack is that all aircraft surfaces
generally present a greater frontal area to the oncoming airflow. Hence, at any
given speed, the form drag of an aircraft as a whole increases as angle of attack
is increased.

Speed

Form drag is a function of the dynamic pressure acting on an aircraft and


therefore increases with the square of lAS. Note that for level flight, the effects of
angle of attack and airspeed on form drag, oppose one another. The effect of
airspeed is, however, the more powerful. Hence at slow speed/high angle of
attack, the form drag is lower than it is at high speed/low angle of attack.

Interference Drag
The total parasite drag produced by an aircraft is greater than just the sum of the
skin-friction drag and form drag generated by the individual components which
are exposed to the airflow. Additional drag is caused by the mixing, or
interference, of converging airflows at the junction of various surfaces, such as
the wing/fuselage junctions, the tail section/fuselage junctions and the wing/
engine nacelle junctions. This additional drag is referred to as interference
drag.

Wherever the airflows from the various surfaces of the aircraft meet, a wake is
formed behind the aircraft. The additional turbulence which occurs in the wake
causes a greater pressure difference between the front and rear surfaces of the
aircraft and therefore increased drag.

Suitable filleting and blending of shapes to control local pressure gradients can
aid in minimizing interference drag. A fairing is a part of the skin (external
surface) of an aircraft added to encourage smoother blending of different
airflows and reduce eddying and the resultant drag.

Fig. 5-12.
A wing-root fairing.

Interference drag also increases with the square of lAS.

Induced Drag
Induced drag will be present whenever the wings are producing lift. To that
extent, it is often said that induced drag is a part of lift. It arises from the
downwash induced by the wingtip and trailing edge vortices which, for a given
amount of lift being produced, tilts the total reaction force further backward
through the induced downwash angle. (Refer again to Figs. 4-7 to 4-9). This
extra rearward tilt, in effect, increases the length of the drag vector and it is this
increase in drag which is known as induced drag.

Principles of Flight Drag 5-9


A diagrammatic explanation of induced drag is given in the following illustration.
Fig. 5-13a shows the forces acting on a section of a hypothetical wing which we
can imagine as being infinitely long-i.e. having no wingtips and, therefore, no
vortices and no way for a spanwise flow to develop. With no induced
downwash, the effective angle of attack is the same as the geometric angle of
attack-which you will recall is the angle between the remote relative airflow
and the chord line.

Fig. 5-13b shows the same section at the same geometric angle of attack, but
this time of a real wing of finite length and therefore having vortices. The
induced downwash decreases the effective angle of attack-the magnitude of
the TR and its vertical component (lift) is reduced as a result.

To restore this loss of lift, the geometric angle of attack must be increased by the
downwash angle until the effective angle is the same in Fig. 5-13c as it was in
Fig. 5-13a. In doing this, the TR becomes more tilted to the rear resulting in an
increase of the length of the drag vector. This increase in drag is induced drag.
drag

TR
Fig.5-13.
a. With no induced downwash, lift
the effective angle of attack is
the same as the geometric angle.

airflow

lift reduced

b. With induced downwash,


the effective angle of attack
and lift are reduced.
induced downwash

induced drag
IV I

c. To restore lift to its former value,


the effective angle of attack must
be increased. This tilts the TR
further toward the rear, increasing
the length of the drag vector-
producing induced drag.

5-10 Drag The Commercial Pilot Series


The angle through which the TR tilts toward the rear is determined by the pressure
distribution and the direction of the effective airflow. By definition, the lift and
drag components of the TR must be resolved with respect to the remote RAF
which reflects the direction of flight. This is the nub of induced drag. Obviously,
the smaller the angle of induced downwash, the lower will be the induced drag.

Factors Affecting Induced Drag


There are a number of factors affecting induced drag, including aspect ratio (AR),
wing planform shape, and C, (which includes considerations of angle of attack,
weight and airspeed). These factors are incorporated in the coefficient of induced
drag (Coil which is:

where 7t = the fixed ratio, 22/7;


e =wing efficiency factor (see below);
AR = aspect ratio.
Aspect Ratio

The effect of aspect ratio on the production of vortices has been covered in the
previous chapter. The higher the AR, the nearer the wing will be to becoming
infinitely long (Fig. 5-13a). High AR wings produce smaller vortices and, in
comparison with a wing of lower AR, proportionally less of the airflow swept by
the longer span is affected by the vortices. Consequently, the induced downwash
angle, when averaged over the whole of the high AR wing, is smaller and the
induced drag low.

Wing Planform Shape

For a wing of given span, an elliptical planform shape produces the smallest
vortices and therefore the lowest induced drag. Because of their difficulty in
construction, not many aircraft have been built with this planform shape-perhaps
the most famous example being the World War II Spitfire. However, for wings
with straight leading and trailing edges, the judicious use of taper and washout of
the wing sections toward the tips can produce a similar reduction in induced drag.
Most straight wings produce between 5 to 15% more induced drag than an
elliptical wing and this is accounted for by the wing efficiency factor (e) in the
above equation for Co;.

Coefficient of Lift

From the pilot's point of view, where the AR and planform shape of the aircraft are
fixed, the important factors in determining induced drag are angle of attack,
airspeed and aircraft weight. These are incorporated in the CL' factor in the
above equation which can be seen to have a powerful effect on the amount of
induced drag generated.
• Angle of attack. Induced drag increases as the angle of attack is increased.
The strength of the vortices is determined by the pressure difference above
and below the wing. When the wing is at the zero-lift angle of attack (C, = 0)
there are no vortices and therefore no induced drag. As the angle of attack is
increased, vortices form and increase in strength up to the angle for C, max.
Induced drag therefore increases with angle of attack to be at a maximum at
the stalling angle.

Principles of Flight Drag 5-11


• Airspeed. It can be shown that induced drag is inversely proportional to the
square of lAS. This is the opposite to the effect of airspeed on parasite drag,
which is directly proportional to lAS'- When the factors of angle of attack
and airspeed are combined, induced drag is greatest at low airspeeds and
high angles of attack. For an aircraft just after take-off for example, induced
drag can be as high as 75% of total drag. When an aircraft is manoeuvring at
high speed, although induced drag is proportionally lower, it is still
significant because of the high angle of attack being used.

• Weight. Increased weight means that higher angles of attack must be used
to produce a given amount of lift at any given speed. Induced drag increases
in proportion to weight squared (W2).

An alternative way of looking at induced drag is as follows. The production of


vortices is an inevitable consequence of the production of lift with a wing of
finite span. These vortices result in an induced downwash which is over and
above the downwash necessary to produce lift. To produce a rotary motion in
any fluid requires energy-an example is the energy required to stir a large
volume of water in a drum with some sort of paddle. In flight, the energy
required to create the vortices (to stir the air) must come from somewhere.
Ultimately, that demand is placed on the engine by requiring higher power to be
used to offset the induced drag when it is desired to maintain a given speed.

Consider it this way. Whenever the wing is 'working hard' to produce lift-i.e. at
a high angle of attack-induced drag will be high and will cause a reduction in
airspeed or require an increase in power to compensate.

Measures for the Reduction of Induced Drag


Many aircraft have measures incorporated in the design of the wings to reduce
the effect of induced drag. As has already been mentioned, washout combined
with taper is one measure. Washout is a reduction of the angle of incidence*
(and therefore the geometric angle of attack) of the wing sections toward the
wing tips. (* The angle of incidence is the angle between the chord of the wings
or tail plane with respect to the fore and aft axis (or line) of the airframe.)

wingroot
section

Fig. 5-14. Washout is a decreased angle of incidence toward the tips in the construction of
wings. Combined with taper, it can reduce induced drag.

NOTE: Another important reason for using washout is to change the stalling
characteristics of an aircraft as discussed in Chapter 8.

Other measures include wing fences (to straighten and control the spanwise
flow); and the modification of the wingtips (drooping, winglets, wingtip tanks) as
shown in Fig. 5-15.

5-12 Drag The Commercial Pilot Series


straight wing

wing let wing fence

Fig. 5-15. Modification ofwingtips can reduce the strength of the vortices formed.

Total Drag
The Coefficient of Drag
The total drag on an aircraft is a combination of parasite drag and induced drag.
The coefficient of total drag (C 0 ) is therefore:

Co = Co parasite + Co induced
A typical CUJve of Co (for the aircraft as a
whole) plotted against angle of attack is
shown in Fig. 5-16. Again, actual values of C0
are given so that the reader has a feel for the
sort of numbers involved.

Note that the minimum Co will generally


occur when the wing is at about zero degrees
angle of attack. Whereas, at that angle the CL
is very small and little lift or induced drag is
produced, the aircraft still generates parasite
drag. As the angle of attack is increased, Co
increases through induced drag and
increments of parasite drag. Beyond the
stalling angle, the increase in C0 becomes
more rapid due mainly to the effect of
separation and increased form drag.
-~ o· ~ ~ 1r 1~

By itself, the C0 curve does not have a lot of


angle of attack a
utility and can be misleading. It must not be
taken as indicating the magnitude of total Fig. 5-16. The coefficient of drag curve.
drag. In flight, small angles of attack are
normally associated with the higher speed
range of an aircraft. In that range, although
the Co is low, the total drag is high-and
increases with the square of airspeed.

As with the CL and lift CUJves therefore, be careful in distinguishing between the
coefficient of drag curve and the drag curve.

Principles of Flight Drag 5-13


The Drag Curve
The drag formula is very similar to the lift formula, which is as could be expected
since both are components of the same force. For flight in the low subsonic
speed range, the drag formula is:
Drag= Co 1/z pV' S

(Note: When calculating the drag of an aircraft with this formula, 'S' stands for
the total frontal area. However, for convenience, when considering the drag of
an aerofoil alone, or comparing drag with lift, as for the lift formula, 'S' stands for
the wing planform area).

It is normal and useful to plot the total drag of an aircraft against lAS in straight
and level flight. The curve which results is more descriptive than the Co curve in
telling what happens to drag in flight. The way in which drag varies with speed
is an important consideration in aircraft performance and will form part of the
discussion again in later chapters.

slow fast

~I
"'"I
total I
Fig. 5-17.
drag
The drag curve. %, "&'li"-'
'6 .,__0
,~ oc lAS'
"'o- ~'li
10-<;: <I
oc lAS' :9

Slow
{high a)
r lAS

speed for minimum drag


Fast
(Iowa)

and lifUdrag max.

The drag curve shows, that for straight and level flight:
• Total drag is high at slow speeds (high angle of attack) due mainly to the
contribution of induced drag.
• Total drag is also high at high speeds (low angle of attack), and consists
mainly of parasite drag.
• Minimum drag is experienced at an intermediate airspeed where induced
drag and parasite drag are equal. This is also the speed at which the ratio of
lift to drag is at a maximum.

In straight and level flight, lift equals weight, and at the minimum drag speed
sufficient lift is produced to counteract the weight but with the minimum
amount of drag. For level flight at a constant speed, thrust must be sufficient to
counteract the drag and thus the minimum amount of thrust will be required for
this at the minimum drag speed.

5-14 Drag The Commercial Pilot Series


The minimum drag airspeed is a major consideration-a number of factors of
aircraft performance are related to it. The drag curve-drag versus lAS (angle of
attack)-is an extremely important relationship. It is really a summary of the
main factors we need to know about drag.

Lift/Drag Ratio
The aerodynamic performance and efficiency of the aircraft are determined by
the lift/drag ratio at different angles of attack. In a sense, lift is the benefit
obtained by moving the aircraft through the air, and drag is the penalty paid for
it. Aerodynamically, to obtain the greatest benefit for the least penalty-in other
words, to obtain the greatest efficiency from the wing-the aircraft must be
flown at the airspeed which gives the angle of attack for the best lift/drag ratio.

The lift/drag (LID) ratio at different angles of attack can be obtained by


comparing CL against C0 since:
lift (l) CL Y, pV' S
drag (D) Co Y, pV' S

When the value of CLI C0 is plotted against angle of attack we obtain the LID
ratio curve. A typical example is given in Fig. 5-18.

The peak of the curve indicates the angle of


attack for maximum L/0 ratio-in other
words, the most efficient angle. For most
aircraft, this is normally about 4'a. In straight
and level flight at a given weight, this angle is 25
associated with one !AS. L
At higher speeds (lower angles of attack), the D 20
LID ratio falls off rapidly. The aircraft is less
efficient aerodynamically. or 15

At lower speeds (higher angles of attack) the


LID ratio also reduces, but less rapidly until
the stalling angle where there is a marked
reduction. Again, in this speed range, the 5
aircraft is not as efficient aerodynamically.
4'
In the majority of aircraft, the pilot does not angle of attack a
have an instrument which gives a direct
reading of angle of attack. Angle of attack is Fig. 5-18. Lift/drag ratio versus angle of attack.
however related to lAS, hence when a
specific angle of attack such as the most
efficient angle is required, this is achieved by
flying the aircraft at the related !AS.

Some important in-flight performance requirements are obtained at the angle of


attack for best LID ratio, such as the maximum cruise range and the maximum
gliding range. To obtain the required performance, it is important that the
aircraft be flown at the particular lAS associated with this angle of attacl{. These
factors will be discussed in more detail later.

Principles of Flight Drag 5-15


An Alternative Classification of Drag
Another way of classifying drag is to place it into two groups; zero-lift drag and
lift-dependent drag.

• Zero-lift drag is, as its name implies, the drag experienced when the
aircraft is flown at the zero lift angle of attack. As previously mentioned, this
situation rarely occurs in flight-only momentarily during aerobatic
manoeuvres or in a truly vertical climb or dive. At this angle of attack, all of
the drag is made up of skin-friction, form and interference drag. As the wing
is not producing lift, no induced drag will be generated.

• Lift-dependent drag consists of induced drag plus increments of skin-


friction, form and interference drag all of which arise when angle of attack is
increased and lift is being generated.

Under this system, total drag is therefore the sum of zero lift drag and lift
dependent drag. The coefficient of zero-lift drag remains constant, with the
amount of zero-lift drag produced relying on the other factors in the drag
equation. The coefficient of lift dependent drag is as described for induced drag
on page 5-11, but modified to include the increments of skin-friction, form and
interference drag at different angles of attack. Co total therefore = Co zero lift (a
constant figure) +Co induced (suitably modified to include the increments of what
we have referred to previously as parasite drag).

5-16 Drag The Commercial Pilot Series


ReviewS

1. The total drag on an aircraft is the sum of all the components of


aerodynamic force which act ....................... and ............................ to the
direction of night.

2. Total drag can be placed in two groups - induced drag and parasite drag.
List the types of drag which make up parasite drag.

3. In an airflow past a solid surface, the layer of air which is retarded by


viscosity adjacent to the surface is called the ....................... ................... .

4. The two types of flow in the boundary layer are called ........................... flow
and ............................ flow.

5. Turbulent bounda1y layers are much (thinner/thicker) than laminar


boundary layers and generate (more/less) skin-friction drag.

6. The point at which the boundary layer changes from laminar flow to
turbulent flow is called the ........................... point .

7. With an aerofoil which has a smooth surface, a laminar flow boundary layer
can generally be maintained from the leading edge to the point
of.......................... ............................ at small angles of attack.

8. As the boundary layer progresses past the low pressure peak on the upper
surface of an aerofoil, the ........................... pressure gradient causes the flow
to become turbulent.

9. As angle of attack is increased, the transition point on the upper surface of


an aerofoil moves (forward/rea1ward) and skin-friction drag (increases/
deceases).

10. Which of the following factors affect skin-friction drag - speed; shape;
surface condition; size; wetted area; angle of attack?

I I. The type of drag which results from the pressure difference between the
forward and rearward-facing surfaces of an aerodynamic body is
called ................ drag.

12. The front and rear pressure difference is caused by the breakdown (or
separation) of streamline flow and the formation of a turbulent wake. The
point at which this occurs is called the ............................ point.

I 3. Separation of the streamline flow around an aerofoil occurs when the flow
in the lower part of the bounda1y layer slows to a stop and begins
to ........................... .

I 4. As the angle of attack of an aerofoil is increased, the bounda1y


layer ..................... , the separation point on the upper surface
moves ........................... ; and form drag (increases/decreases).

I 5. Passing the stalling angle, the separation point moves rapidly (fmward/
rearward); a large turbulent wake is formed: and form drag (increases/
decreases) dramatically.

Principles of Flight Drag 5-17


16. Streamlining reduces form drag by making the curvature of surfaces more
gradual particularly toward the rear. The reason for the reduction in form
drag is that ............................. is delayed and the .................. .formed to the rear
is smaller.

17. If the form drag of a flat shape is rated at I 00%, this can be reduced by
streamlining to as little as ........ %.

18. The effectiveness of streamlining is determined by fineness ratio, which is


measured by ...................... divided by ........................ .

19. Which of the following factors affect form drag- streamlining; size; angle of
attack; airspeed?

20. The additional drag caused by the mixing of different airflows around an
aircraft is called ............................... drag.

21. Induced drag is present when the wings are producing ................... .

22. Whenever the downwash induced by the ·wingtip vortices is present, the
(geometric/effective) angle of attack must be increased if the same amount
of lift is to be produced. This results in a further rearward tilt of the TR force
with respect to the (effective airflow/flightpath) resulting in induced drag.

23. Aspect ratio (AR) has a (significant/insignificant) effect on the strength of the
vortices produced. Wings with a high AR generate proportionally (more/
less) induced drag than those with low AR.

24. Induced drag is highest when the angle of attack is (high/low).

25. Induced drag is (inversely proportional/proportional) to IAS 2 •

26. Induced drag is therefore greatest at (high/low) speeds and (high/low)


angles of attack.

27. State four design measures which can be taken to reduce the induced drag
on a wing of given AR.

28. Sketch a graph showing a typical Co versus angle of attack curve.

29. Write down the drag formula.

30. Draw a graph showing a typical cmve of drag versus airspeed for straight and
level flight. Indicate the stalling speed and the speed for minimum drag.

31. The lift/drag ratio can be obtained by dividing the coefficient of .................. by
the coefficient of... .................. .

32. Draw a typical curve of lift/drag ratio versus angle of attack. Indicate the
most efficient angle of attack, and the stalling angle.

33. In an alternative classification, drag can be divided into two groups- zero-lift
and lift-dependent drag. Zero lift drag does not contain ..................... drag.
Lift dependent drag is made up of .......................... drag and increments
of ............... drag, .............. -.................... drag and .......................... drag.

5-18 Drag The Commercial Pilot Series


Lift Augmentation
Introduction

In the design of aerofoils, there is no difficulty in providing for lift at high speeds.
In the cruise or at higher speeds, the wings are operated at a low angle of attack
(low CL) and the required lift is mainly derived from airspeed. The main
problem in designing aircraft for high-speed operation is in keeping parasite drag
as low as possible for efficient and economical use of engine power. Aerofoil
sections used for high speed aircraft are therefore characterized by having little
or no camber, a low thickness/chord (tic) ratio, and the point of maximum
thickness placed well back.

The high-speed aerofoil suffers however from having a low CLmax. Wings
constructed with this type of aerofoil therefore have a relatively high stalling
speed and provide little manoeuvring capability at low speeds.

An aircraft which is required to operate and manoeuvre safely at low speeds


must incorporate a high-lift aerofoil in its wing construction. This type of aerofoil
is highly cambered, has a high t/c ratio, and the point of maximum thickness
well forward. By comparison with the thinner high-speed section, it provides a
much higher CL at all normal operating angles of attack, and thus a lower stalling
speed and more manoeuvrability. The penalty paid is that much more parasite
drag is generated at high speeds.

Hence high-speed aerofoils have good drag characteristics at high speed, but a
poor lifting performance at low speeds. Low speed/high lift aerofoils have much
better lifting capability at low speeds but a poor drag performance at high
speeds. The GP aerofoil is aimed at compromising between the lift and drag
characteristics of the two.

Whatever the type of aerofoil section used, almost all aircraft wings have some
sort of device for lift augmentation at slow speeds and high angles of attack.
These devices aim to provide the advantages of high lift at low speeds, without
incurring the disadvantage of generating high drag at high speeds. There are a
number of ways in which this can be done, but the main devices used are flaps,
slats and slots.

Trailing-edge Flaps
Most aircraft are fitted with trailing-edge flaps-hinged trailing-edge surfaces
usually fitted on the inner sections of the wings.

Fig 6-1. Typical trailing-edge flaps fitted to light aircraft.

Principles of Flight Lift Augmentation 6-1


The flaps are lowered to change the shape of the aerofoil section over that part
of the wing to which they are attached. Usually, they are hydraulically,
electrically or mechanically operated and actuated by an electrical switch or
mechanical selector handle from the cockpit. Some flap systems have set
stages of operation, while others allow variable settings to be made within their
range of travel. All have some type of indicator which shows the amount of flap
selected, even if in some cases this may rely on the positioning of the flap
selector handle. When selected, the flaps on each wing should travel
simultaneously and symmetrically, and it is important that this be checked
visually before flight. In most aircraft, there are airspeed limits for operating the
flaps and for flight with flap extended which must be observed for structural
safety reasons.

All types of flap work on the principle of increasing the effective camber of the
aerofoil section over that part of the wing to which they are attached. When
lowered, the increase in camber results in a greater pressure differential being
generated above and below the wing. This effect is illustrated at Fig. 6-2 which
shows the type of trailing-edge flap called the simple or plain flap.

~---------------
--- ------
------ --- ---
~-
---~
-------=-----
c ?E -
~---------------
-----------------
~----------------

Fig. 6-2. Flaps operate by changing the effective camber of the aerofoil section.

Effects of Trailing-edge Flap


There are several effects of lowering trailing-edge flap:

Increased CL The main effect of lowering flap is that CL is increased over all
normal operating angles of attack. This means that at any given airspeed, more
lift is produced with flaps down than with flaps retracted. The higher CLmax also
results in a reduced stalling speed. With flaps extended, aircraft are more
manoeuvrable at low speeds; and can safely approach and land at lower speeds
resulting in shorter landing distances.

The way in which CL increases with different


increments of flap lowered is shown at Fig.
6-3. With the initial 2a - 3ao of flap travel, the
---
/? full flap

:'!? '"~,, ·~
30°
increase in CL is large, but this increase tails
off as further amounts of flap are lowered.
Simple flaps can achieve an increase of CL
up to about sao of deflection, but the gain in
I
lift over the last part of their travel (from
about 6a - sao) is marginal.

Note that the angle of attack for CLmax (the


critical or stall angle) reduces as more and __-LJ_ _ _ _ _ _ _ _ _+
more flap is used. This is discussed in more o· a
detail shortly.
Fig. 6-3. Lowering flap increases CL.

6-2 Lift Augmentation The Commercial Pilot Series


Increased C0 • Whenever flaps are lowered, C0 increases. As the drag is higher,
to maintain any given speed with the flaps lowered requires more thrust than
with them retracted. This increase in drag with flap is, however, of advantage on
the approach to land. It enables steeper approach angles with better forward
visibility and better obstacle clearance to be flown safely at a comparatively slow
speed. In most circumstances, the higher drag during the landing roll-out also
helps to reduce the landing distance.

steeper approach
at slower speed \ ,

. d' t
sho rt er Ian d1ng 1s ance , " t't[J-- -,;-
'\ h h t
1.- __ I..._f-_-_-_-_-_-_-_-=-_-_-_-_-_-_--::---:_-.:o--~-L.p!p s ah7g~:;~;~~~
11
a
longer landing distance

Fig. 6-4. The use of flap can provide for a steeper approach to be made safely at a slower
speed, giving better obstacle clearance and shorter landing distances.

Decreased lift/drag ratio. With a few exceptions, when flap is lowered on an


aircraft, the increase in C 0 is proportionally greater than the increase in CL. The
lift/drag ratio is thus reduced with flap extended. For most types of trailing-edge
flap, the greatest increase in lift is obtained over the initial part of flap travel. The
use of partial flap-in the order of 15 - 20' (sometimes variously called the lift
flap, optimum flap or take-off flap setting)-gives the advantage of increased lift
with a relatively unimportant increase in drag. The use of the full flap setting-
sometimes called the drag flap or landing flap setting-where there is a much
larger increase in drag for a small increase in lift, is normally restricted to the
approach and landing.

Reduced angle of attack. At any given speed, the angle of attack is lower with
trailing- edge flaps lowered, than with them retracted, and the stalling angle of
attack is lower. This is reflected in Fig 6-3 which shows that the CLmax occurs at
a progressively lower angle as the flap setting is increased. Note however, that
the angle of attack referred to is the geometric angle of attack in which the
original chord line is used as a reference. As is shown in Fig. 6-5, the aircraft will
fly level and stall at a lower nose attitude with flaps extended than it does with a
'clean' wing and this is what the pilot sees. Bearing in mind that the flaps do not
normally extend over the whole of the trailing edge of the wing, if an average
'new' chord line is drawn for the flaps extended position and used as a
reference, the wing will stall at a similar effective angle as the original geometric
'flaps retracted' angle.

flaps up flaps extended

reduced geometric a
Fig. 6-5. With trailing-edge flap extended, the aircraft will
fly level and stall at a lower angle of attack.

Principles of Flight Lift Augmentation 6-3


Rearward movement of the centre of pressure. When flaps are lowered, the
shape of the pressure envelope around the wing changes as shown in Fig. 6-5.
Most of the increase in pressure differential occurs toward the rear of the
aerofoil section, where the flaps are situated. As a result, not only is the TR force
increased, but the centre of pressure (CP) moves rearward. This movement of
the CP has a tendency to cause a nose-down pitch as the flaps are travelling. In
some aircraft types, this nose-down pitch is very evident. On others, the nose-
down pitch may be offset by other factors-this subject will be covered in
greater detail in a later chapter.

Fig. 6-6. As flaps are lowered, the


centre of pressure moves rearward.

Summary

When trailing-edge flaps are lowered:


• CL increases;

• Co increases;
• the l.jD ratio is reduced;
• the stalling angle of attack is reduced; and
• the CP moves rearward.

Note: The effects of operating trailing-edge flap on the pitch attitude of an


aircraft, and the tendency to 'balloon' or sink, are discussed in Chapter 9.

Types of Trailing-edge Flap


There are many types of trailing-edge flap. The characteristics of some of the
more common are covered below, i.e. simple, split, slotted and Fowler flaps.
The increase in CLmax which can be obtained varies from about 50% with the
simple flap to 90% with the Fowler flap. The gain in CLmax is generally achieved
at the expense of increased mechanical complexity.

The Simple (or Plain) Flap

The effects of lowering the simple flap have


been covered earlier. The effectiveness of the
simple flap is limited by the relatively early
separation of the boundary layer at moderate
to high angles of flap deflection. This limits
the CLmax which could potentially be y________
------- ............ ,
---- ', ~

achieved. In effect, the flap itself becomes


stalled, and this leads to a lower overall gain
----- ' ............
' ',,'
~-------- ........................

in lift and a large increase in drag. ----------~' ......~'---~'-:"":---

Fig. 6-7. The simple (or plain) flap.

6-4 Lift Augmentation The Commercial Pilot Series


----
The Split Flap -------------:~~----
In the split flap arrangement, only the lower surface of the
a..----------
...---
-
-------- ---
__
--
.................................. --
----::::::.:-:::---
_
......

.................
- ......
aerofoil section moves. Because the upper surface retains
the same shape, the early flow separation which occurs with
the simple flap is avoided. The gain in CLmax with split flap is
--,
.........................
............. .......

therefore higher than for the simple flap. However, as a


larger wake is formed, the drag generated by a split flap is
higher.
Fig. 6-8. The split flap.
The Slotted Flap

A slotted flap is a type of simple flap in which a slot is opened -----


up ahead of the flap when it is lowered. Air, moving from --------- :::::::::
high to lower pressure through the slot, accelerates around
the nose of the flap and is directed down over the upper
surface of the flap. This provides a form of boundary layer
control, in which the boundary layer is re-energized-given
greater kinetic energy. This delays separation of the
boundary layer enabling a higher CLmax to be achieved with
lower drag, than for a simple flap without a slot.
Fig. 6-9. The slotted flap.
The Fowler Flap

The Fowler flap is similar to the slotted flap, but in


addition to being deflected downward, the flap also
moves back. In addition to gaining the benefit of the
slot, the wing area is increased giving a further
increase in lift. Because of the geometry of this
arrangement, there is also a larger rearward
movement of the CP. The Fowler flap is the most
efficient of the trailing-edge flaps, giving the greatest
flap retracted
increase in CL for the lowest increase in drag.
Fig. 6-10. The Fowler flap.

A comparison between the typical gains made in CLmax with the use of different
types of trailing-edge flap, together with an indication of the drag produced and
the change in geometric stalling angle, is given at Fig. 6-11.

Fig. 6-11.

.....
o· a
+
drag generated

Principles of Flight Lift Augmentation 6-5


Slats and Slots
A slat is a small aerodynamic surface placed ahead of a
main aerofoil, to form a slot through which air can flow (Fig. Fig. 6-12.
6-12).

Some older aircraft had fixed slats fitted forward of most of the leading edges of
the wings, but this is not common today because of the drag which the fixed slat
generates at high speed. When they are fitted to modern aircraft, fixed slats
normally cover a short part of the wingspan at a particular site-e.g. at the
wingtips, or near engine nacelles-where they are used to control the local
airflow.

To avoid the drag penalty at high airspeed, most slats are retractable and open
and close automatically. At low angles of attack, the effective angle of attack of
the slat is negative and it is held firmly in the closed position. As the angle of
attack of the wing is increased it is accompanied by an increase in the upwash
over the leading edge. When, for these reasons, the effective angle of attack of
the slat reaches the point where it begins to produce 'positive' lift, it moves up
and forward on its tracks, opening up the slot. If the angle of attack of the wing
is again decreased, this process is reversed and the slat closes. The slat is thus
in operation when it is needed for lift augmentation at high angles of attack, and
closed to avoid the extra drag at low angles of attack.

Fig. 6-13. Automatic slats open when the required angle of attack is reached.

The principle of operation of the slat-or rather, the slot formed when it is
open-has already been covered. Air from below the wing accelerates through
the slot and becomes directed tangentially back along the upper surface, adding
kinetic energy to the boundary layer. As illustrated in Fig 6-14, the effects are:
• separation is delayed, and the pressure envelope over the upper surface is
'flattened out';
• the stall is delayed and will occur at a higher angle of attack; and
• CLmax is increased.

,...___ with slat, CL max ----~>­


increased, stall
delayed

~ ~,
plain wing~ ( \
\ \
\ ''
"
o· a

Fig. 6-14. Slats provide an increase in coefficient of lift and a higher stalling angle.

6-6 Lift Augmentation The Commercial Pilot Series


Leading-edge Flaps
A number of large aircraft and some high-performance aircraft are fitted with
leading-edge flaps, which are normally operated in conjunction with trailing-
edge flaps. The principle of operation is the same as for trailing-edge flaps-that
of increasing the camber of the wing. The effects of extending leading-edge flap
are the same as for the trailing-edge variety, except that if they are used on their
own:
• the stalling angle of attack is increased; and
• the CP moves forward.

There are a number of ways in which leading edge flap can be mechanised. Fig.
6-15 illustrates two-a type of 'drooping' flap, in which the whole of the leading
edge surface moves forward and down, and the Krueger type in which the lower
leading- edge surface hinges downward from the nose. Note that if the first type
of flap is extended further to form a slot, it becomes in effect a large slat. Unlike
automatic slats, the control of leading-edge flap is activated from the cockpit and
may be staged.

leading-edge droop Krueger flap

Fig. 6-15. Two types of leading-edge flap.

Combined High-lift Devices

Larger air transport aircraft are usually equipped with a cruise configuration
combination of the high-lift devices discussed thus far,
i.e. leading-edge flaps (or slats) combined with multi-
element slotted trailing edge flaps. A typical
arrangement is illustrated in Fig. 6-16 with the devices
shown retracted for the cruise; partially deployed for
take-off; and fully deployed for landing. In this latter
configuration, note the extensive use of slots to provide
take-off configuration
boundary layer control; obtain the highest possible
increase in CLmax.; and the lowest possible safe
approach and landing speed.

Fig. 6-16.
landing configuration

Principles of Flight Lift Augmentation 6-7


Spoilers
Although not a lift augmentation device (in fact, quite the reverse), we cover
spoilers in this chapter for convenience.

Most advanced jet transports and most gliders have spoilers on the upper
surfaces of their wings. These are hinged surfaces, which when extended,
disturb the airflow over the upper lift -producing part of the wing, thereby
decreasing lift and increasing drag. Glider pilots use mechanically-operated
spoilers to reduce airspeed and/or steepen their descent path without increasing
airspeed. On large jet aircraft, the spoilers are hydraulically operated and
deployed after touchdown to 'dump' the lift and increase the weight on the
wheels. This improves the effectiveness of the wheel brakes; increases drag
during the roll-out and, in some aircraft, improves directional control in strong
cross-wind conditions during the landing run.

Some jet transport aircraft also utilize differential spoilers to augment the
ailerons in the control of the aircraft in roll. When the spoiler on one wing is
deployed, that wing loses lift and the aircraft will roll in that direction. With this
arrangement, low rates of roll are controlled by the ailerons. When a higher rate
of roll is demanded by the pilot, the appropriate spoiler is deployed or partially
deployed to achieve the increased rate of roll.

6-8 Lift Augmentation The Commercial Pilot Series


Review 6

1. All types of flap work on the principle of increasing the


effective ........................ of the aerofoil section of which they are a part.

2. When flap is lowered:


(a) CL is (increased/decreased) resulting in a (higher/lower) stall speed.
(b) C0 is (increased/decreased).
(c) The LID ratio is (improved/reduced).

3. With trailing-edge flap lowered, the geometric stalling angle of attack is


(reduced/increased).

4. Sketch a split flap, a slotted flap, and the Fowler flap.

5. The effect of leading-edge slat is to:


(a) (delay/promote) separation.
(b) (reduce/increase) stalling speed.
(c) (reduce/increase) stalling angle.

6. An effect of lowering leading-edge flaps is that the aircraft will stall at a


(higher/lower) angle of attack.

7. Spoilers operate by (disturbing/smoothing out) the airflow over the top of the
wing, which ........................ the lift and .............................. the drag.

Principles of Flight Lift Augmentation 6-9


6-10 Lift Augmentation The Commercial Pilot Series
Flight Controls

Introduction

What an aircraft does at any given time in the air-whether it flies level, climbs,
descends, turns, etc-is determined by the pilot placing it in a certain attitude
with a certain power setting. The attitude of an aircraft is its position in flight
relative to the horizon-e.g. nose high/nose low; banked or wings level. Attitude
is not the same thing as angle of attaclc At different times, an aircraft can have
the same attitude but be at a different angle of attack and following a different
flightpath.

:!!! flightpath

Fig. 7-1. Same nose attitude, different angle of attack and different flightpath.

If an aircraft is placed in a certain attitude, its performance will be determined by


the power setting. Conversely, if a certain amount of power is set, the
performance which results will depend on the attitude being held by the pilot.
Hence, we often say about flying:

power + attitude = performance.

Power (or more specifically, thrust) is determined by what we do with the


engine controls, and in particular, the throttle(s). The attitude of the aircraft is
controlled with the main flight controls. These are surfaces which, when
deflected, alter the pattern of the airflow around the wings and the tail section,
causing changes in the aerodynamic forces that they generate. In this chapter,
we cover the effects of the flight controls (and associated controls) in changing
and maintaining the attitude of the aircraft in the air.

The Primary Flight Controls


To describe the attitude of an aircraft, or its position
in flight, the three mutually perpendicular reference
axes which pass through its centre of gravity (CG) are
used. Any change in aircraft attitude is about the CG
and can be expressed in terms of rotation about
these three aircraft axes.

• Rotation about the lateral axis is known as pitch;


• Rotation about the longitudinal axis is known as
roll; and
• Rotation about the normal axis is known as yaw. Fig. 7-2 The aircraft axes.

Principles of Flight Flight Controls 7-1


The conventional flight controls used to affect movement about these axes are:
The elevator (hinged to the trailing edge of the tailplane) which controls
pitching of the nose up or down, and is operated from the cockpit with fore
and aft movements of the control wheel.
• The ailerons (hinged to the outer trailing edge of each wing) which control
rolling of the aircraft, and are operated by rotation of the control wheel.
• The rudder (hinged to the trailing edge of the fin) which controls the yawing
of the nose left or right, and is operated by the rudder pedals.

power elevator
and
thrust

Fig. 7-3. The primary controls.

NOTE: Some aircraft are fitted with a control


column or 'stick', instead of a control wheel. The
control column serves exactly the same function.
Moving the control column f01ward or backward
operates the elevator; moving it sideways operates
the ailerons. In this manual, the use of the term
control column applies equally to the control
wheel.

Fig. 7-4. The control wheel and the


control column.

Other associated controls include:


• trim tabs (situated on the trailing edge of the control surfaces), and usually
operated by trim wheels or handles in the cockpit, and
• the wing flaps (situated on the inner trailing edge of each wing), and
operated by a manual lever or electrical switch. (Covered in the previous
chapter).

Principle of Operation
The aerodynamic principle of operation of conventional flight controls is the
same as described in the previous chapter for the simple flap. However, by
comparison with flaps (which change the camber of the wing in one direction),
control surfaces are hinged so that they change the camber of the parent
aerofoil in two directions. The maximum angle through which control surfaces
can be deflected is usually also much less than that of the flaps.

7-2 Flight Controls The Commercial Pilot Series


How Control in Flight is Achieved
When flying by visual reference, the pilot sets the required attitude with
reference to the horizon which he or she can see in the general area forward of
the nose. Once the aircraft has been held and has settled in that attitude for a
short while with the appropriate power set, the instruments inside the cockpit
are then checked to see if any adjustments are required to the attitude selected.
(For example, if level flight is required, the altimeter may be checked.) For most
aircraft at normal flying speed, quite small changes in the pitch attitude and in
roll can have a significant effect on the desired performance, hence flying
attitudes must be carefully selected and held if a reasonable degree of accuracy
is to be achieved.

Control in pitch, meaning movement of the nose up or down, is achieved by


forward and backward movement of the control column, which moves the
elevator. This alters the aerodynamic force produced by the tailplane which
rotates the aircraft about its CG to change the pitch attitude.

control column control column


back forward upward
__..,. ~ aerodynamic force

.,!.~ A]i
1

DOWN -~~~===o::::::::*:•
\ down
downward elevator
aerodynamic force

Fig. 7-5. The elevator is the main means of achieving control in pitch.

Some aircraft have a fixed horizontal stabilizer with a moving elevator, which
operates by changing the camber. Other aircraft have an 'all-moving' or 'all-
flying' tail in which the angle of attack of the tailplane is changed. Each design
however, serves the same purpose of providing control in pitch and longitudinal
stability.

__/ __/
fixed horizontal stabilizer all-moving tail
plus moving elevator (or "all-flying"/slab tail)

Fig. 7-6. Separate horizontal stabilizer plus elevator; all-moving (or all-flying tail).

The primary effect of moving the elevator is to pitch the nose up or down. If the
elevator alone is used, it does not cause the aircraft to roll or yaw. There is
therefore no seconda1y effect of using elevator. However, a consequence of
using elevator is to cause the airspeed to change. In normal flight attitudes,
pitching the nose down leads to a higher airspeed and, conversely, pitching the
nose up leads to a lower airspeed.

Principles of Flight Flight Controls 7-3


Control in roll, (and control of bank angle), is achieved by rotating the control
wheel or by moving the control column to the left or right. This moves the
ailerons, which are on the outer trailing edge of the wings. To roll to the left the
control wheel is rotated to the left. The left aileron goes up, causing a decrease
in lift from the left wing. At the same time the right aileron goes down, causing
an increase in lift from the right wing. The result is that the aircraft will roll to the
left. The roll can be stopped and controlled at the desired bank angle with the
control wheel/control column.

L
L

~
L
L

--li
t ... t 7 CG
,iii'

Fig. 7-7. Ailerons: one up, one down- produces rolling motion about CG.

The primary effect of moving the ailerons is to roll the aircraft. A secondary or
further effect is to cause yaw (movement about the normal axis). When the
aircraft is banked, the lift vector is tilted in the direction of the bank. If the
ailerons alone are used to roll the aircraft and no other control action is taken,
the tilted lift vector combined with the weight of the aircraft produce a resultant
force which will cause the aircraft to slip sideways toward the lower wingtip.
Once this slip occurs, the airflow impinging on the greater area of 'fin' surface
behind the CG will cause the aircraft to 'weathercock' in the direction of the
bank, i.e. to yaw. In summary, the sequence is: roll-slip-and then yaw.

r
roll

w
Fig. 7-8. The secondary effect of aileron is yaw.

Another secondary effect of aileron in the yawing plane is called adverse yaw.
This varies from aircraft to aircraft depending on design, and is generally more
pronounced when large aileron deflections are used to roll the aircraft quickly.

7-4 Flight Controls The Commercial Pilot Series


When aileron is applied to roll the aircraft, the camber
of the outer section of the upgoing wing is increased.
Both the lift and the drag of that wing are increased.
Conversely, the camber of the downgoing wing is
decreased with a resulting decrease of lift and drag.
This difference in 'aileron' drag between each wing
manifests itself as adverse yaw-or yaw in the opposite
direction to the application of aileron. Adverse yaw is
only present while the ailerons are deflected and, if no
action is taken with the rudder to correct it, will cause
the nose to yaw through several degrees (until it is roll to the left
balanced by the 'weathercocking' forces), and the
aircraft will skid. Once the aircraft has reached the Fig. 7-9. Application of aileron can
desired angle of bank and the ailerons are returned cause an adverse yaw.
more or less to a neutral position, adverse yaw is
considerably reduced.

Any tendency for adverse yaw and the resulting skid can be easily counteracted
by the use of rudder in the direction of roll. When rolled, most aircraft will
require such use of rudder coordinated with the use of aileron to prevent
skidding, with the rudder pressure being reduced when the roll is stopped.

The ailerons (or lateral control systems) of many t ..._lift reduced


aircraft are designed in such a way as to reduce or
counteract the effects of adverse yaw. In the Frise-type
aileron, for example, the nose of the aileron protrudes
below the wing when it is deflected upward thereby
increasing drag at the same time as the lift is reduced.
In this way, the drag of each of the wings is more
drag increased
evenly balanced when aileron is applied.
Fig. 7-10. The Frise-type Aileron design
can reduce adverse yaw.

Other methods for counteracting the effects of adverse yaw include:

• Differential ailerons. For a given movement of the control column, the


downgoing aileron is deflected through a smaller angle than the upgoing
aileron, thereby reducing the difference in drag and adverse yaw.

• Coupling of controls. On some aircraft, the rudder is coupled to the


ailerons so that when there is a given aileron control input, the rudder moves
automatically to counteract the adverse yaw.

• The use of spoilers. As covered in the previous chapter, spoilers are used
on some large aircraft to achieve control in roll. When used for this purpose,
the spoiler is deployed only on the downgoing wing to reduce its lift, and the
increase in drag on that wing offsets the tendency for adverse yaw. The
spoilers are usually used in conjunction with the ailerons or, in some cases,
as the sole means of lateral control at high speeds.

Principles of Flight Flight Controls 7-5


Control of yaw is achieved through the rudder, which is in turn operated by the
pilot using the rudder pedals. When the fore and aft axis of the aircraft is
correctly aligned with the direction of flight, the aircraft will be properly balanced
(or coordinated). A small coordination ball (or balance ball) on the flight
instrument panel is the prime indicator to the pilot that the rudder is being used
correctly. The coordination ball is normally kept centred using rudder
pressure-so that the aircraft is not slipping or skidding sideways. Some
examples of the use of the coordination ball are illustrated in Fig 7-11.

skid to the right correctly balanced slip to the left correctly balanced
right rudder required left rudder required
Fig. 7-11. The coordination (or balance) ball is the prime indicator of whether rudder is be!'ng used correctly.

The prim my effect of rudder is to yaw the aircraft. A secondary or further effect
is to cause a roll, since:
• strong yawing of the nose to one side will speed up the outer wing, which
will then produce fractionally more lift;
• once the aircraft begins to skid, the wing which is to the rear is slightly
shielded ('blanketed') from the oncoming airflow, resulting in less lift; and
• in aircraft with dihedral (wing tips higher than wing roots), the forward wing
in the skid has a slightly higher effective angle of attack resulting in more lift.

yaw
~
skid

outer wing
travels faster rear wing
blanketed dihedral effect (exaggerated)
by fuselage

Fig. 7-12. The secondary effect of rudder is roll in the direction of yaw.

Summary of the Main Flight Controls

Plane Axis CONTROL Primary Effect Further Effect

pitch lateral Elevator pitch - •


roll longitudinal Ailerons roll yaw

yaw normal Rudder yaw roll

* There is no 'further effect' of elevator. The airspeed change which follows


movement of the elevator is more accurately described as a consequence.
Moving the elevator has no effect on aircraft attitude other than in pitch.

7-6 Flight Controls The Commercial Pilot Series


The Effect of Airspeed on the Controls
Increased airspeed over the flight-control surfaces makes them more effective.
In most light aircraft, the cockpit controls are mechanically linked to the control
surfaces. In this type of system, the force required to move the controls (stick
force) increases with airspeed. The flight controls will feel firmer and less
movement will be required to achieve the same effect. At slow speed, the stick
forces are reduced and the control column and rudder must be moved through
greater distances to achieve the same response from the aircraft in pitch, roll
and yaw.

In larger aircraft, the flight control system is usually hydraulically powered. The
stick force is provided synthetically and the force required to move the controls
will generally remain constant regardless of airspeed. However, as for light
aircraft, larger control movements are needed to obtain the same aircraft
response at low speeds.

The Effect of Slipstream


The slipstream is the body of faster-moving air which
is accelerated rearward by the propeller. On single-
engined propeller aircraft, the slipstream usually
envelops the fuselage and flows back over the
empennage. To a degree depending on aircraft
design, the slipstream makes the elevator and rudder
more effective at high power settings and low
airspeeds. The ailerons (and on some high T-tail
aircraft, the tailplane/elevator) are outside the
slipstream and are not affected by it. The influence
of the slipstream on rudder and elevator
effectiveness phases out as speed is increased and/or Fig. 7-13. Slipstream increases
power is reduced. rudder and elevator effectiveness.

A further effect of slipstream is in changing the aerodynamic balancing force


developed by the tailplane when power is changed. As discussed in more detail
in Chapter 9, the tailplane usually develops a small downward force to balance
the aircraft . An increase in power and slipstream, as shown in Fig. 7-14,
strengthens the airflow over the tailplane and increases the magnitude of the
downward force, causing a tendency for the nose to rise. To prevent any
unwanted change in attitude, the pilot must stop the nose from rising with
forward elevator pressure as power is added.

low
thrust
aerodynamic greater
force aerodynamic
nose rotates force
upward about CG

Fig. 7-14. Increasing power usually causes a nose-up tendency.


(which can be opposed with elevator pressure).

Principles of Flight Flight Controls 7-7


Conversely, if the aircraft is trimmed with power on, a reduction in power and
slipstream will cause the nose to drop-which can be countered with back
pressure on the control column to hold the nose up in desired position.

Yet another effect of slipstream is the yawing effect which occurs when power is
increased or decreased in a single-engined aircraft. In travelling back from the
propeller, the slipstream 'corkscrews' back around the fuselage and strikes the
left side of the fin (for propellers rotating clockwise when viewed from the
cockpit-the usual direction of rotation in modern aircraft). This will tend to
yaw the nose of the aircraft to the left as power is increased, but can be easily
counteracted by applying sufficient right rudder to keep the coordination ball
centred.

(2) increased slipstream


on left side of fin

(for clockwise rotation of slipstream)

Fig. 7-15. Counteract the yawing effect of slipstream with rudder.

Conversely, if the coordination ball is centred and power is reduced, less right
rudder pressure (or possibly left rudder pressure) will be needed to keep the
aircraft balanced with the ball in the centre.

Summary of the Effects of Slipstream on the Controls

For most single engined propeller aircraft, as power is increased the stronger
slipstream:
• increases the effectiveness of the elevator and rudder, particularly at slow
speed;
• tends to make the nose rise; and
• yaw to the left.

Conversely, if power and the slipstream are reduced:


• the effectiveness of the elevator and rudder is reduced ;
• the nose drops; and
• tends to yaw to the right.

Unconventional Control Configurations


The layout of the control surfaces in some aircraft is different from the
conventional configuration described in the foregoing. These unconventional
configurations include the canard configuration, and the use of the vee (or
butterfly tail), elevons and !ailerons, where two of the functions of the elevator,
rudder and ailerons are variously combined in one control surface.

7-8 Flight Controls The Commercial Pilot Series


The canard configuration (Fig. 7-16a) employs a foreplane (as
opposed to a tailplane) in which the horizontal stabilizer is !'
placed ahead of the wings. The canard may have a separate canard
(or forep!ane)
conventional-style elevator, or be of the all-moving slab type
for control in pitch. Fig. 7-16a.

A design feature sometimes seen in light aircraft is the vee (or


butterfly) tail in which the function of the elevator and rudder
are combined. If the control column is moved fore and aft,
both control surfaces move up and down to provide control in
pitch. If rudder is applied, the control surfaces move
asymmetrically to provide control in yaw. A number of
advantages are claimed for this type of configuration,
including less drag, lower weight and better spin-recovery
characteristics. However, the need for a complicated Fig. 7-16b. The vee tail;
differential gearing mechanism between the elevator and
rudder controls must be seen as a disadvantage.

With elevons, the function of the elevator is combined with


that of the ailerons. This type of arrangement is sometimes
seen on delta-wing aircraft. Control in pitch is provided by
having the elevons move up and down in unison. Control in
roll is achieved by having the elevons move asymmetrically.
Fig. 7-16c. Elevons.

Tailerons also combine the function of the elevator with the


ailerons. In this arrangement, a slab tailplane is employed,
with the left and right halves of the tailplane able to be moved
in unison for control in pitch, or independently for control in
roll.
Fig. 7-16d. Tailerons

Trim Controls

An aircraft is in trim in pitch, roll and yaw when it maintains a constant attitude
without the pilot having to exert any steady pressure on the controls. An aircraft
which is properly trimmed is far more pleasant to fly than one which is out of
trim. It will virtually fly 'hands off and require control inputs only to manoeuvre
but not to maintain an attitude or heading.

The trim controls, which are operated from the cockpit by trim wheels, handles
or electrical trim 'buttons', are devices to relieve the pilot from having to hold
constant pressure on the control column or rudder. The main need for trim is in
pitch, and so all aircraft are fitted with elevator (or pitch) trim. More advanced
aircraft also have rudder trim (yaw) and some also have aileron trim (roll).

In the usual trim system, operation of the trim wheel or handle in the cockpit
varies the angle at which a trim tab on the trailing edge of the control surface is
set. The trim tab operates by creating a small aerodynamic force acting near the
trailing edge of a control surface which is used to hold the surface at the desired
angle of deflection. As shown in Fig. 7-17, the control surface will maintain its
angle of deflection when the moment created by the trim tab is equal and
opposite to the moment of the control surface itself.

Principles of Flight Flight Controls 7-9


trim force
Fig. 7-17. The control surface will f
hinge
maintain its position when its
aerodynamic moment (F x d) is
balanced by the moment of the trim
tab force (f x D).
main aerodynamic fore~
F
control force

The correct method of trimming is to hold the aircraft in the required attitude
with steady pressure on the controls and then trim this pressure off. The trim
controls operate in the natural sense-e.g. if steady elevator back pressure is
needed, then rotate the top of the trim wheel or move the trim control backward
to relieve the pressure. Similarly, if right rudder pressure is required, move or
rotate the rudder trim control (if fitted) to the right. The trim controls should be
moved gradually and steadily. Trim controls are very powerful and if they are
moved too quickly or suddenly this can result in a rapid change in aircraft
attitude and, possibly, over-stressing of the airframe. For the same reason, the
attitude of the aircraft should be changed with the main flight controls and not by
using the trim controls.

(1) hold steady pressure on (4) When trimmed, control


control column pressure is relieved
t
I
I

----- ~ )liJfr;~
~ / (3) trim tab

@------------------------// moves

(2) turn trim wheel slowly in same direction


as control pressure
Fig. 7-18. The correct method of trimming.

If the trim control is an electrical button (or switch), it will be spring-loaded in


the central OFF position. To remove a steady pressure being held on a flying
control, move the button or switch in the natural and instinctive sense until the
pressure is relieved-for example if f01ward control pressure is required, the
trim switch is held in the f01ward direction. When the control pressure is
relieved, the trim is released and should return to the OFF position.

Trim controls are very useful devices. Develop the habit of using them whenever
trimming or re-trimming is required. Flying an aircraft which is correctly
trimmed is much more comfortable and usually produces more accurate results.

7-10 Flight Controls The Commercial Pilot Series


Balancing of Controls
There are two types of balancing:

• Aerodynamic balancing, which is the function of facilitating the 'easy'


movement of controls by the pilot. It is achieved by designing the control
surface so that its centre of pressure is at an appropriate distance from the
hinge line.

• Mass balancing, which is the function of eliminating 'flutter' of a control. It is


achieved by arranging the distribution of the mass (or weight) of the control
surface so that its centre of gravity is at an appropriate distance from the
hinge line.

Aerodynamic Balance
When a control surface is deflected (for example, an up elevator-by holding
the control column back) the aerodynamic force produced by the control
surface itself opposes its deflection. This causes a moment to act on the control
surface about its hinge line (Fig. 7-19) which tries to return it to its original faired
(i.e. streamlined) position. The pilot must overcome this moment to maintain
the desired control position and feels this as stick force.

__/distance
Note that the aerodynamic force produced
by the deflected control acts through the
centre of pressure for the control surface.
The greater the distance between this centre
of pressure and the hinge line, the greater the
moment of the force which resists the pilot's
input and the higher the stick force. The
main reason for aerodynamic balancing is to
ease the difficulty with which a control can
be moved.
Fig. 7-19. The control force moment determines the
ease with which a control can be moved.

By altering the design of the control surface and the positioning of its hinge line,
it is possible to adjust the control force moment so that the stick forces are
neither too light or too heavy. Methods for providing aerodynamic balance
include the use of inset hinges, horn balance and balance tabs.

Inset Hinges

If the hinge line of the control surface is inset, as illustrated in Fig. 7-20, the
distance between the hinge-line and the CP of the control is reduced-thus the
control force moment (and the stick force) will also be reduced. In addition,
when the hinge line is inset in this way, the nose of the control surface protrudes
up (or down) into the airflow when the control is deflected. The acceleration of
the airflow around the nose of the control surface (when it is in this protruded
position) causes a decrease in pressure in that area, which results in the CP of
the control surface moving closer to the hinge-line, further reducing the control
moment. There is, in effect, an aerodynamic force acting on the control ahead
of the hinge line which helps to keep it deflected.

Principles of Flight Flight Controls 7-11


hinge-line
'

hinge-line
aileron I control
I surface

I
I I
I I

+--1----" inset hinges distance reduced


?

Fig. 7-20. The inset hinge line reduces the control moment and the stick
force required from the pilot.

Horn Balance

Horn balance is achieved when a control surface is designed with a portion


which protrudes ahead of the hinge line, as shown in Fig 7-21. The protruding
portion (the 'horn') can be shielded or unshielded. Both types operate on the
same principle as the inset hinge. As the horn is ahead of the hinge line, the
overall centre of pressure of the control surface is brought closer to the hinge
line and the control moment is reduced.

~ control surface
~ force

~-····.·····
··~
~ o__:,~, !';

I I
I I
If<;
distance reduced '

1
shielded horn - - - - - - , control surface
force

~ I I
I I
!f<;
distance reduced

Fig. 7-21. Shielded and unshielded hom balance.

With both the inset hinge and the horn balance, the designer must be careful not
to bring the centre of pressure too close to the hinge line. If this is done, the
stick forces may be too light, making the controls too sensitive and the aircraft
difficult to fly. In an extreme case, if the CP of the control surface moves fmward
of the hinge line, the stick force will be reversed and the control will have to be
forcibly prevented from going to full deflection of its own accord. This is known
as an aerodynamically 'overbalanced' control, which can be defined as one

7-12 Flight Controls The Commercial Pilot Series


which has its centre of pressure too close to the hinge line. At the other end of
the scale, a control which has its centre of pressure too far behind the hinge line
and which is difficult to manipulate, is known as an aerodynamically
'unbalanced' control.

Balance Tabs

On conventional tailplanes, a balance tab is sometimes incorporated as part of


the elevator. It is mechanically connected to the elevator by a linkage which
causes it to move in the opposite direction.

If the pilot exerts back pressure on the control column, the elevator is raised and
the balance tab goes down. As shown in Fig. 7-22, the elevator balance tab unit
now generates a small upward aerodynamic force which helps to move the
elevator up, thereby reducing the control load required of the pilot.

(1) back pressure (2) elevator (3) balance tab


on control goes up goes
column

creating-

main aerodynamic force (4) small aerodynamic


from horizontal stabilizer force which helps
and elevator move elevator up
Fig. 7-22. The balance tab.

NOTE: Although it is similar in appearance to the trim tab (Figs. 7-17 and 7-18)
the balance tab serves a different function as has been described. The trim tab
will only move when the pilot moves the trim control. The balance tab moves
automatically as the elevator is moved. If the aircraft is fitted with a balance tab,
its movement should be checked in the pre-flight inspection by moving the
elevator one way and noting that the tab moves in the opposite direction. On
more sophisticated aircraft, balance tabs may also be fitted on the ailerons and
rudder.

Anti-balance Tabs

In aircraft fitted with an all-moving tail ('all-flying' or 'slab' tail) the centre of
pressure of the control surface and its hinge line are relatively close. Compared
with the conventional tailplane/elevator arrangement, for a given deflection of
the control surface and a given stick force, the all-flying tail has the potential to
provide a much more powerful and stronger pitching force.

___ramie ~rce
'

CP small aerodynamic
force provided by
small hinge moment anti-balance tab

Fig. 7-23. The antibalance tab improves control 'feel' and prevents overbalance.

Principles of Flight Flight Controls 7-13


Because of the small hinge moment inherent in the all-moving tailplane, the
elevator control force tends to be light. There is also the possibility that the
control will overbalance-that is, when it is deflected, the centre of pressure of
the control surface could move forward of the hinge line giving it the tendency to
go to full deflection by itself. To avoid the possibility of overbalance, and to
provide better elevator control 'feel', all-moving tailplanes are often fitted with an
anti-balance tab, as shown in Fig. 7-23.

The anti-balance tab moves in the opposite direction to the balance tab
previously described (hence the name). If the trailing edge of the all-moving tail
is moved down, the anti-balance tab also moves down to provide a small
aerodynamic force to oppose the control column movement. Conversely, if the
trailing edge of the all-moving tail moves up, the anti-balance tab also moves up.
In this way, a well-designed anti-balance tab will provide the right amount of
'stick force'-i.e. the elevator will not feel too light or too heavy.

Movement of the tab is automatically provided through its linkage with the all-
moving tail. However, in most cases, the anti-balance tab is also linked to the
trim wheel so that it can be used as a trim tab as well.

Mass Balancing
All structures twist and bend (flex) under load-the fuselage and wings of an
aircraft are no exception. Indeed, the wings of most large aircraft are noticeably
flexible-with the wing tips bending upward by several metres as the aircraft
becomes airborne and the wings take up the aerodynamic loading. In flight, the
wings of large passenger aircraft can be clearly seen to flex up and down as the
load varies in turbulence. In smaller aircraft, twisting and bending of the wings
and fuselage may not be noticeable, but it will be present to some degree as
these structures can never be completely rigid.
aileron hinge CG

c t.-L
gust makes wing flex upward

Fig. 7-24.
Flexural aileron flutter.
(exaggerated for purposes
of explanation). t aileron la;s behind,
increasing wing flex

wing reaches elastic limit and begins


to spring back down

aileron lags behind, increasing downward flex


which continues to lower elastic limit
and the cycle repeats itself

7-14 Flight Controls The Commercial Pilot Series


This structural flexibility can lead to control 'flutter' at high speeds on some
aircraft. Flutter is a vibration or high-speed oscillation of a control surface
which, if left unchecked, can impose severe loadings on the wings and
empennage of an aircraft and lead to structural failure.

Consider a wing which is subject to a transient upward and downward bending


motion due perhaps to turbulence. Aileron flutter is liable to occur if the centre
of mass (CG) of the control surface is some distance behind the hinge line and,
because of inertia, it lags behind the rest of the outer wing in its movement up
and down. When this occurs, the changes in camber of the outer wing which
are caused by the 'lagging' aileron magnify the flexing of the wing caused by the
original oscillation and may make it self-generating-i.e. flutter. This process,
called flexural aileron flutter, is illustrated in Fig. 7- 24.
torsional axis

~~---~---!
A similar process called torsional aileron flutter can
occur if the wing is able to twist about its lateral axis.
In this case, the lag due to the inertia of the aileron, ~---~
<::.·:: -'\..:,:,l··~ ~-

causes the outer wing to twist alternatively nose up


and nose down (or oscillate) as shown in Fig 7-25. Fig. 7~25. Torsional aileron flutter.

Another effect which can occur if the wing twists about its torsional axis, is
called aileron reversal. As the aileron is moved down with the intention of
raising the wing, the upward force at the rear of the outer wing can cause the
leading edge to twist downward such that the overall angle of attack of the outer
wing is reduced. With this effect being mirrored by the other wing (angle of
attack increased), the aircraft will roll in the opposite direction to the pilot's input
of aileron. Aileron reversal is avoided by constructing the wing with sufficient
stiffness about the torsional axis, and through the use of spoilers to control roll at
high speeds.

The elevator and rudder control surfaces are also liable to flutter but, as they are
attached to shorter and stiffer structures, this is less likely than with the ailerons.
The flutter mechanism also requires that the control system linkages are
sufficiently elastic and 'stretch' enough to allow the control surface to deviate
from its normal position. Modern aircraft with hydraulically powered control
systems are less sensitive to control flutter because of the inherent rigidity of the
control system. (Although the hydraulic supply lines may be flexible, a constant
hydraulic pressure can be maintained at the control servos, which prevents the
control surface from moving).

Mass balancing must be applied to control systems which are prone to flutter.
This means that extra mass (weight) must be added to the control surfaces to
bring their respective centre of mass (CG) at, or close to, the hinge line. This
decreases the tendency for the control to lag behind through inertia when the
aircraft structure bends or twists. In aircraft which already have aerodynamic
horn or inset hinge balancing, this extra mass can be concealed within the
control surface in the horn or forward of the inset hinge line. If this is not
possible, the mass may have to be fixed externally on an arm which projects
ahead of the hinge line.
Fig. 7-26. Mass balancing. Extra mass is added where
indicated to bring the CG closer to the hinge line.

Principles of Flight Flight Controls 7-15


Review 7
I. A saying, often used in flying is:
.................................... plus ................................. equals performance.

2. Sketch an aircraft and draw in and name the three mutually perpendicular
axes which run through its CG.

3. Rotation of an aircraft:
(a) about the lateral axis is called ..................... :
(b) about the longitudinal axis is called ........................ :
(c) about the normal axis is called ......................... .

4. Conventional flight controls are:


(a) elevator, to achieve control in .................. :
(b) aileron, to achieve control in ..................... :
(c) rudder, to achieve control in ................. .

5. The secondary effect of aileron is to cause ......................... .

6. The secondary effect of rudder is to cause ........................... .

7. A consequence of using elevator is to cause a change in .......................... .

8. The adverse yaw which can follow the use of aileron, is a yaw in the (same
direction as/opposite direction to) the direction of aileron application.

9. List four measures used for counteracting or reducing the effects of adverse
yaw.

I 0. Increased airspeed over the control surfaces will make them


more ......................... and stick forces will be .......................... .

II. For most single-engined aircraft, as power is increased the stronger


slipstream
(a) (increases/decreases) the effectiveness of elevator and rudder;
(b) tends to make the nose (rise/drop); and
(c) yaw to the (left/right).

12. To trim out a stick force, the trim tab itself must be deflected in the (same
direction as/opposite direction to) the control surface deflection.

13. Aerodynamic balancing is achieved by adjusting the distance of the (centre


of pressure/centre of gravity) of the control surface from its hinge line.

14. State three methods of achieving aerodynamic balance.

15. The stick forces of an 'overbalanced' control will be (too heavy/too light).

16. An antibalance tab moves in the (same direction as/opposite direction to)
the deflection of the parent control surface.

17. Mass balancing is a matter of adjusting the CG of the control surface so that
it is (closer to/further from) the hinge line.

7-16 Flight Controls The Commercial Pilot Series


Stalling and Spinning
Stalling
Introduction

The stall is a condition of flight in which the angle of attack of the wing exceeds
its critical (or stalling) angle. The aerodynamics of the stalling process for an
aerofoil have been covered in detail in previous chapters. We have seen that
when a wing is taken beyond its stalling angle, the airflow breaks away (or
separates) from the upper surface and, as a result, there is a large reduction in
the lift produced. In simple terms, the wing operates beyond its angle of attack
for CLmax.

A fixed-wing aircraft is operated normally with an angle of attack from a few


degrees above the 'zero-lift' angle (usually associated with high speeds) up to
the angle for CLmax (usually associated with slow speeds in straight and level
flight/higher speeds when manoeuvring). Except for deliberate stalling practice
and some aerobatic manoeuvres, stalled flight is normally avoided because of
the loss of lift (usually involving a loss of height) and the possibility, in an
extreme case, of loss of control.

Most aircraft do not have a direct indication of angle of attack in the cockpit and,
as we shall see shortly, the speed at which an aircraft will reach the stalling
angle of attack depends on a number of factors. There is frequently a need to fly
the aircraft with the wings close to the stalling angle-for example, on take-off
and during the approach to land. To enable pilots to do this safely and
confidently, there is a need to understand and recognise the symptoms of the
stall so that the necessary action can be taken to avoid a stall if these symptoms
should appear. And, in the event that an inadvertent stall does occur, there is a
need for the pilot to be practised in, and be able to apply, the appropriate
recovery technique promptly and effectively.

The Stall and the Lift Formula


If an aircraft is to fly level, there must be sufficient lift to balance the weight.
When discussing the lift formula in Chapter 4, we saw that lift is a force made up
of three basic factors:
• the lift coefficient (CL) which depends basically on the wing section and
angle of attc.ck, and is a measure of how effective the particular aerofoil is in
generating lift;
• YzpV2 , the expression for the dynamic energy contained in the moving
airstream which, as explained in Chapter 3, is measured by the airspeed
indicator in terms of indicated airspeed (lAS); and
• the wing area (S)-which is conventionally taken to be the planform area of
the wings.

A given wing section (cross-sectional shape) has a certain value of CL at a given


angle of attack, which means that CLmax will occur at a fixed angle of attack for
that wing section. If the shape of the wing section is changed (for instance, by
lowering flap) then the CLmax will have a new value which will be found at a
different angle of attack, which then becomes fixed for that configuration. Note
carefully that the angle of attack for CLmax, for a given configuration, is fixed.

Principles of Flight Stalling and Spinning 8-1


The wing surface (S) is also a fixed value for a given wing configuration, e.g.
flaps up or down.

Thus, from a practical point of view, the lift formula can be expressed as:
Lift = CL. lAS . s
and, since S is for practical purposes, a constant, we can say that:
Lift = CL . lAS
Hence, as far as the pilot is concerned, the production of lift in any given aircraft
configuration depends only on the angle of attack and the lAS. For flight at a
constant height where lift must remain equally opposed to the aircraft weight,
the pilot may choose to fly at any lAS within the normal range obtainable, but
whatever value is chosen, it must be matched by one, and only one, angle of
attack (to provide the correct CL for the aircraft to remain level). If, for example,
it was desired to fly at a high lAS, the angle of attack must be relatively low-to
provide the low CL. Alternatively, if a slow airspeed was desired, the angle of
attack must be relatively high-to provide the high CL required.

Consider now a typical light aircraft in straight and level flight at a given speed
(Fig. 8-1 refers). If the throttle is closed, thrust will be reduced and the aircraft
will decelerate since the drag is no longer being overcome. To remain in level
flight, the pilot must match the reduction in airspeed with increases in CL (angle
of attack). Soon, an lAS must be reached where the associated angle of attack
reaches the value for CLmax. Any further reduction in airspeed can no longer be
matched with an increase in CL to compensate and the aircraft will stall. The lAS
at which that occurs is called the basic stall speed, which is defined as the lAS at
which an aircraft reaches its stall angle in straight and level flight while in the
'clean', power-off configuration.

100KIA~
4'"
I 80 KIASL
8' "
I 60 KIASL
12° a
r
~ ~ ~

J J J
Fig. 8-1. In straight and level flight, a reduction in /AS is compensated by an increase in CL
so that lift continues to equal weight.

Symptoms of the Stall


When approaching the stall, certain symptoms (or indications) will become
apparent to a degree which will va1y from one aircraft to another and from one
configuration to another. These are:

• Reducing airspeed. Note that we did not say 'low' or a 'reduction in'
airspeed because it is possible to fly an aircraft at a low speed for as long as
the fuel lasts, without necessarily stalling. But, if the airspeed is reducing
and continues to reduce, the stalling angle of attack must inevitably be
reached.

8-2 Stalling and Spinning The Commercial Pilot Series


• Reducing control effectiveness. Again note we do not say 'less effective
controls' because it is possible to have less effective controls and yet not
stall. But if control effectiveness is reducing and is allowed to continue
doing so, it is certain that the stalling angle must ultimately be reached.

The reduction in control effectiveness is allied to the reducing airspeed. It


occurs because the speed at which the airflow passes a deflected control
determines the lift force from that control, and through that, its
effectiveness. A consequence of reduced control effectiveness is the need
to increase the amount of control movement to produce the same effect as
the speed falls off.

• Buffet (or judder). As the wings approach the stalling angle, the
separation point begins to move more rapidly forward over the upper
surface and, as a result, there is a growth in the the turbulent wake behind
the wing. Some of this turbulent flow impinges on the aft fuselage and tail
unit causing a shaking-known as pre-stall or control buffet-which can be
felt through the airframe as well as through the controls. This buffet is a
good indication that the stalling angle is very close and, if the angle of attack
is further increased, that the wing is sure to stall.

The amount and intensity of pre-stall buffet varies considerably between


aircraft designs. For example, in high-wing aircraft, the wake will generally
pass 'over the top' whereas in low-wing aircraft, the wake envelops most of
the aft fuselage and control surfaces.

• The sink. Referring back to the CL cUlve in Fig. 4-2, you will note that the
peak of the curve (for the GP-type aerofoil) is well rounded and that for a
few degrees angle of attack before CLmax is reached, there is a significant
drop-off in the rate of increase in CL with angle of attack. This, coupled with
the reduction in elevator effectiveness at slow speed, generally means that
the angle of attack approaching the stall cannot be increased fast enough to
prevent loss of lift and some pre-stall sink taking place.

Once the aircraft sinks, there is a brief period where the nose attitude is
relatively high and the relative airflow becomes directed more from below.
This causes a rapid increase in angle of attack and, in effect, accelerates the
onset of the stall proper.

Other signs

The foregoing symptoms occur with all aircraft to some greater or lesser degree
as the stalling angle is approached and reached. There are some other 'tell-tale'
signs which may also accompany an approach to the stall but, as they may not
always be present, they cannot, technically, be called symptoms. They are:

• Lower noise level. When approaching the basic stall without power, there
is a reduction in the noise level in most aircraft. This factor is greatly subject
to individual perception.

• High nose attitude. In many stalls the nose attitude is indeed high, but as
the aircraft can be made to stall in any nose attitude, it cannot be taken as a
sure sign that the stall is being approached. For instance, when on a
descending approach to land with the relative airflow coming from below, it
is quite possible to stall the aircraft with a nose-low attitude.

Principles of Flight Stalling and Spinning 8-3


• Low airspeed. As will be explained shortly, it is possible to bring the aircraft
to the stall at an lAS higher than the basic stalling speed. In fact, in any
situation where the wings are producing more lift than is required for level
flight-such as in a turn or manoeuvring-the stalling speed will be higher,
and could even be close to cruising airspeed values.

• Stall warning devices. Most modern aircraft are equipped with devices
which warn the pilot of high angle of attack situations. These devices take
the form of warning lights (typically Piper aircraft), or a whistle noise which
increases in pitch as the angle of attack gets closer to the stalling value
(typically Cessna aircraft). In aircraft fitted with hydraulic control systems
which prevent buffet from being felt through the controls, 'stick shakers' will
normally be fitted to artificially introduce a buffet-type warning.

52 KIAS

<J~~~~
RAF 50 K!AS
a. exceeds 16"
'wake expands
buffet sets in
close to 50KIAS
aircraft sinks and
stall angle is exceeded

more than 50KIAS height


nose has dropped
loss

unstalled the aircraft

~
le•,elflighttsres~ ~~~ J
(power now re-applied)
RAF

Fig. 8-2. Sequence of events in the stall and recovery without power.

The Stall
At the stall, the separation point moves forward rapidly, the streamline flow over
most of the upper surface of the wing separates and a large turbulent wake is
formed. Underneath the wing the flow remains streamlined and there is a
relatively high pressure which provides some lift. However, as the low pressure
area over the upper surface collapses, net production of lift is sharply reduced.
With the formation of the larger turbulent wake, there is a substantial increase in
drag.

Approaching the stall, the centre of pressure (CP) will have moved gradually
forward to be at about 15 - 20% chord by the time the critical angle is reached.
As the flow over the upper surface breaks down at the stall, the CP moves
rapidly rearward to about the mid-chord position. This rearward movement of
the CP results in a nose-down pitch which tends to reduce the angle of attack
and unstall the wings. However, as the aircraft has at the same time lost lift and
is sinking with increased drag, the relative airflow is from below, which usually
results in the aircraft remaining above the stalling angle of attack-particularly if
the pilot continues to hold the stick back.

Thus the point of stall is characterised by the nose pitching down as the angle of
attack goes beyond CLmax. The degree of pitch down depends largely on:

8-4 Stalling and Spinning The Commercial Pilot Series


• The shape of the wing section. In general, the more rounded the leading
edge (typical of the high-lift aerofoil), the higher the nose attitude at the stall,
and the greater the amount of nose down pitch. Higher speed wing sections
with sharp leading edges can be expected to stall with a lower nose attitude
but with a more abrupt pitch-down.

• Configuration. Stalling with flaps down, involves a relatively low nose


attitude and the pitch-down motion at the stall is not large. The use of
power during the approach to the stall is associated with a higher nose
attitude (to be explained shortly) and the nose drop is more pronounced.

Very often, an aircraft will roll (or 'drop a wing') when the stall occurs. This is
caused by one wing reaching the stalling angle ahead of the other and can occur
for a number of reasons which will be discussed shortly.

Stall Recovery

Since a stall involves flight at angles of attack above that for CLmax, it follows that
the recovery must entail reducing the angle of attack to a value lower than
CLmax. This a achieved by fmward movement of the control column so that the
down elevator raises the tail and lowers the nose. The degree of fmward
movement of the control column which is necessary depends greatly on aircraft
type, stabilizer/elevator design and aircraft configuration. In general however,
the amount is not large, and can often be more accurately described as a
'relaxation' of back pressure on the stick.

Fig. 8-2 showed that fmward stick had lowered the nose sufficiently for the
relative airflow approaching the wing to reduce an angle of attack lower than the
stalling angle and normal flight (albeit in a descent) was resumed. Once
sufficient airspeed has been attained, the nose of the aircraft can be raised,
power applied and level flight re-established. Although recovery from the stall
has been successful, it has incurred a considerable loss of height and this is
clearly unsatisfactory should an unintended stall have taken place when height
above the ground is limited.

Reduction in height loss during stall recovery can be facilitated through the use
of full power. When the control column is eased fmward and power applied,
forward thrust accelerates the aircraft faster than is the case without applying
power and it then becomes possible to raise the nose of the aircraft at an earlier
stage. When the level flight attitude is regained, power is reduced to normal
cruise setting.

Note: Power does not recover an aircraft from the stall. This can only be
achieved by decreasing the angle of attack. The use of power merely helps to
reduce the height loss on recovery.

Interim Summary
If the amount of lift is to remain constant, the two practical variables in the lift
formula, CL and lAS, must work in the opposite sense if one or other is changed.

The CLmax is a fixed value which is reached at a given angle of attack for a given
wing configuration and the !AS at which CLmax is reached depends on the
amount of lift required. Thus, any alteration to lift requirement (for a given
configuration) does not affect the stall angle but does affect the stall speed. (We
will shortly be examining the factors affecting the stalling speed).

Principles of Flight Stalling and Spinning 8-5


To recover from the stall, the angle of attack is reduced with fmward movement
of the control column and height loss is reduced with proper use of power.

Factors Affecting Stalling Speed


Weight

For level flight, the amount of lift must be equal and opposite to the weight.
Thus a heavier aircraft requires more lift (more CL . lAS). So, at the stall, where
CLmax is a fixed quantity for a given wing configuration, it follows that only the
lAS can be increased to provide the greater lift requirement. Accordingly, the
stall angle is reached at a higher lAS. The stall speed therefore increases with
increase in weight and decreases with decrease in weight.

It is possible to calculate the new stall speed at different all-up weights by using
the following formula:

new stall speed = basic stall speed X Jnewweight


old weight

For example, if an aircraft's basic stall speed is 60 kt while at 1200 kg, what is its
new stall speed when the all-up weight is 2000 kg?

new stall speed = 60 X ~


=6ox0
=60X1·28

= 76·8 kt (77 kt)

L
L ---- ----
77 KIAS

angle of
attack

1200 kg
---- 2000 kg
w ----
w
----

Fig. 8-3. Stall speed increases with an increase in weight.

Load Factor

For an aircraft to manoeuvre, the wings must produce lift over and above that
required to balance the weight. For example, if an aircraft is to fly a level turn at
60" angle of bank, the lift produced by the wings must be twice as great as that
required to maintain straight and level flight with the wings level. The reason is
that to turn (or manoeuvre) the lift force must produce two components-a

8-6 Stalling and Spinning The Commercial Pilot Series


vertical component to balance the weight and the centripetal force component
directed toward the centre of the manoeuvre. Without the centripetal force, the
aircraft would simply fly a straight course, and without a vertical component,
gravity would take over and it would simply descend.

The measure of how much extra lift is being produced by the wings, is called the
load factor. Load factor is defined as the lift being produced at any particular
time by the wings, divided by the weight. It is expressed in terms of 'g' or
multiples of aircraft weight. Hence in straight and level flight, where the lift is
equal to the weight the load factor is 1'g'. In the example of the 60° turn, where
the lift is twice the weight, the load factor is 2'g'. (Load factor will be covered
again in more detail in Chapter 11, Turning).

The effect of load factor on the stalling speed is the same as for an increase or
decrease in weight. If the load factor is increased, the angle of attack of the
wing must be increased (at the same speed) to produce the extra lift required.
This means in turn that the angle of attack for CLmax must be met at a higher lAS,
and thus the stalling speed must increase with increased load factor. The stall
which can be made to occur at the higher speed when manoeuvring is called
the high-speed or accelerated stall.

The relationship between load factor and accelerated stall speed when
manoeuvring is similar to that expressed by the formula for a change of weight,
i.e.:
accelerated stall speed = basic stall speed x square root of the load factor, or
V (accelerated) = V (basic) x '</'g'
(Where V (basic) is the stall speed, straight and level in the same configuration
and at the same weight).

Hence, if an aircraft has a load factor of 2'g', its stalling speed will be 1·4 times its
basic stalling speed (since the square root of 2 is 1·4). If, for example, an aircraft
has a basic stalling speed of 50 KJAS, its stalling speed with 2'g' applied will be
1-4 x 50, or 70 KJAS. With 3'g' applied, the stalling speed will increase to
approximately 86 KIAS, and with 4'g' (outside the limits for most light aircraft)
the stalling speed is doubled, at I 00 KIAS.

I so KIAS 1\

~~~tr~g~~d_ _ _ _
·~ level

Fig. 8-4. The high-speed or


accelerated stall.

attempted 3g

-
intended flightpath
pullout

actual flightpath (stalled)

Principles of Flight Stalling and Spinning 8-7


Fig. 8-4 illustrates what would happen if this same aircraft (with a basic stalling
speed of 50 KIAS) attempted a 3'g' pullout from a dive at 80 KIAS.

Most light aircraft are not equipped with an accelerometer to give a read-out of
load factor in the cockpit. There is however, a direct relationship between load
factor and angle of bank in a level turn. This will be covered in more detail in
Chapter II.

Altitude

For straight and level flight, a given aircraft weight must be balanced by an equal
and opposite lift force, and this requirement applies no matter what the altitude
might be. We have seen that, for a given aircraft configuration, the amount of lift
gene1:ated depends only on the angle of attack (CL) and airspeed. Thus, in our
simplified version of the lift formula, for level flight L = W = CL . !AS. Hence,
regardless of altitude, at any given angle of attack and indicated airspeed the
same amount of lift will be generated. This means that the stall angle-the angle
of attack for CLmax will be reached at the same lAS in straight and level flight,
regardless of altitude.

If we now consider the stall speed in terms of true airspeed (TAS) it is a different
matter. You will recall that the expression for the dynamic energy in the moving
airstrearn-YzpV2 -is measured by the airspeed indicator as !AS. Consider now an
aircraft climbing at a constant lAS. As altitude is increased, density is reducing.
For the 1/zpV' function (that is lAS) to remain constant, V (theTAS) must increase.
Thus, although the stall !AS remains constant, as altitude is increased, the stall
speed in terms of TAS is increased.

This increase in the TAS at which an aircraft will stall at increased altitude is
significant to its performance. For example, if a take-off is contemplated at a high
altitude airfield, the TAS (the actual speed through the air) at which the aircraft
can become airborne with a safe margin above the stall !AS, is higher than at
lower altitude, and the required take-off distance is therefore increased.

Power

It was illustrated in Fig. 8-1 that as the !AS reduces in level flight and angle of attack
increases during the approach to the stall, the nose attitude becomes progressively
higher. If power is used during this phase, the thrust line is inclined upward and,
as shown in Fig. 8-5, there is an upward component of thrust which helps to offset
the weight. When the weight of the aircraft is being supported in this way (by both
aerodynamic lift and a component of thrust) the angle of attack required to
maintain level flight at any given lAS is reduced by comparison to that required
when the wings alone must provide all of the support for the weight. Thus, at any
comparable !AS, the angle of attack power-on, is lower than with power-off, and
the stalling angle is not reached until the aircraft has slowed to a lower !AS. In
other words, the use of power in the approach to the stall, reduces the stall speed.

An additional factor to consider, is the slipstream effect which is allied to the use
of power at the stall. Slipstream has the following effects which will be apparent
at the stall to a greater or lesser degree depending on the amount of power
applied and aircraft design:

• The increased speed of the airflow over the tail unit results in improved
elevator (and rudder) effectiveness. The improved elevator effectiveness
better enables the pre-stall sink to be counteracted resulting in a slightly
higher nose attitude at the stall proper.

8-8 Stalling and Spinning The Commercial Pilot Series


• In a single-engine aircraft, the slipstream flowing past the inner sections of
the wings, tends both to increase the speed of the airflow and reduce the
effective angle of attack in those areas. Insofar as lift is concerned, these
two factors tend to oppose one another resulting in no net change to the lift
produced. However, they do combine to delay separation over the inboard
sections which means that, power-on, the wing is more inclined to stall first
nearer the wingtips.

As a result of the above factors, for most light aircraft the power-on stall tends to
be more definite, with a greater tendency to wing drop. A further effect is that
with more thrust to oppose the drag, the approach to the stall is prolonged and
the symptoms are more easily identified.
increased slipstream L
over empennage
increases elevator
and rudder effectiveness

Fig. 8-5. The effects of power


at the stall.

w
Flap

In Chapter 6, it was explained that the lowering of flap increases the effective
camber of the wing; increases the value of CLmax and, with trailing edge flap,
reduces the geometric angle of attack (Fig. 6-3 refers). From our simple lift
formula (L =CL . lAS) we can see that if CLmax increases, lAS must decrease at
the stall angle if lift is to remain constant (to oppose an unchanged weight).
Hence we can conclude that the use of trailing edge flap will:
• reduce the stall speed (lAS); and
• reduce the stall angle of attack (which translates in level flight, to a lower
nose attitude).

Lowered flaps increase drag-particularly when fully lowered. In the approach


to the stall (without power), there is a faster reduction in speed and a shortened
period in which the symptoms are evident.

At the stall with flaps down and power applied there is often a greater tendency
for the aircraft to roll (drop a wing). This will be covered in more detail shortly.

Slats

The effect of slats has also been covered at Chapter 6. In short, with slats
extended, boundary layer separation is delayed and the CLmax is increased; the
aircraft therefore stalls at a slower speed and a higher angle of attack.

Slats typically produce very high nose attitudes at the stall and this carries the
risk of the tail striking the ground when the aircraft is operated close to the stall
and near the ground just prior to landing. This problem is normally overcome by
combining the use of slats with that of trailing-edge flaps, where the respective
effect of each device on the stall angle tend to cancel out.

Principles of Flight Stalling and Spinning 8-9


The use of slats alone has little effect on the sequence of events in the approach
to the stall-except that the high nose attitude should be apparent, and the
collapse of the intense low-pressure envelope above the wing, when it occurs,
tends to make the stall more abrupt.

Contamination of Wing Surfaces

When the wings are covered with ice, hoar frost, or such contamination as dirt,
dust or bird droppings, early separation of the boundary layer is encouraged-
particularly if the contamination is on the forward upper surface where most of
the lift is produced at high angles of attack. Even thin layers of frost or dust, can
result in a reduced CLmax being experienced (through earlier boundary layer
separation) at a lower angle of attack. Any reduction in the CLmax means that
the aircraft will stall at a higher speed than normal.

A further cause of an increased stall speed is the additional weight of the


pollution, and this is particularly so in the case of ice.

Any ice or frost at all, even if only the texture of fine sandpaper, should be
removed from the wing prior to flight. Similarly, it pays to remove any
accumulation of such things as insect remains or bird droppings from the wing.

Summary

The main factors affecting the stall (lAS) of a given aircraft are:
• weight-increased weight means an increased stalling speed;
• load factor-any manoeuvre which increases load factor, including turning,
increases the stalling speed;
• power-increased power decreases the stalling speed;
• flap (or slat) extension-decreases the stalling speed; and
• condition of the wings-any contamination increases the stalling speed.

While the stall lAS is not affected by altitude, the stall TAS increases with
increased altitude.

While the stall speed varies depending on the above factors, for a given aircraft
configuration, the stall angle remains the same. If the configuration is changed:
• by lowering trailing-edge flaps, the stall angle decreases and this is reflected
in a lower nose attitude in the basic stall; or
• by extended slats alone, the stall angle increases, giving a higher basic-stall
nose attitude.

Wing Drop at the Stall


Many aircraft exhibit a tendency to roll (or 'drop a wing') at the point of stall.
This is an undesirable tendency and, as we will explain shortly, recovery from a
stall in which a wing-drop occurs requires careful application of the correct
technique.

There is only one underlying cause for a wing-drop at the stall- that is, one wing
has reached its stalling angle ahead of the other. This can be caused by a
number of reasons, including:

8-10 Stalling and Spinning The Commercial Pilot Series


• Differences in the surface condition of the wings. The presence of unequal
amounts of ice, other contamination, or damage can result in flow
separation from one wing ahead of the other.
• Unbalanced flight approaching the stall. As was explained in Chapter 7,
when an aircraft with dihedral slips or skids, one wing develops (by design)
a higher effective angle of attack. If small amounts of slip or skid are
induced just as the stall is approached it can result in one wing reaching the
stall angle ahead of the other.
• Use of aileron near the stall. As will be explained in more detail shortly, the
use of aileron near the stall-can result in the outer sections of the wing
with the 'down' aileron reaching its stall angle ahead of that with the 'up'
aileron. This outcome from the use of aileron is also likely if the stall is
approached in unbalanced flight i.e. with 'crossed controls'.
• Approaching the stall during climbing or descending turns. As will be
explained in Chapter II, the angles of attack of each wing are different in
climbing or descending turns. Thus, as the stall is approached in such turns,
one wing reaches its stall angle ahead of the other and roll will result at the
stall.
• The use of flap approaching the stall. It is possible when flaps are extended,
that through mechanical wear and tear or slight manufactming differences,
they extend unevenly causing slight overall differences in the angle of attack
between the wings. If the flaps do extend evenly, their use nevertheless
tends to accentuate a wing drop for other reasons. This is because, with
flaps down, the effective angle of attack of the inner wing sections is
reduced (by comparison with the outer) and when the wing stalls it tends to
stall first nearer the wingtips. If one wing stalls ahead of the other in these
circumstances, the wing drop tends to be more sudden and have a faster
rate of roll.
• The use of power. As explained previously, when power is applied
approaching the stall, propeller slipstream tends to reduce the effective
angle of attack and delay the stall over the affected inboard sections. The
use of power, particularly in conjunction with flaps, tends to encourage
stalling of the outer sections first and accentuate any wing-drop.
Design Measures for Preventing Wing Drop

A stall which develops from the wing root first is desirable since, if a wing-drop
occurs, the rolling moment (and the rate of the ensuing roll) will be relatively
small. This characteristic is inherent in aircraft with a rectangular wing planform
shape (many training aircraft). Aircraft with tapered wing planforms are more
inclined to tip stalling, but the designer usually counteracts this by incorporating
washout-a lower angle of incidence toward the wingtips-as has been
previously described (Fig. 5-14 refers). Washout enables the inner sections of
the wing to reach the stall angle first, thereby reducing the tendency for tip-
stalling and a rapid roll if the wing should drop. In other aircraft, stalling of the
inboard sections first may be achieved by fitting flow strips (stall strips or
'spoilers') to the inboard leading edges as shown in Fig. 8-6. These strips
encourage early flow separation at the higher angles of attack.

Fig. 8-6. A flow strip.

-------- --
_____________
-- :::
Principles of Flight Stalling and Spinning 8-11
Use of Aileron Near, and During, the Stall
In spite of the design measures mentioned above, many training aircraft exhibit
a moderate tendency toward wing-drop, particularly if the stall is approached
with flap down and partial power on. This behaviour is quite normal and
satisfactory recovery can be achieved if the correct technique (to be described
shortly) is applied.

As we have seen, the wing drop at the stall is caused by one wing reaching the
stalling angle of attack slightly ahead of the other. When this occurs, the aircraft
will roll because of the imbalance of lift between the wings-i.e. the stalled wing
will 'drop'. As the roll develops, the airflow approaches each wing at a different
angle and, as a result, the angle of attack of the downgoing wing increases,
while that of the upgoing wing decreases (Fig. 8-7).
angles exaggerated for
purposes of explanation

upgoing angle
of attack

Fig. 8-7.
C velocity vector of
downgoing wingtip
relative airtlow downgoing wing downgoing angle
of attack

The increased angle of attack of the downgoing wing places it further into the
stall. By comparison with the upgoing wing, it therefore has less lift and higher
drag. The resulting lateral imbalance of lift and drag causes the aircraft to yaw
and to roll 'automatically'-to enter a state called autorotation. (Autorotation is
the basis of, and precedes the spin, and will be discussed in more detail in the
next section on Spinning.)

In addition, as the wing is dropping, since overall lift has been reduced in the
stall, the aircraft will pitch nose down.

To avoid the possibility of inducing a wing drop, large amounts of aileron


should not be used approaching the stall. The wing which has its aileron
deflected downward may be taken beyond the stalling angle which will cause a
wing drop.

higher lift

~
upgoing
wing
much less lift

~
--------
/ ...... stalled
less drag
'
I I

downgoi~'g-- ---A------'
wmg -<:J Aileron down
but left roll
and left yaw continue

much higher drag

Fig. 8-8. Avoid the use of large amounts of aileron near, or at, the stall.

8-12 Stalling and Spinning The Commercial Pilot Series


For the same reason, aileron should not be used as a wing drops at the stall in
an attempt to 'pick it up'. If aileron is used in this way, the angle of attack on the
downgoing wing will be further increased. This will take it further beyond the
stalling angle, adding to (not preventing) the unwanted roll and subsequent yaw.
The use of aileron to 'pick a wing up' at the stall can thus lead to autorotation, and
from there to a spin.

Recovery from the Wing-drop Stall


Note: The 'wing-drop stall' is sometimes referred to as the 'fully developed stall'.

When the wing drops at the stall, the first action-as it is with any stall recovery-
must be to reduce the angle of attack by easing the stick forward-straight forward.
In this process, the temptation to use aileron to pick up the dropping wing (an
instinctive thing to do) must be resisted. Then attend to the roll-yaw sequence and,
since aileron should not be used for the reasons explained, sufficient rudder is used
to prevent further yaw and so interrupt the snowballing autorotative sequence.

The angle of attack of the wings is normally reduced quite quickly by forward sticl{.
Thus a situation is soon met where the aircraft is unstalled, the rolling has stopped
(or almost so) and it is in a wing-down, nose-low attitude with increasing airspeed.
At this point (as the wings are unstalled) aileron is used, coordinated with rudder, to
roll the wings level at the same time as the nose is raised and full power applied to
bring the aircraft back to normal level flight. If flap was used in the approach to the
stall, it is raised in stages as the normal level flight attitude is regained. Once the
flaps are fully raised, the power may be reduced to the normal cruise setting.

The recovery sequence explained above, assumes that the nose of the aircraft had
not dropped greatly before recovery action was taken, in which event the use of full
power early in the sequence is appropriate. If the nose had been allowed to drop
well below the horizon, it is better to delay the application of power until the nose is
raised (after recovery) above the horizon.

Summarizing, the recovery from the wing-drop (fully developed) stall is:
• stick centrally forward;
• use sufficient ('top') rudder to prevent further yaw;
• apply full power;
• use ailerons smoothly to roll the wings level;
• raise the nose;
• centralize rudder.

In practice, these actions are carried out more or less simultaneously, and when
properly executed, the amount of height loss is not great.

If flap has been used, raise it in stages to recover to the cruise flight configuration
and then reduce power.

Recovery at Onset
The primary purpose of practising stalling is to be able to recognise the symptoms
leading up to the stall, so that it can be prevented-rather than 'recovered from'
once it has occurred. The most telling symptom on most aircraft is the pre-stall
buffet. In addition, there are the associated signs while in non-accelerated or level
flight of low airspeed, high nose attitude and activation of stall warning devices.

Principles of Flight Stalling and Spinning 8-13


For recovery at the onset of the stall when these signs are evident, the following
actions should be taken simultaneously:
• apply full power;
• stop the aircraft from rolling and yawing;
• lower the nose attitude to slightly above the horizon.

When the airspeed has shown a positive increase, the flaps (if extended) are
raised in stages and the aircraft returned to level flight at the normal cruise
power setting.

Note that the sequence of this recove1y is different from the full stall recove1y
where f01ward stick must always be the first action. The reason for the
difference, is that at the onset the aircraft has not actually stalled and, therefore,
full power application as the nose is gradually and smoothly lowered will bring
the aircraft safely back to n01mal flight with little or no loss of height. If the nose
was lowered before power is applied in these circumstances, an immediate loss
of height is incurred.

Spinning
A spin is a condition of stalled flight in which the aircraft follows a spiral descent
path about a vertical axis. As well as being in a stalled condition, the aircraft is
yawing and rolling with one wing producing more lift than the other. The spin is
an advanced stage of autorotation with the downgoing wing more deeply stalled
than the other. Greater drag from the stalled lower wing results in further yaw,
leading to further roll, and so on. Aerodynamic conditions may also fluctuate in
the spin, causing the nose to pitch up and down.

You can induce a spin deliberately by yawing an aircraft that is stalled, or is just
on the point of stalling. Once established in a spin, the aircraft is in motion about
all three axes. In other words, it is:

• stalled; q_
Spin aXIS
• rolling;
I
• yawing;
• pitching;
I up elevator

• slipping; and
• rapidly losing altitude at
a relatively low
airspeed.

relative airflow

Fig. 8-9. An aircraft in a controlled spin to the left.

8-14 Stalling and Spinning The Commercial Pilot Series


When established in a spin, the aircraft descends on a tight spiral path about the
vertical spin axis, with the wings at a large angle of attack. The total
aerodynamic reaction force from the aircraft is directed vertically upward and
counterbalances the weight acting vertically downward-in a sense, the aircraft
is parachuting downward. For each of the wings (even though they are in a
stalled condition) the total reaction force can be resolved into lift and drag
components acting perpendicular and parallel to their respective relative
airflows. With the outer wing spinning rapidly about the inner, the direction of
the relative airflow affecting the two is different, as shown in Fig. 8-9. The outer
wing is travelling faster, has a lower angle of attack than the inner, and thus
produces more lift. It is this difference in lift which produces the high rate of roll
in the spin. At its higher angle of attack, the inner wing produces more drag
which causes the continuous yaw. There will also be some side-slip present.

Autorotation
Autorotation is the basis of the spin. It occurs when a difference between the
angle of attack of the wings develops and when they are at, or beyond, the point
of the stall. The two main features to autorotation are:
• 'auto-roll'-with its higher angle of attack, the dropping wing is more deeply
stalled; it generates less lift and wants to keep dropping which causes the
aircraft to continue rolling; and
• 'auto-yaw'-the dropping wing also generates much more drag and wants
to continue yawing the nose of the aircraft in the same direction as the roll.

If a wing drops or a roll is induced in flight, perhaps by a gust or some action by


the pilot, the angle of attack on the downgoing wing is greater than that of the
upgoing wing (Fig. 8-7 refers). In normal flight at angles of attack well below
the stalling angle, the downgoing wing develops more lift through its higher
angle of attack-and the upgoing wing less. Hence there is a natural tendency
for any roll to be damped out, and for the aircraft to roll level of its own accord.

critical angle -..., no roll


equal lift

v
I
INCREASED
more
lift
DRAG
1 I
I I .1
I I
CL no roll I I
I I
equal lift I
j\ ~
more
lift rising
I I
wing 1I
DECREASED
I I LIFT

less dcag / : :
Co less lift'
i
f
II coil I
I I coli ) _.-----,.. 1
/ rising 1 _.-----,..1 1 ~I

~
wing . I / 1 . I DROPPING

l"<•"'"/" / 'I WING


w1ng / 1
tendency
tendency to keep
to recover / / I rolling
..- more drag I AUTO-ROLL
- _~:_::~--- I

normal in-flight range stalled fiight


Fig. 8-10. Lift and drag
effects on a dropping wing. angle of attack

Principles of Flight Stalling and Spinning 8-15


At the stalling angle or beyond, the increased angle of attack of the downgoing
wing causes it to stall (or to become more deeply stalled) and develop even less
lift. The result is that a dropping wing in a stalled condition will continue to
drop and the rolling motion will tend to continue of its own accord. This effect
is illustrated on the familiar CL curve in Fig. 8-10.

It can also be seen from Fig. 8-10, that compared with a roll in normal flight,
when a wing-drop occurs in stalled flight, the downgoing wing also experiences
a large increase in drag which yaws it in the direction of the roll. The strong yaw
leads to more roll, which leads to more yaw, and so it goes on-autorotation has
set in and the aircraft is about to enter a spin.

Autorotation can occur with only the dropping wing stalled and with the
(unstalled) rising wing producing considerable lift. Or it can occur with both
wings stalled as illustrated Fig 8.-1 0.

Spin Characteristics
Different aircraft will spin in different ways and it is not possible to describe a set
of spin characteristics which will apply to all. In general terms, most aircraft will
go through an early incipient stage of the spin (one or two turns) where the rate
of pitching, rolling and yawing may fluctuate and the airframe is subjected to
some buffeting. If control action is taken during the incipient stage, recovery will
be almost immediate.

After these first few initial turns, most aircraft will settle into a stable spin with
steady rates of pitching, rolling and yawing. With the wings more deeply stalled,
it is likely the spin will be 'flatter' (nose up, and more toward the horizon) with
little, if any buffeting. Spin recovery action at this stage, will probably require
about two full turns to take effect. Some light aircraft may not be able to achieve
this fully developed stage of the spin.

For a given aircraft type, a rearward CG will encourage a flatter spin making it
more difficult to lower the nose during recovery. If the aircraft spins with the CG
outside the aft limit, recovery may not be possible-a very good reason for
ensuring that the aircraft is never flown outside of its approved CG limits.

Conversely, a fmward CG normally results in a steeper spin with a higher rate of


descent and a higher rate of rotation. It may make recovery much easier and, in
fact, may even prevent a spin occurring.

Most light aircraft have to be 'held into' the spin for it to continue-i.e. have full
up elevator and full rudder held in the direction of the spin. In general, having
the flaps down and/or power on worsens the spin conditions and will make the
aircraft less inclined to recover. For most light aircraft, intentional spinning with
flaps down is prohibited.

Spin Confirmation
Most of the gyroscopic instrument indications are unreliable in a spin. However,
there are two instrument indications which can be reliably used to confirm the
fact and the direction of the spin. They are:
• a consistently low but fluctuating airspeed indication (if the aircraft is in a
low nose attitude but not in a spin, the airspeed will be high and increasing);
and
• the turn indicator which always indicates a turn in the direction of the spin.

8-16 Stalling and Spinning The Commercial Pilot Series


Spin Recovery
The recovery from a fully developed spin will be covered in detail by your flight
instructor. Recovery techniques vary between one aircraft type and another and
the following is not intended to replace proper flight instruction. In general, the
standard recovery techniques will include:
• checking the throttle is closed and the flaps are up;
• applying full opposite rudder to the direction of the spin;
• a pause (to allow the rudder to take effect in reducing the yaw);
• easing the control column centrally forward to reduce the angle of attack;
then
• as soon as the spin stops, smartly centralizing the rudder, levelling the wings
and easing out of the ensuing dive.
Some aircraft may take one or two further turns in the spin before the opposite
rudder and forward elevator have an effect in stopping the rotation. During
these one or two turns, the nose may go down and the rate of rotation will
increase temporarily. Others will stop spinning immediately these recovery
actions are taken, in which case you must be prepared to smartly start easing
out of the dive to avoid an excessive build-up of airspeed.

It is emphasised that the spin is a stalled condition of flight which develops from
autorotation. An unintentional spin c,an be easily avoided by:
• not allowing the aircraft to stall (or if it does, by taking prompt and correct
recovery action); and
• by avoiding the conditions which lead to autorotation-i.e. the coarse use of
aileron or rudder when the aircraft is approaching (or is at) the stalling angle.

Do Not Confuse a Spin With a Spiral Dive

A manoeuvre that must not be confused with a spin is the spiral dive, which can
be thought of as a steep turn that has gone wrong. In a spiral dive the nose
attitude is low, the wings are not stalled (unlike in a spin), the airspeed is rapidly
increasing and the rate of descent is high-a spiral dive is really just a steep
descending turn.

Because the wings are not stalled in a spiral dive, when recovering from it there
is no need to move the control column forward as you do when recovering from
a spin. To recover from a spiral dive:
• close the throttle (to reduce acceleration);
• smoothly roll the wings level with ailerons; and
• ease out of the dive with elevator (you can expect to feel increased 'g'-
loading); beware of a high-speed stall.

Principles of Flight Stalling and Spinning 8-17


Reviews

I. The stall is a condition of flight in which the angle of attack of the


wing .................. its critical (or stalling) angle.

2. For a wing in a given configuration, the angle of attack for CLmax is (fixed/
variable).

3. State four symptoms of an approaching stall.

4. The point of stall is characterized by the nose pitching (up/down).

5. For a GP aerofoil at the critical (stall) angle of attack, the centre of pressure
(CP) is at about ......... %chord. When the wing stalls, the CP moves rapidly
(fmward/rearward).

6. To recover from the stall, the angle of attack must be (increased/reduced).

7. Which of the following is correct? If weight is increased, the aircraft will


stall:
(a) at a higher speed but same angle of attack.
(b) at a higher speed but lower angle of attack.
(c) at the same speed but lower angle of attack,

8. When the aircraft is subject to a 'g' loading (i.e. it is being turned or


manoeuvred), its stalling speed in the manoeuvre = basic stall speed
X ................ .

9. If altitude is increased, the stall lAS (decreases/increases/stays the same)


and the stall TAS (decreases/increases/stays the same).

I 0. The stalling speed, power on, is (higher/lower) than with power off.

II. For a level-flight stall with trailing-edge flap lowered, the stall speed is
(higher/lower) and the nose attitude is (higher/lower) than with flap up.

12. Contamination on the forward upper surface of the wings, if it is the texture
of fine sandpaper, is (unlikely/likely) to increase the stalling speed.

13. The term used for a reduction of angle of incidence toward the wingtips
is ...................................... .

14. Write down the standard recove1y actions for a wing-drop stall.

15. In autorotation the dropping wing has (more/less) lift and (more/less) drag
than the rising wing.

16. Write down the two reliable instrument indications of a spin.

17. Write down the standard spin recovery technique.

8-18 Stalling and Spinning The Commercial Pilot Series


Straight and Level Flight
Introduction

There are four main forces acting on an aircraft in flight-lift, weight, thrust and
drag. These four forces are very similar to the forces which act on any surface
vehicle. The main difference is that the weight of a surface vehicle is supported
by the ground, whereas the weight of an aircraft in flight is supported by the
aerodynamic lift force mainly produced by the wings.

In the case of a surface vehicle-a car for example-the thrust force is provided
by the engine turning the wheels which then drive the car along, and the drag is
due to tyre friction and air resistance. With a propeller-driven aircraft, the thrust
is provided by the engine/propeller combination. The motion of the aircraft is
opposed by the drag force resulting from air resistance.

ground reaction lift

t
I
I
~
drag~ L!P~thrust drag ~~==~?~~??~thrust
~<¢>

weight weight

Fig. 9-1. The four main forces acting in flight.

The driver of the car controls the 'thrust' developed by the road wheels by
va1ying foot pressure on the accelerator. The pilot of the aircraft controls the
power developed by the engine and the thrust of the propeller by moving the
throttle forward or back with his or her hand (or in the case of an aircraft with a
variable-pitch propeller, by varying manifold pressure and/or propeller rpm).

In this, and the next two chapters, we will be covering the forces acting in flight
under the headings Straight and Level Flight, Climbing and Descending, and
Turning.

Principles of Flight Straight and Level Flight 9-1


The Forces Acting
In steady straight-and-level flight, the aircraft is maintaining a constant altitude,
lAS and heading-i.e. it is not climbing or descending, accelerating or
decelerating, or turning. For these conditions to apply, the forces acting on the
aircraft must be in equilibrium. The lifting forces must counterbalance the
weight (otherwise, the aircraft will gain or lose altitude) and the thrust must be
equal and opposite to drag (othe1wise the aircraft will gain or lose speed). In
coordinated (or 'balanced') flight, the wings must also be level so that the
aircraft does not turn.

lifting forces
counterbalance weight

Fig. 9-2. The conditions for oD


unaccelerated level flight.

Of the forces which go toward supporting the weight in level flight, some lift may
be provided by the fuselage and other parts of the aircraft other than the wings.
This may include lift from the tailplane, although it will usually be generating a
small amount of downward 'negative' lift which, in effect, adds to the weight of
the aircraft and requires more lift from the wings. (Such negative tailplane lift is
called 'trim drag').

In addition, the thrust line of the engine will rarely be exactly parallel with the
direction of flight. In the normal cruising attitude, the thrust line will be ve1y
close to the direction of flight, but will become increasingly inclined upward as
speed is decreased and the angle of attack increases. This upward inclination
provides a small vertical component of the thrust which helps to support the
weight. In level flight, the vertical component of thrust does not normally make
a significant contribution to supporting the weight unless the nose attitude is
comparatively high (e.g. approaching the stall) and a high power setting is used.

The wings provide by far the greatest contribution to counterbalancing the


weight in level flight and it is normal to disregard any lifting effect of the fuselage
and inclined thrust line. Hence, although it is not strictly cmrect for all
conditions of level flight, for practical purposes it can be taken that the only force
which balances the weight is wing lift and that the thrust line will remain the
same as the direction of flight.

wing lift
Fig. 9-3. The 'lifting forces'.
For practical purposes, wing
lift may be taken as the only fuselage and
force balancing the weight in other lift
level flight.
tailplane lift
(usually negative)

9-2 Straight and Level Flight The Commercial Pilot Series


Pitching Moments
In flight, the four main forces do not all act through the same point. Weight acts
through the CG and the position of the CG can change-from flight to flight
depending on the aircraft loading, and during flight as fuel is consumed (and in
large aircraft, when passengers or freight move!).

The lift acts through the centre of pressure (CP) of the wings. As you will recall
from Chapter 4, the position of the CP changes with angle of attack, generally
moving forward as angle of attack increases, and then rearward as the stalling
angle is exceeded.

With these changes, the CP and the CG of an aircraft rarely act through the same
line. The usual design arrangement is to have the CP behind the CG. As a result,
the lift/weight forces set up a couple which causes a nose-down pitching
moment.

Similarly, the thrust and drag forces rarely act through the same line and form
another couple which can cause a nose-up or a nose-down pitching moment
depending on the arrangement of the forces. Usually, the thrust line acts below
the line of drag and a nose-up pitching moment results.

Aircraft are normally designed so that the pitching moments of the lift/weight
(L/W) and thrust/drag (T/D) couples oppose one another. They are, however,
rarely in perfect balance and it is the function of the tailplane to provide the
necessary balancing (or stabilizing) force as shown in Fig. 9-4.

LIW couple
~
lift I nose-down
ij moment
pitching

I
I
CP I
t
tailplane dr
I
~
stabilizing T/D couple
moment

~ I
I
I
11
1I .
we1ght

Fig. 9-4. The pitching moments in flight.

As is normal, and in the interests of keeping things simple, most of the following
diagrams in this manual show the forces as if they act from a common point,
and the balancing force of the tailplane-which is usually always present but
relatively small-is omitted.

Principles of Flight Straight and Level Flight 9-3


Variable Effects on the Couples
If an aircraft is trimmed for level flight and there is a change in the magnitude or
line of action of any of the forces, the strength of the respective couple will
change and cause a nose-up or nose-down pitching moment. The more
significant of these effects occur through changes to the power setting, and
changing the configuration by raising and lowering the flaps and undercarriage.

Changes in Power

An increase in power (and thrust) increases the strength of the T/D couple
resulting in a nose-up pitch. Conversely, a reduction in power and thrust
weakens the T/D couple and results in a nose-down pitch. These changes in
nose attitude are easily countered with the use of elevator as power is changed.

dra
.:::-::-=-=-::s;-:_-=-:-:;_·II>
--------- thrust
)"-"'"""'
Fig. 9-5. Increasing power and thrust causes a nose-up pitch (and vice-versa).

Lowering and Raising the Flaps

When the flaps are lowered:


• lift increases and the CP moves reatward, resulting in an increased nose-
down pitching moment;
• drag increases and the line through which the total drag acts is usually
changed-raised with a high-wing aircraft resulting in a stronger nose-up
pitching moment, and lowered with a low-wing aircraft resulting in a nose-
down pitching moment;
• depending on design, in some aircraft, the airflow over the tailplane is
altered. If the downwash is reduced, a stronger nose-down pitching
moment generally results. If the downwash is increased, the nose will
generally tend to pitch up.

~LIW

~LIW
lift vector increases
and moves back -
+1 couple
lift vector increases A couple - nose-down pitch 1
and moves back - T
+ nose-down pitch I I
I I
I drag vector increases
I
I and moves up
I T/D
I - nose-up Pitch
I couple

~~\~'---p--~
increased downwash
TID
may cause
couple nose-up pitch

low-wing aircraft I w I high-wing aircraft I w

Fig. 9-6. Typical effects of lowering flap on the balance of couples.

9-4 Straight and Level Flight The Commercial Pilot Series


Hence, whether the nose pitches up or down (or indeed pitches at all) when the
flaps are lowered depends on which of the various moments predominates. In
general, most aircraft tend to pitch nose down when the flaps are lowered and
nose up when they are raised. Most high-wing training aircraft, however, give
the opposite response and pitch nose-up when the flaps are lowered. As it
varies from type to type, it is not possible to be more specific. The strength of
any pitching moment experienced will depend on lAS, the rate of flap extension
and retraction, and the amount of flap used. As the flaps are often lowered or
raised when flying at relatively low levels it is good airmanship to anticipate and
deal promptly with any unwanted pitching moment arising from this cause.

Another effect of lowering and raising the flaps can be 'ballooning' or 'sinking'.
Ballooning is a transitory tendency to gain altitude (or have a reduced rate of
descent) when the flaps are lowered. It arises from the relatively sudden gain of
lift as the flaps are being lowered. It will be more pronounced at the higher
speeds in the flap operating range (where, with the faster airflow, the increase in
lift is greater) and will be particularly noticeable if a large amount of flap is
lowered quickly. Such ballooning as is normally experienced is not of great
consequence and may not even be noticeable in a light aircraft when flap is
lowered on the approach to land.

The sinking which can result when the flaps are raised quickly is a different
matter-particularly if this is done at slow speed and at low level. When the
flaps are raised lift is reduced and, if the same nose attitude is held, height will
be lost (or rate of climb reduced) until the aircraft is able to accelerate to
recover the loss of lift through increased speed. In addition, as the flaps are
raised, the stalling speed increases. If sinking is experienced near the ground,
there will be a natural tendency for the pilot to raise the nose attitude to counter
it, which could well result in the aircraft stalling. For these reasons you should
avoid raising the flaps when flying at low levels and slow speeds.

Raising and Lowering the Undercarriage

For retractable-undercarriage aircraft, raising the undercarriage usually causes a


nose-up pitch, while lowering it will usually result in a nose down pitch.

Summary

The response in terms of the change of pitch attitude for most aircraft to changes
in power, flap and undercarriage configuration occurs in the natural sense, i.e:
• power up-nose up;
• flap up-nose up;
• undercarriage up-nose up.

And, in the opposite direction:


• power down (or reduced)-nose down;
• flap down-nose down; and
• undercarriage down-nose down.

An exception to the foregoing is that, with most high-wing training aircraft, when
the flaps are raised, the nose may go down, and when the flaps are lowered, the
nose may come up.

Principles of Flight Straight and Level Flight 9-5


Increasing and Decreasing Speed in Level Flight
To accelerate in level flight, thrust must be greater than drag. Conversely, to
decelerate, thrust must be less than drag. The driver of a car can control speed
simply by increasing or decreasing foot pressure on the accelerator. In flight,
speed is controlled differently. As will become clearer when we discuss
climbing and descending, the speed of an aircraft is determined fundamentally
by its attitude and the altitude (or height) by the amount of power applied.

If the aircraft is trimmed in level flight at constant speed, the forces will be in
equilibrium-the total lift will balance the weight and the thrust will balance the
drag. In addition, there will be no net nose-up or nose-down pitching moment.
Under these conditions, if the pilot increases power with the intention of
accelerating to a higher speed but without moving any of the other controls, the
aircraft will rotate nose-up and gain height. The speed may also increase, but
this will depend on the nose attitude which results. Conversely, if power is
reduced with the intention of slowing to a lower speed, the nose will drop and
the aircraft will descend. The speed may decrease (quite possibly, it will remain
constant) but once again this will depend on the nose attitude which results
from changing the power.

Under the same conditions just described, if the pilot increases power but
prevents the nose from rising with elevator (i.e. holds the same attitude), the
aircraft will still climb as the speed increases and the wings gain more lift. The
converse applies when power is reduced and the same nose attitude is held.
The speed will reduce and the aircraft will descend as lift is reduced at the lower
speed.

There is thus an interdependence between height and speed on the one hand,
and power and attitude on the other. This interdependence is summarized in
the adage 'power plus attitude equals performance', which can be universally
applied to all phases of flight. Applied to straight and level flight it means that a
change of performance-i.e. to fly at a faster or lower airspeed-must be
brought about by controlling both the power and the nose attitude.

Hence, to fly at a faster speed but remain in level flight, the nose must be
prevented from pitching up at the same time as power is applied and, while the
aircraft is accelerating, the angle of attack must be progressively decreased
(nose attitude lowered) to keep the lift equal to the weight The rate at which the
nose attitude must be lowered depends on a number of factors including the
amount of power applied, aircraft speed and the rate at which it accelerates.
The proper coordination of elevator control with the use of throttle and
increasing speed comes with practice and experience. The actions required are
summarised in Fig. 9-7 below.
lift
to remain in level flight,
lower the nose attitude
as power is increased
and as the aircraft
is accelerating

drag......._~:::::::~;;,;;;'!~~ thrust
I I
I I
Fig. 9-7. Increasing speed in
It I
level flight. increase power
to increase thrust

weight

9-6 Straight and Level Flight The Commercial Pilot Series


As the airspeed increases, the resisting force (drag) will also increase and
eventually, when the drag builds up to be equal to the thrust, the system of
forces will be in a new state of equilibrium. At the new, and higher speed, the
nose attitude for level flight will be lower.

The converse applies to a reduction of speed in level flight. As the power is


reduced and while the aircraft is slowing down, the nose attitude must be
progressively raised. With the decreasing speed drag is reduced and, if the
reduction in power has not been large, the aircraft will settle at the new and
lower speed when the thrust and drag forces come once again into balance. At
the lower airspeed, the aircraft will have a higher nose attitude.

lift
to remain in level flight,
raise the nose attitude
as power is reduced
and as speed
is decreasing

drag+--~~ thrust
~II
Fig. 9-8. I I
Reducing speed in level flight. It I
reduce power
to decrease thrust

weight
As you will be aware from our previous consideration of drag (Chapter 5), below
the minimum drag/maximum UD ratio airspeed, drag begins to increase again
as speed is reduced. A large reduction in power may result in the aircraft
decelerating into this region and if it is desired to fly level at speeds well below
the minimum drag speed, the power will have to be increased again (and
adjusted as required) to maintain the speed. The aircraft will otherwise
continue to decelerate until the stalling angle of attack is reached and level flight
will no longer be possible.

From the foregoing, it is also clear that there is a direct relationship between
speed, nose attitude (angle of attack), and power (thrust). In level flight, a high
speed is associated with a low nose attitude-and a low speed with a high nose
attitude. If the aircraft is to remain in level flight when the speed is changing, the
use of elevator to control the nose attitude must be properly coordinated with
changes of power and speed.
~--li>- increasing lAS -lower nose attitude required--+~

Fig. 9-9. lift lift lift


Maintaining
level flight.

~--- ~dG'~--
¢
slow faster much
speed faster

weight weight weight

+--+-- decreasing lAS- higher nose attitude required+--+--


Principles of Flight Straight and Level Flight 9-7
Performance in Straight and Level Flight
The thrust required for steady (unaccelerated) straight and level flight at any
given indicated airspeed is of course equal to the total drag at that speed (T = D)
Thus the curve of 'thrust required' versus !AS for a given aircraft (Fig. 9-1 0) is
exactly the same as the drag curve versus lAS for that aircraft.

'thrust stall
I
required' I
= I
total I
I
drag I
curve is for level
flight and valid for
constant aircraft
I minimum thrust I I
weight and altitude

I '
Fig. 9-10.
I ''
The 'thrust required'curve. t v'

i
Slow Fast
lAS
(high a) (Iowa)

speed for minimum drag


and litudrag max.

Compare this cwve with the drag curve at Fig. 5-1 7 and note:
• High thrust is required at high speeds to overcome what is mainly parasite
drag.
• High thrust is also required at low speeds to overcome what is mainly
induced drag.
• The minimum thrust required for level flight occurs at the minimum drag/
best 1)0 ratio speed.
• Over a section of the curve on either side of the minimum thrust speed, the
aircraft can be flown with the same thrust at two different speeds-one
higher than the other. Two such speeds are annotated A and B on the cwve
and we will refer to these speeds again shortly.

Rather than refer to performance in terms of thrust, it is more appropriate and


relevant to consider the power required and power available in level flight.
The use of these parameters leads to the determination of a number of factors in
aircraft performance including maximum level flight speeds, absolute ceiling,
range and endurance, and best rate and angle of climb speeds.

The Power Required Curve


The power required to move an aircraft through the air at a constant speed can
be determined by multiplying the thrust required by TAS-i.e. the force applied
x distance over time. In level flight, as thrust = drag, we can say that:
power required = drag x TAS.
The power required (PRl curve can therefore be easily obtained by multiplying
drag x TAS and plotting the results against TAS. Shown in Fig. 9-11 is a
comparison between a typical PR cwve and the drag curve for the same aircraft.

9-8 Straight and Level Flight The Commercial Pilot Series


drag
'thrust stall
required' I
I
= I
total I PR
drag I

Fig. 9-11.
A comparison between the
drag curve (thrust required)
and the power required curve.
power
required
curves are for level
PR flight and valid for
constant aircraft
weight and altitude

TAS

minimum drag I
max UD speed

Points to note with the above graph are that:


• Although the drag (and therefore, thrust required) is the same at points A
and B, the power required at those two different speeds is different. (Note
the points at which the vertical dashed lines intersect the PR curve.) Higher
power is required at the higher TAS. This serves to illustrate the important
fact that power and thrust are not the same thing-Nevertheless, when
we speak of 'increased power' we probably mean 'increased thrust' and in
general terms it will be true-an increase in power generally leads to an
increase in thrust and vice versa. We just have to be careful in distinguishing
between the two when it comes down to questions of performance.
• The bottom of the drag curve (i.e. minimum drag) does not coincide with
the bottom of the PR curve (i.e. minimum power). Hence, the minimum
power speed is lower than the minimum drag/ best UD ratio speed.

• If the PR curve is now considered by itself, the speed for minimum drag/ best
LJD ratio occurs where a line drawn from the origin of the graph is tangential
to the curve (see Fig. 9-12). This point on the PR curve is the best
aerodynamic speed for flying for range, as we will discuss in more detail in
Chapter 15.

Fig. 9-12.

power= drag x TAS


power
power required
drag=--
TAS PR

For minimum drag, the ratio of power to TAS, must be at a minimum.


This occurs on the power required cuJVe where the line from the
0
origin is tangential. At any other point, the ratio of power to TAS,
i.e. ratio of A to B, or Tan El is higher. B TAS

Principles of Flight Straight and Level Flight 9-9


The Power Available Curve
Aircraft piston engines are rated in horsepower-the unit used in the old 'Imperial'
system to measure power. The power output of the engine to the propeller shaft
can be fairly readily and accurately calculated at different rpm. At any given rpm
this output is relatively independent of TAS-although as speed is increased there is
generally an increase in efficiency and power output owing to the effects of ram air
pressure in the induction system and lower exhaust pressures. A typical curve of
power delivered to the propeller (Fig. 9-13) is usually one which is fairly flat but
reflecting some increase in engine efficiency and power available as TAS increases.
power delivered
to propeller

power available
from propeller

power
Fig. 9-13. / losses due to propeller inefficiency
The power available from I
the propeller.

TAS

Owing to inefficiencies of the propeller, not all of the engine power is converted to
propulsive power. (Discussed in more detail in Chapter 12.) Most propellers are
least efficient at low speeds with their efficiency increasing to a maximum of about
80% over the higher speed range of the aircraft. The cu1ve for (propulsive) power
available from the propeller is therefore typically steeper and more curved than that
for power delivered at the propeller shaft, as shown in Fig. 9-13.

The power available (PA) curves which we use in the following diagrams is the
lower of the two in Fig. 9-13, i.e. the (propulsive) power available from the
propeller.

Maximum and Minimum Speeds in Level Flight


When the power required and power available cwves for level flight in an aircraft
are compared, a number of factors relating to performance emerge. The power
required curve is related to the aerodynamics of the aircraft, and can be thought of
as the aerodynamic demand for power to meet a certain performance. The power
available curve is related to the engine/propeller combination, and can be thought
of as the ability of that combination to meet that demand for power under given
circumstances. Each set of curves shown in the following diagrams is valid for one
aircraft weight, one altitude, and one configuration.

The maximum speed for level flight at any given power setting occurs where the
cwves intersect at the higher TAS (Fig. 9-14). At that point, the power required =
power available (and the thrust developed will equal the drag). As power is
reduced from maximum, the PA curve moves down on the graph; the intersection
of the curves moves to the left and; as you would expect, the maximum speed at
the lower power setting is reduced.

The minimum speed for level flight occurs either where the curves intersect at the
lower TAS or at the stalling speed. For most aircraft, the stalling angle of attack (or
some condition of instability or loss of control effectiveness) will be reached before

9-10 Straight and Level Flight The Commercial Pilot Series


maximum power available and thrust are required to counterbalance the drag.
Whichever of these speeds is reached first as the aircraft is slowed down,
becomes the minimum speed possible in level flight.

~...-- full power


stall
power
:-90%\\
available
PA

-
--
-
I
__.1 -sa%- p A

r-
Fig. 9-14.
and I
-:- 70/
--
;t('
Maximum and minimum I
speeds in level flight. power
I
required
PR
I I
I
I
I
I
+
minimum TAS maximum
speed I speed

The Effect of Weight


In level flight at any given airspeed, if the weight of
the aircraft is increased, the angle of attack will also
have to be increased if the lift is to remain equal to stalling speed
the weight, This increase in angle of attack results in increased
an increase in drag and therefore of power required
,,I.
to maintain level flight (PR ~ drag x TAS). Increased ,, II
II
aircraft weight thus has the effect of moving the PR II
p I'
;I
1
1 max.
curve upward and to the right, which means that the 1t- speed
power required to maintain level flight at all speeds is I }I : 1 reduced
increased. As shown in Fig. 9-15, there will also be a I II
slight reduction in the maximum level flight speed I R to maintain II
and, as we have already seen, an increase in the II all speeds
increased
:I
stalling speed. Conversely, at lower aircraft weights, -1!,
less power is needed to maintain level flight at any TAS
given speed, there will be a slight increase in the
maximum speed attainable, and the stalling speed Fig. 9-15. Increased weight increases
the power required to maintain speed
decreases. in level flight.

The Effect of Altitude


An increase in altitude affects both the power required and the power available.

Power Required

The drag at any aircraft weight and lAS remains constant regardless of altitude.
However, if an aircraft is flown at constant lAS but increasing altitude, theTAS at
that lAS steadily increases. Since the power required to fly at any given lAS is a
function of TAS x drag, it follows that the power required to maintain an
indicated airspeed increases as altitude is gained (in spite of the fact that the
drag remains constant). Looked at another way, as an aircraft is climbed at

Principles of Flight Straight and Level Flight 9-11


constant lAS, an advantage is gained because its real speed (TAS) is steadily
increasing. This advantage comes at the expense of an increase in power
required.

When successive PR cUives are plotted against TAS to reflect increasing altitude,
they move upward and to the right as shown in Fig. 9-16 (Upward because of
the increase in power required PR; and to the right because successive points on
the cUives representing the same lAS (and drag) occur at a higher true
airspeed.)

Power Available

The decrease in air density at altitude inevitably causes a decrease in the power
available from both jet and piston engines. This reduction in power available
from a piston engine may be largely delayed by the use of supercharging, but
supercharged engines must also ultimately suffer a loss of power as altitude is
gained. As a result, an increase in altitude is reflected in a downward movement
of the PA curve on the graph.

The effect of altitude in increasing the power required and reducing the power
available is shown in Fig. 9- I 6.

Note that as altitude is increased, the


maximum TAS in level flight is reduced.
min.speed
However, with a supercharged engine, increased
an increase in maximum TAS will be sea level
-I-
possible up to such altitude as the
supercharger is able to maintain
I 1:
It/
/ 1,_~..-· .r,lI

~:
maximum boost. Thereafter, the effect II
p !I max.
of altitude will be the same as for a I ~speed
normally aspirated engine. 1 1 reduced
I I
The minimum level flight speed is I I
increased with altitude. The TAS at I : I
which the aircraft stalls will be .j. .j.
increased but the rate at which the low-
TAS
speed intersection of the PR and PA
cutves moves to the right on the graph is
Fig. 9-16. The effect of altitude on power
faster. As represented in Fig. 9-16, this required and power available.
means that in most aircraft an altitude
will be reached where the minimum
level flight speed is above the stalling
speed.

Hence, in general terms, the effect of altitude is to reduce the maximum speed
and increase the minimum speed in level flight. Ultimately, an altitude will be
reached where the two coincide and level flight will only be possible at one
speed. When that point is reached, the aircraft will be at its absolute ceiling, as
is discussed in more detail in the next chapter under climbing.

9-12 Straight and Level Flight The Commercial Pilot Series


Review 9

1. The four main forces acting on an aircraft in flight are ............ , .............. ,
............. and ................... .

2. For straight and level flight at constant airspeed, ............... must


counterbalance the weight, and ...................... must counteract the drag.

3. The four main forces in flight rarely act through the same point. In most
circumstances there will be two couples acting to pitch the nose up or
down. These couples are the ................... couple and the ................... couple.

4. If the CG is ahead of the CP, the lift/weight couple will tend to pitch the nose
(up/down).

5. If the line of drag is above the line of thrust, this will tend to pitch the nose
(up/down).

6. It is the function of the ................................ to balance out the lift/weight and


thrust/drag couples in flight.

7. If power is reduced (thrust decreased) the nose of the aircraft will tend to
pitch (down/up).

8. For a low-wing aircraft, when the flaps are lowered you should expect the
nose to pitch (up/down).

9. For a high-wing aircraft, you should expect the nose to pitch (up/down)
when the flaps are lowered.

I 0. After an increase in speed in straight and level flight, the nose attitude will be
(higher/lower) than it was previously.

II. The curve of 'thrust required' for level flight is the same as the aircraft's ....... .
CUlVe.

12. The minimum thrust IAS is the speed for minimum ................. and
maximum ......................... ratio.

13. In straight and level flight at any given lAS, the power required = .............. .
X ...... .

I 4. For a given piston-engined aircraft in straight and level flight, the minimum
power IAS is (higher than/equal to/lower than) the minimum thrust lAS.

15. Given that drag can be equated to power/TAS, the minimum drag speed can
be found on the power required cUlve at the point where the line drawn
from the origin:
(a) is tangential to the cmve;
(b) intersects the bottom of the cmve.

16. The propulsive power available from the propeller at any given TAS is:
(a) the brake (horse)power measured at the propeller shaft;
(b) brake power minus the loss due to inefficiency of the propeller.

Principles of Flight Straight and Level Flight 9-13


17. The power available from the propeller at a given power setting, and the
power required for level flight, can be graphed against TAS. State what the
upper and lower intersections of these curves represent.

18. If the weight of the aircraft is decreased in level flight:


(a) the power required to maintain any given lAS is (increased/decreased);
(b) the maximum speed attainable will be (increased/decreased);
(c) the stalling speed will be (increased/decreased).

19. The power required to maintain any given lAS in level flight (increases/stays
the same/decreases) at higher altitude.

20. The power versus TAS (power required) curve for level flight moves (up/up
and to the right) on the graph as altitude increases.

21. The power available versus TAS curve moves (down/down and to the right)
as altitude is increased (setting aside the effect of supercharging).

22. When the point is reached where the bottom of the power required curve
just coincides with the power available curve, the aircraft is said to have
reached its ............................. ceiling.

9-14 Straight and Level Flight The Commercial Pilot Series


Climbing and Descending
Climbing
As an aircraft climbs it is gaining potential energy by virtue of the gain in altitude.
There are two ways in which this can be achieved:
• by making a zoom, where there is a short-term gain in altitude and a loss of
airspeed; or
• by climbing at a steady speed.

The Zoom

Provided it has sufficient speed, the aircraft will zoom climb if the nose attitude
is raised above the normal straight and level flight attitude for the speed
concerned. The climb is achieved by converting the velocity of the aircraft
(kinetic energy) into a gain in altitude (potential energy). Hence, speed must
reduce as altitude is gained in the zoom-in effect, speed is traded for altitude.
If the power setting is not changed, the altitude gain will depend on the initial
speed-a high speed conferring greater potential to gain altitude and vice versa.
If the nose attitude is not too high, the aircraft will probably gain some altitude
and then settle in level flight at a slower speed. With a very high nose attitude,
speed will continue to reduce until the aircraft stalls.

An extreme example of the zoom climb can be seen in an aerobatic glider


which, without the benefit of thrust, is able to gain sufficient speed (and kinetic
energy) to be flown through a looping manoeuvre without reducing speed
below the stalling speed.

The Steady Climb

Aircraft are usually climbed at a steady speed and rate of climb. In the normal
climb, power in excess of that required to counterbalance the drag is used to
gain altitude (potential energy). The greater the excess of power available, the
greater the ability to climb at a steady speed.

zoom climb:
decreasing airspeed and rate of climb

steady climb:
constant airspeed and rate of climb

Fig. 10-1. The zoom climb and the steady climb.

Principles of Flight Climbing and Descending 10-1


The Forces Acting in the Climb
Before we examine the forces acting in a climb, consider the behaviour of a
vehicle at rest, first on a level surface, then on a slope (Fig. I 0-2).

On the level surface, the weight of the vehicle acts at right angles to the surface.
There is no force acting across the surface to move it and it will stay where it is,
unless it is driven or pushed away. However, when the van in our example is
parked on a slope, there are now two components to its weight:
• one at right angles to the slope (WI)
which acts to hold the vehicle against
the slope; and
• One which acts parallel to the slope,
which we can call the rearward
component of weight (RCW). If the van
had its parking brake released and was
out of gear, the rea1ward component of
its weight would cause it to roll w
backwards down the slope-the size of
the RCW and the rate of backward Fig. 10-2.
acceleration depending on the
steepness of the slope.

Although an aircraft in a steady climb does not have a solid slope to move on,
the principle of forces is exactly the same as a van on a hill. As shown in Fig. I 0-
3, an aircraft in a steady climb follows a flightpath which is inclined upward from
the horizontal by the climb angle (which we show as 8). Weight, as always, acts
vertically downwards, but in a climb it has the same two components as the van
in our previous example-one component which acts at right angles to the flight
path (WI), and the other which acts back down the flight path and parallel to
it-the rea1ward component of weight (RCW). The component of weight (WI)
in effect holds the aircraft against the flight path (when it is balanced by the lift)
The rea1ward component (RCW) acts in the same way as it did with the van-if
there is no force to oppose it, the aircraft will accelerate backwards down the
flight path. Hence, to climb at a steady speed, the rearward component of
weight must be opposed by thrust, acting parallel to and along the flight path.
Climbing does not therefore depend on generating more lift-which acts only to
balance the component of weight at right angles to the path of the climb.

lilt
Fig. 10-3. The forces acting in a 1.- -
climb (ignoring drag).

rearward component of weight


parallel with fiightpath
(W sin 0)

----- --
If.,,
:£.1
------((~limb angle 0~ / component of weight
~----------- -\A'"'" perpendicular to flightpath
I (W cos 8)
\
_jW1
weight W

10-2 Climbing and Descending The Commercial Pilot Series


Note that the relative strength of the two components of weight depends on the
climb angle e. From simple trigonometly, we see that WI = W Cos e, and RCW
= W Sin 0. In practical terms, this means that as the climb angle becomes
steeper, the rearward component of weight (RCW) increases and the
perpendicular component (WI) diminishes. In a 90' climb, WI disappears and
the whole of the aircraft weight becomes the 'rearward component'.

In a steady climb at constant speed, the forces must be in equilibrium (if they
are not, the aircraft speed and/or the angle of climb will be changing). Ignoring
for the time being, the effect of aerodynamic drag, it can be seen from Fig. I 0-3
that the conditions for a steady climb will be met when:
• sufficient thrust is provided to balance the rearward component of weight,
i.e. Tl = RCW; and
• the lift, acting at right angles to the flightpath (as it does) balances the
remaining component of weight at right angles to the flightpath, i.e. L = WI.

That the forces are in equilibrium, can be confirmed by drawing in the


resultants. It can be seen from this that the weight (W) is equal and opposite to
(and supported by) the resultant of the lift (L) and the component of thrust
which we have called TI.

To complete the diagram of the forces acting in the climb, we must add in the
aerodynamic drag which is always present and acts parallel to the flightpath.
We have done this in Fig.l 0-4 and it can be seen that to maintain the balance of
forces, thrust must be increased by an amount equivalent to the drag, such that
total thrust (TT) is equal to the rearward component of weight plus the drag, i.e.
TT= RCW+ D.

The aircraft will maintain a steady climb, with the forces in equilibrium, when
the resultant (R) of the 'forward and upward' forces-total thrust and lift,
balances the resultant (Rl) of the 'rearward and downward' forces-the two
components of weight plus drag.

Fig. 10-4. The forces acting in a


steady climb.

NOTES:

I. In Figs. I 0-3 and I 0-4, we have assumed that thrust acts parallel to the
flightpath. In reality, it will normally be inclined upwards at a small angle.

Principles of Flight Climbing and Descending 10-3


2. In a steady climb, as the angle of climb (e) increases, the requirement for
aerodynamic lift decreases. From Figs. I 0-3 and I 0-4, it can be seen that lift must
remain equal and opposite to the component of weight which is perpendicular to
the flight path. This component (WI = W cos e) reduces as the climb angle
becomes steeper, becoming zero in a 90" climb where, accordingly, no lift is
required. However, for training aircraft, where normal climb angles are quite
small (perhaps about 5") the lift force is only marginally less the weight. [Cos 5" =
0.99 which means that in a 5" climb, lift is 99% of the value for weight].

3. To climb at a given airspeed, sufficient thrust must be available to overcome


the drag at that airspeed and the rearward component of weight. Therefore, the
ability to climb (and how well the aircraft will climb) depends on having a surplus
of thrust (and power) available after overcoming the drag at the chosen climb
speed.

Climb Performance
The climb performance of an aircraft encompasses the:
• speed in the climb;
• rate of climb-i.e. the rate at which it gains altitude in feet per minute as
indicated on the vertical speed indicator (VSI); and
• angle of climb-the steepness of the flightpath over the ground, for which
there is no direct instrument indication in the cockpit.

Climbing Speed

For any given power setting, the speed in the climb is determined by the nose
attitude selected by the pilot. The higher the nose attitude, the lower the climb
speed and vice-versa. The recommended climbing power setting(s) and speed
(s) for a specific aircraft are given in the Pilots Operating Handbook.

Rate of Climb

At any given speed, the rate of climb is determined by the excess of power
available over the power required-i.e. the amount of power applied over the
power required to overcome the drag at that speed (in straight and level flight).
Shown in Fig. I 0-5 are typical power available/power required curves for a piston-
engined aircraft. The excess of power available (PA) over power required (PR) is
represented by the shaded area between the two curves.

full

Fig. 10-5.
At any given TAS, the rate of power
climb is determined by the (P)
excess of power available
over power required.

0 min. speed
TAS max. speed
level flight level flight

10-4 Climbing and Descending The Commercial Pilot Series


The diagram at Fig. I 0-5 is based on the PR cwve for straight and level flight. It is
quite normal (and convenient) to do this for other than steep climb angles since
the angle of attack (and therefore the drag) at any given speed in the climb is
virtually the same as the angle of attack and the drag at the same speed in level
flight. Thus, any errors present in using this curve will be negligible. Note that:
• At any given TAS, the rate-of-climb performance is determined by the length
of the vertical lines between the cwves-i.e. the difference between PA over
PR. For the typical aircraft represented, there is a range of speeds where
there is little difference in the amount of excess power available and
therefore little difference in the rate of climb.

• The PA curve shown in Fig. I0-5 is for full power (maximum MP/rpm). At
this power setting, it is possible to climb the aircraft at all speeds from just
above the stalling speed to just below the maximum speed for straight and
level flight (where admittedly the climb performance will be very poor). If a
lower power setting is used, the PA curve moves downward on the graph,
the excess of PA ovel' PR reduces, and the climb performance over the
range of speeds possible will be reduced.

Maximum Rate of Climb

The maximum rate of climb occurs at the speed where there is the greatest
excess of power available_ This is more clearly seen if the excess power (EP)
available is extracted from the PNPR curves and is plotted separately against
TAS as shown in Fig. I 0-6. (This curve contains the same information on excess
power as Fig. 10-5, but presented in a different way).

excess
power
available I
EP
Fig. 10-6. (max. rpm)
The excess power cutve.

stall
1\
max. !peed
TAS
speed level flight
speed for
maximum
rate of climb

The climb speed for maximum rate of climb is the speed which coincides with
the peak of the excess power curve for maximum MP/rpm. At any other point
on the curve, the rate of climb will be lower, although on either side of the
maximum rate of climb speed, it can be seen that there is a range of airspeeds
where there is little difference in the amount of excess power available and,
therefore, of rate of climb.

Maximum Angle of Climb

It can be seen from Fig. I 0-4 that in a steady climb the total thrust (TT) must
balance the drag (D) plus the rearward component of weight (RCW). From Fig.
I 0-3, we saw that RCW = W sin e. That is, in a steady climb:
total thrust (TT) = D + W sin e.

Principles of Flight Climbing and Descending 10-5


Transposing the factors in this equation:
. total thrust - drag
sm 9 = weight (where e = the angle of climb)
Now, total thrust less the drag, represents the excess of thrust available at any
given power setting. Therefore, the angle of climb in still air is determined by
the excess of thrust available over the drag at the climbing speed. The
maximum angle of climb occurs at the speed where this excess of thrust over
drag is at a maximum.

Note that here, for angle of climb, we have introduced the term excess thrust
whereas before, we used the term excess power in relation to the rate of climb.
Thrust and power are not the same thing. In simple terms:

• Thrust is a force. An aircraft propeller generates thrust by accelerating a


mass of air toward the rear. The amount of thmst generated depends on the
increase in velocity given to the mass of air entering the propeller disc-the
greatest increase in velocity (and therefore the highest thrust) occurs when
the aircraft is stationary. As airspeed is increased, the propeller is less able
to cause further acceleration to the air entering the disc, and thus the thrust
drops off. Eventually, a high speed is reached where the propeller is unable
to accelerate the air passing through it, and no thrust will be produced.

• Power is the rate of doing work. It is measured by the combination of the


force applied and the velocity which it produces. The power required curve
combines the thrust required for level flight with theTAS gained. The power
available curve gives an indication of how capable the engine/propeller
combination is in effectively applying thrust at different true airspeeds.
Whereas the 'raw' thrust available from the propeller decreases with speed,
when it is combined with TAS to obtain power available, a slight increase
with speed is shown. Thmst is the force produced by the propeller. Power
is the measure of how effectively that thrust can be put to work e.g. to fly the
aircraft level or to climb at different true airspeeds.

Fig. 10-7a shows typical thrust available and drag ('thrust required') curves for a
piston-engined aircraft. Note the difference between these cmves and the PA
and PR curves. The maximum thrust is available when the aircraft has zero
forward speed and decreases (rapidly at first) as airspeed is gained. The power
available on the other hand increases as airspeed is gained. The difference
between the drag and the power required cmves has already been illustrated at
Fig. 9-10.

max. thrust
available (TA)
I excess power
EP
/ (EP)
at max. rpm

I
drag
(or 'thrust
required')

stall TAS
Vx Vx Vy
Fig. 10-7. (a.) (b.)

10-6 Climbing and Descending The Commercial Pilot Series


Because of the shape of the thrust available and drag/'thrust required' curves for a
piston engined aircraft (Fig I0-7a), it is not easy to distinguish where the maximum
excess of thrust available over drag (and therefore the best angle-of climb speed-
Vx) occurs. This is typically at a speed which is just above the stalling speed.
Indeed, on some aircraft the maximum angle of climb speed is the lowest safe
speed above the stalling speed at which the aircraft can be climbed. If the aircraft
is climbed at any higher speed the excess of thrust available over drag decreases as
does the angle of climb.

Vx -the maximum angle-of-climb speed-can also be found (and is more clearly


seen) on the excess power curve at the point where a line drawn from the origin of
the cmve is tangential to the curve. This is shown in Fig. I 0-7b with the maximum
rate-of-climb speed Vv shown for comparison. At Vx, the ratio of excess power (i.e.
rate of climb) to TAS (forward speed) is greatest and, therefore, so is the climb
angle. At VY, while the rate of climb is higher, the ratio of rate-of-climb to forward
speed is lower, as is the climb angle. This difference between the two speeds is
shown again in the next diagram (Fig. I 0-8).

To summarize:
• The maximum angle-of-climb speed is lower than the maximum rate-of-climb
speed. For a piston-engined aircraft, the maximum excess of thrust over drag
(which is required for maximum angle of climb) typically occurs at a speed
which is close to the stalling speed. The maximum rate of climb speed, which
relies solely on the amount of excess power available is usually much higher.
• The maximum angle-of-climb speed occurs at the point on the excess power
curve where a line drawn from the origin is tangential to the cmve. The
maximum-rate of climb speed lies below the peak of the excess power curve.

Normal Climbing Speed

Training aircraft are usually climbed at maximum power. The normal climb speed
is usually higher than the speed for maximum rate of climb and is a compromise
between speed and rate of climb. As we have just seen, there is a range of speeds
over which there is little variation in excess power and therefore rate of climb.
Climbing at the higher end of this range means little is lost in terms of time to climb
to the desired altitude, but the higher speed permits better engine cooling, better
fmward visibility, and an overall reduction in time taken en-route.

t
ALTITUDE GAINED MAXANGLE Vx
MAX RATE
CLIMB Vv
IN A GIVEN TIME CLIMB

Vy
Fig. 10-8. Vx
normal

Note: Refer to your Pilot's Operating Handbook for the actual values of the various
climb speeds for your particular aircraft. Typically, the best angle-of-climb speed Vx
is in the order of I 0 or 15 knots less than the best rate-of-climb speed VY.

Principles of Flight Climbing and Descending 10-7


Factors Affecting Climb Performance
Anything which acts to reduce the amount of power available over the power
required (i.e. the excess power available) will affect the performance of the
aircraft in the climb. The factors include; the amount of power applied, the
airspeed being maintained, the weight of the aircraft, flap and/or undercarriage
extension, altitude, temperature, manoeuvring and, insofar as climb angle over
the ground is concerned, the wind.

The Amount of Power Applied


Reduced power decreases climb performance since the amount of excess
power over power required is reduced. Both the rate and angle of climb will be
reduced. If the aircraft is not climbing as well as expected, first check that full
power (or the recommended power) is set.

Airspeed.

Flying faster or slower than the recommended airspeed can degrade the climb
performance which is being aimed for.

Flap and/or Undercarriage Extension

Extended flaps (and particularly full flap) will decrease climb performance
since, for almost all aircraft, the L/D ratio is reduced and the power required to
maintain any given speed is increased. The best rate-of-climb performance is
invaliably achieved with the flaps up. The angle-of-climb performance with full
flap will also be much poorer than with the flaps up.

The general tendency will also be for the angle of climb to be lower with smaller
amounts of flap deflection because of the poorer UD ratio. However, for some
aircraft, the Pilots Operating Handbook may recommend the use of a 'take-off
flap setting for a 'short field' take-off-i.e.to obtain the shortest take off distance
and best obstacle clearance in the climb-out at a safe speed above the stalling
speed. Although the UD ratio is poorer in these circumstances, the use of the
take-off flap setting results in a relatively small increase of drag but permits a
slower safe climbing speed where this increase in drag is virtually offset by the
higher thrust available. (Refer again to Fig. 10-?a).

As a general rule, an aircraft should not be flown in an extended climb at the


maximum angle-of-climb speed. In this type of climb, full power is being used at
a very low airspeed. As a consequence, cooling is poor, and the conditions are
conducive to overheating, detonation and engine damage. Use the maximum
angle climb only for as long as is necessary and be careful with your handling of
the aircraft as the speed employed is usually very close to the stalling speed.

In aircraft with retractable undercarriage, leaving the landing gear extended


increases the drag and reduces the climb performance.

Weight

An increase in weight will degrade climb performance. The power required to


maintain any given speed in level flight increases with increased weight (Fig. 9-
15 refers). There is therefore a smaller excess available for climbing. Heavy
weights can markedly degrade both the rate and angle of climb. The best climb
performance is achieved at the lighter aircraft weights.

10-8 Climbing and Descending The Commercial Pilot Series


Altitude

The decrease in air density as altitude is gained causes a decrease in both


engine and airframe performance. As shown in Figs. I 0-9 and I 0-10, the power
required and power available curves move closer together and the climb
performance gradually deteriorates as the excess power available reduces. If
the climb is continued, the power required/power available curves coincide and
all of the power available will be needed to maintain the aircraft in level flight at
the speed where the curves coincide. At this point the aircraft will have reached
what is called its absolute ceiling. The service ceiling-which has a more
practical application-is reached at a lower altitude, when the rate of climb has
been reduced to 100 feet per minute.

PA sea level note the changes to max.


rate of climb ( +------)
and to max. angle of climb ( 'V )
as altitude is increased

EP

TAS TAS

Fig. 10-9. The effect of altitude on the excess power available for climbing.

Fig. 10-10.
Climb performance deteriorates
gradually as altitude is gained. - - - - - - - - - absolute ceiling

100 fpm- service ceiling


200 fpm

500 fpm

sea level

Temperature

At any altitude where the air temperature is higher than the standard figure,
climb performance will be reduced because of the reduction in air density.

Manoeuvring

Any manoeuvring in the climb-e.g. turning-will absorb some or all of the


excess power available and climb performance will be reduced.

Principles of Flight Climbing and Descending 10-9


Wind

Headwind and tailwind components affect the climb angle over the ground,
but not the rate of climb as shown in Fig I 0-11 below.

ALTITUDE GAINED
t .ililllwind 0
same climb speed and power
windilill·~

IN A GIVEN TIME
same
altitude
gain

Fig. 10-11. The effect of


wind on the climb angle. no wind tailwind

DISTANCE TRAVELLED over the ground IN A GIVEN TIME

Caution: Raising the Flap after a Baulked Approach


In Chapter 9, because of the possibility of the aircraft sinking as lift is lost, a
caution was given about raising the flaps when flying at low levels and low
speeds The most likely requirement for this is during a go-round (or baulked
approach) with full flap lowered.

With full flap lowered, the climb performance of most aircraft is significantly
reduced. When there are no obstacles ahead, the climbout with full flap is not
normally a problem and, to prevent the possibility of sinking, the usual practice is
to establish the aircraft in a positive climb at a safe height before starting to raise
the flaps. If however, circumstances dictate that the flaps should be raised at a
lower height (e.g. where this is necessary to gain a better climb angle to avoid an
obstacle), they should be raised slowly to enable the aircraft to gradually
accelerate to the higher airspeed. In this way, the loss of lift as the flaps are
raised is balanced by the gain in lift through the increasing airspeed. If done
carefully, the aircraft will not sink and a safe margin will be maintained above
the stalling speed.

Descending
There are two ways in which an aircraft is normally descended-in a glide or in
a power-on descent.

Gliding involves closing the throttle fully so that no thrust is produced. In fact,
with the throttle fully closed, in most circumstances the propeller will be driven
by the airflow-it will 'windmill' and create a small amount of additional drag.
Flying speed in the glide is maintained by lowering the nose attitude and by
allowing the potential energy of the aircraft (altitude) to be converted to kinetic
energy (speed) in a controlled fashion. In effect, altitude is traded for speed.
The speed maintained in the glide is determined by the nose attitude selected
and, within the range of gliding speeds available, the rate of descent is
determined by the speed being maintained.

In a power-on descent, speed is also controlled by the nose attitude, but the rate
of descent is determined and controlled by the amount of power applied.

10-10 Climbing and Descending The Commercial Pilot Series


Gliding
The forces acting in the glide are shown in Fig. 10-12. In this diagram, it is
assumed that any thrust from the propeller is negligible with the throttle closed.
TR lift
fj,>
,')
,'I
'
''
component of weight
parallel to flightpath

Fig. 10-12. '


The forces acting in a glide. '' '
' '
''' ''
'
t_' '
'

weight

In the glide, only three of the four main forces are present-weight, lift and drag.
In a steady glide at constant speed these three forces are in equilibrium. As
shown in the diagram, if the forces are resolved with relation to the flightpath,
the perpendicular component of weight is balanced by the lift, and the
component of weight parallel to the flightpath (which provides the speed of the
aircraft) is equal and opposite to the drag. To maintain a given speed in the
glide, the nose attitude of the aircraft must be adjusted so that this component of
the weight remains equal to the drag. When the forces are resolved vertically,
the resultant of lift and drag-the total reaction force (TR)-is equal and
opposite to the weight.

Gliding Performance
Glide Angle

The glide angle at any given speed in the glide is directly determined by the lift/
drag (LID) ratio. If the LID ratio is high the glide angle will be shallow. If the LID
ratio is poor, the glide angle will be steep.
TR

w
Fig. 10-13. The lift/drag ratio determines the angle of glide.

Principles of Flight Climbing and Descending 10-11


As can be seen from Fig. 10-13, the glide angle (distance travelled horizontally in
the airmass per unit of altitude lost) is in direct proportion to the l!D ratio. The
greatest range (and the shallowest angle) in the glide will be achieved when the
aircraft is flown at the angle of attack for the best UD ratio. This angle of attack
is usually about 4°. The pilot has no direct indication of angle of attack in the
cockpit, hence flying at the recommended glide or descent speed will ensure
that the aircraft is at or close to this most efficient angle of attack.

If the glide speed is higher or lower than recommended, the UD ratio will not be
as good and the glide angle will be steeper.

best LID
ratio I higher angles
of attack:
I steeper glide
I
Liftldrag I
ratio I
I
lower angles
I
Fig.10-14. Gliding at the recommended of attack: I
speed gives the best UD ratio and the steeper glide I
greatest distance in the glide. I
higher 1 lower
<:=l=i>
speeds l speeds

oo •
4° Angle-of-attack 16°
t
best glide angle of attack

When undershooting on a glide approach, there is a great temptation to raise the


nose attitude and 'stretch the glide'. The higher nose attitude may give the
appearance of a decreased glide angle, but in fact the aircraft will be descending
on a steeper path at a higher angle of attack. To obtain the best gliding
performance, keep to the recommended speed. Flying at a higher or lower
speed will result in a steeper glide angle and reduced range.

Fig. 10-15. Flying at the correct


gliding speed ensures the best
range is achieved in the glide.

9
I

Use of Flaps

In almost all aircraft, the lowering of flaps reduces the UD ratio and the glide
angle is steeper as a result. When the flaps are lowered, the nose attitude must
also be lowered to maintain the gliding speed. The stalling speed will also be
lower and it will be possible to glide with flap down at a slower speed and still
maintain a safe margin above the stalling speed. However, at this lower speed,

10-12 Climbing and Descending The Commercial Pilot Series


the UD ratio will be poorer and consequently the glide angle will be steeper
than with the flaps up and gliding at the recommended speed. When flying a
glide approach to land, do not lower the flaps until you are sure that the aircraft
has sufficient height to reach the intended landing point at the steeper glide
angle.

The Effect of Weight

Aircraft weight does not affect the glide angle. A decrease in weight means that
the angle of attack for best UD ratio will occur at a slightly lower airspeed.
Although the range in the glide is not altered, the rate of descent will be slightly
lower and the time taken to reach the ground will be longer.

J,.- L
same angle of attack:
/I different weights and
I
/ recommended speed different airspeeds
I
I

Fig. 10-16. The best glide angle is the same at all weights but
airspeed must be lower at lower weights.

The recommended glide speed stated in the Pilot's Operating Handbook is


based on maximum gross weight. The variation in weight for most training
aircraft is not usually great enough to significantly affect the glide if the
recommended glide speed is used at all times-even though, theoretically, a
slightly lower glide speed could be used when lightly loaded.

The recommended descent speed in the Pilot's Operating Handbook will be


suitable for all normal weights in a light training aircraft.

The Effect of Wind

A headwind component reduces the glide distance over the ground. It does not
affect the glide distance through the airmass, or the rate of descent. Conversely
a tailwind component increases the range of the glide over the ground, but has
no effect on the glide angle through the air, or on the rate of descent.

The effect of wind is illustrated in Fig. 10-1 7. In all three cases illustrated (nil
wind, headwind and tailwind), the glide angle relative to the airmass is the
same. If there is no wind, the flightpath of the aircraft will be the same as its
glide angle. When the airmass is moving relative to the ground (when there is a
wind), the fiightpath in the glide will be different-steeper in a headwind,
shallower in a tailwind. Note however, the nose attitude and airspeed in all
three cases will be identical.

Principles of Flight Climbing and Descending 10-13


tailwind component

Fig. 10-17. More ground is covered gliding with a tailwind, less with a headwind.

The Power-on Descent


In a glide, performance is necessarily limited because no power is applied-the
performance of the aircraft (airspeed and rate of descent) is determined solely
by attitude. To give greater control over the rate of descent, thrust can be added
by applying power.

If, for example, the aircraft is in a glide at the recommended speed, the
flightpath may be made shallower and the rate of descent reduced by applying
power and raising the nose attitude at the same time so that the speed is
maintained. As more and more power is applied, and the nose further raised,
the aircraft will eventually return to level flight Yet more power and an even
higher nose attitude and it will climb. In this way, the rate and angle of descent
may be controlled with power, and the airspeed with the attitude selected.

Fig. 10-18. The forces acting


in a power-on descent.
I
component of weight r

\j
assisting forward motion
and balancing drag
I
W

There are a number of combinations of airspeed, and rate and angle of descent
available in a power-on descent and, once again, it is a matter of selecting the
right amount of power and placing the aircraft in the right nose attitude to obtain
the performance you want. Any small adjustments to the power will require a
corresponding adjustment to be made to the nose attitude for the desired speed
to be maintained.

If you wish to descend at the same airspeed but at a higher rate with power on,
there are two things you can do:
• reduce the power and/or

• increase the drag (lower the flaps, extend the undercarriage-if retracted).
When the flaps are lowered, you will usually need a lower nose attitude.

10-14 Climbing and Descending The Commercial Pilot Series


Review 10
Climbing

1. In a zoom climb, kinetic energy is converted to potential energy by


trading ............................... for .............................. .

2. In a steady climb at constant speed, power in excess of that required to


balance the (weight/drag) is used to gain altitude.

3. In a steady climb, if the weight of the aircraft is resolved with respect to the
flightpath, there will be two components-one parallel to the flightpath, [the
rea1ward component of weight (RCW)], and one perpendicular to the
flightpath. The RCW is opposed by (thrust/drag), while the perpendicular
component must be counterbalanced by (lift/thrust).

4. In the full diagram of forces acting in a steady climb:


(a) total thrust must balance the ................................................... of weight
plus the ............................. ; and
(b) the lift must balance the ............................................................ of weight.

5. The perpendicular component of weight which is balanced by the lift is


measured by W cos 8 (where 8 is the angle of climb). For small angles of
climb, it is therefore true to say that the lift is fractionally (greater/less) than
the weight.

6. The speed in a steady climb is determined by the ........................ .


.......................... selected by the pilot.

7. In a steady climb, the rate of climb is determined by the amount


of ............................ available (i.e. applied) over the amount
of ............................... required to overcome the drag at that speed.

8. At any given TAS, the rate of climb which is achievable is determined by the
excess of power ............................ over power ................................ .

9. The maximum (or best) rate of climb is obtained at the TAS where there is
the greatest ............................. of power available over power required.

10. If the aircraft is climbed at less than full power, the excess of power
available will (increase/decrease) and the rate of climb achieved will be
(higher/lower).

11. The maximum (or best) angle of climb will be achieved at theTAS where
there is the greatest excess of ................................ available over drag.

12. The speed for best angle of climb (Vx) can be found on the excess power
versus TAS curve, at the point where the line drawn from the origin
is ............................ to the curve. The speed for best rate of climb (VY) can be
found on the same curve at the ................ of the Clllve.

13. For a piston-engined aircraft, the best angle of climb speed (Vx) is typically
(lowe1)higher) than the best rate of climb speed (VY).

14. The normal climb speed will usually be a faster speed than VY, and at a
slightly lower rate of climb. State three advantages of climbing at this higher
speed.

Principles of Flight Climbing and Descending 10-15


15. Rate-of-climb is measured by the amount of height gained in a
given ...................... . Climb angle is a function of the amount of height
gained in a given horizontal ................................... .
16. If power is reduced, climb performance (will be poorer/will not be
affected).

I 7. The rate of climb with flap extended will be (better than/the same as/poorer
than) the rate of climb without flap.

18. A better rate and angle of climb will be achieved when the aircraft is at
(higher/lower) weight.

19. The effect of altitude on the power required and power available curves is;
(a) the power required curve moves .......... and to the ..................... :
(b) the power available curve moves ........................ .

20. Therefore (from Q 19), the excess of power available over power required
(decreases/stays the same/increases) and climb performance (decreases/
stays the same/increases) as altitude is gained.

21 If the aircraft is manoeuvred (e.g. turned) while in a steady climb at


constant power, the rate of climb will be (reduced/unaffected).

22. If an aircraft is climbed in a tailwind, the climb angle over the ground will be
(less than/the same as) in nil wind conditions. The rate of climb will be
(unaffected/higher/lower).

Descending

23. In a steady speed glide, the component of ............................... which is


parallel to the tlightpath is equal to the drag.

24. For the shallowest glide angle and the best still-air gliding range, the aircraft
must be flown at the airspeed which achieves the angle of attack for the
best ............/ .............................. .

25. If the flaps are lowered in the glide and speed is maintained, the nose
attitude will be (lower/higher); the glide angle (steeper/shallower), and the
rate of descent (higher/lower).

26. Aircraft weight (affects/does not affect) the glide angle. At a lower weight,
the speed for the best glide angle will be (higher/lower).

27. A headwind or tailwind component (does/does not) affect the glide angle
through the airmass.

28. Gliding with a tailwind component will make the tlightpath over the ground
(shallower/steeper). The range in the glide will be (increased/decreased)
but the time taken to reach the ground will be (longer/shorter/the same).

29. In a power-on descent, the speed is controlled by ......................................... .


and the rate of descent is controlled by the amount of .......................... ..

10-16 Climbing and Descending The Commercial Pilot Series


Turning
Introduction
11
In the last two chapters, we considered the forces acting on the aircraft in
straight flight-either level, climbing or descending. We now consider the forces
acting when an aircraft follows a curved path, i.e., one which requires radial
acceleration toward the centre of the curve. We look at four specific cases-
level, climbing and descending turns, and manoeuvring in the vertical plane.
Any manoeuvring in a non-aerobatic aircraft will normally be made up of these
elements. A study of more complex aerobatic manoeuvres is outside the scope
of this manual, although the general principles we discuss here apply equally to
most manoeuvring flight.

Turning
Any object which is constrained to travel on a curved path has, because of its
inertia, a natural tendency to want to continue travelling in a straight line (and
will therefore fly off at a tangent if released). To keep it on its curved path, a
force must be continually applied to it toward the centre of the cwve. From
Chapter I, we know this force as centripetal* force (CPF), which can be
measured by the formula: Wv 2
CPF =
gr

(*centripetal means centre-seeking)

Turning an aircraft is analogous to whirling an object tied to a stling. If the object


is suspended vertically, your hand supplies a lift force which is equal and
opposite to the weight of the object-under conditions which are similar to
straight and level flight. However, if you whirl the object on a curved path a
much greater lifting force must be provided along the string so that its vertical
component is sufficient to support the weight, and the horizontal component-
the centripetal force-is sufficient to keep the object turning. The total lifting
force is greater than the weight of the object and you can certainly feel this
increase.

L
--------
f------------: L
I
I I
I
this component ""I I
counteracts weight I
I
I
- - -)!J

\___ this component


is centripetal force:
it holds the object on its curved path

Fig. 11-1. Centripetal force pulls


w w an object into a tum.

Principles of Flight Turning 11-1


To turn an aircraft, a force toward the centre of the turn must be generated. This
is done by banking the aircraft and tilting the lift force so that it has a horizontal
component. It is this horizontal component of lift-the centripetal force-which
pulls the aircraft into a turn and keeps it turning.

The Level Turn


In straight and level flight, the lift force counteracts the weight of the aircraft. To
turn in level flight, the wings must generate increased lift to provide both the
centripetal force for the turn and a vertical component to counteract the weight.
If the lift is not increased as the aircraft rolls into the turn, the vertical component
of lift opposing the weight will reduce and the aircraft will descend. To generate
this increase of lift at the same speed, the angle of attack of the wings must be
increased by applying backward pressure on the control column.

L
A.----- L +--------------- L
1 ,

I
I
,
- - - - - - - } > - ''

w w w
30° bank angle 60° bank angle

Fig. 11-2. The steeper the angle of bank in the turn,


the greater the lift required from the wings.

The steeper the level turn, the greater the lift force required to maintain the
vertical component of lift equal to the weight. Note also that the steeper the
angle of bank, the greater is the centripetal force and, for any given speed, the
smaller the radius and the 'tighter' the turn will be.

Use of the Controls

The ailerons are used to roll into the turn and to control the angle of bank. At the
same time as the aircraft is rolled, "back pressure is placed on the control column
to increase the angle of attack and increase the lift. The steeper the angle of
bank, the greater the back pressure required to hold the aircraft in the turn.

As has already been explained in Chapter 7, while the ailerons are deflected and
the aircraft is rolling, rudder pressure is used in the same direction as the roll to
prevent any adverse yaw and keep the aircraft balanced. When the roll is
stopped and aileron is being used to control the bank angle, less rudder pressure
will normally be needed to keep the coordination ball centred.

In summmy, during a level turn the ailerons are used to roll the aircraft and
control the angle of bank. In coordination with the ailerons, elevator is used to
control the lift and to keep the aircraft level; with rudder being used to maintain
balance.

11-2 Turning The Commercial Pilot Series


Drag Increases in a Turn

A consequence of increasing the angle of attack of the wings to produce the


centripetal force necessal)' for the turn, is that drag (mostly induced drag) also
increases. If a constant airspeed is to be maintained in a level turn, power must
be increased to provide extra thrust to offset the increase in drag. The steeper
the angle of bank in the turn, the greater the increase in drag. If the power is not
increased, the airspeed will decrease. In medium-banked level turns, the small
reduction in airspeed caused by the relatively small increase in drag is usually
accepted, with no change being made to the power setting. The Joss in speed is
soon recovered once the aircraft is returned to straight and level flight. In steep
turns (45° angle of bank or higher) it is usual to increase power as the turn is
entered to provide the extra thrust needed to offset the potentially much greater
reduction in airspeed.

Load Factor
In straight and level flight, the load on the aircraft is the weight and, with lift
equal to the weight (i.e. L = W), the load factor is said to be I. The pilot
experiences the same force from the aircraft seat as if sitting at a desk on the
ground-i.e. equal to his or her normal weight. It is felt as I 'g' or I times the
normal acceleration due to gravity.

As we have seen, in a turn the lift force must be greater than the weight so that it
can produce a vertical component to oppose the weight, and a horizontal
component (centripetal force) to provide the necessa1y acceleration toward the
centre of the curve. The reaction to this increase in lift is felt as an increase in
the load-the aircraft and evel)'thing in it has now undergone an apparent
increase in weight. The pilot now experiences a greater force from the aircraft
seat and feels heavier than before.

The origin of this increase in the loading-the apparent increase in weight-is


shown in Fig. I I -3. It is the resultant of the weight and the reaction to the
centripetal force imposed (sometimes called 'centrifugal force'). This resultant
force is equal and opposite to the lift and is the same force which is experienced
by say, a racing car travelling at high speed around a steeply-banked cmved
track (and which prevents it from sliding to the bottom of the banked surface).

component supporting force component


supporting W + from the ground 'L' supporting W L

~I \.
I
'centrifugal' 'centrifugal'
reaction centripetal force reaction centripetal force
,-<(------ ------1>- 1<11!(-------- -------1>-
'
'
'
'
'
'
'
'
resultant of
w resultant of
w
centrifugal reaction and weight centrifugal reaction and weight
i.e.' loading' i.e.' loading'
(equal and opposite to 'L') (equal and opposite to L)

Fig. 11-3. The load imposed when manoeuvring is the resultant of the reaction to centripetal force
and the weight.

Principles of Flight Turning 11-3


A measure of how much the loading of an aircraft is increased during
manoeuvres is given by the load factor-which is the ratio of the lift force being
generated by the wings at any given moment compared to the weight of the
aircraft.
LIFT BEING GENERATED
LOAD FACTOR =
WEIGHT

In a level turn with 60° angle of bank, the wings produce a lift force which is
double the weight, i.e. L = 2W. This means that the force on the wings (the
loading) is doubled when compared with straight and level flight. The force on
everything else in the aircraft is also doubled, and the pilot experiences a force
from the seat which is twice the normal weight, i.e. 2'g'. The load factor in these
circumstances is said to be 2. This is true for all 60° bank level turns, irrespective
of airspeed, rate of turn or size of aircraft.

In level turns at small bank angles, the load factor is not very high. At bank
angles above 60°, the lift generated by the wings must increase greatly if the
vertical component is to counteract the weight, othe1wise altitude will be lost.
Any increase in lift results in an increase of wing loading, as shown in the CU!Ve
of load factor plotted against bank angle for a level turn (Fig. 11-4).

3
LOAD
FACTOR
(L/W)

or lgl
2

1
0 10° 20° 30" 40" 50" 60" 70" 80"
BANK ANGLE

Fig. 11-4. Load factor versus bank angle 1n a level turn.

Notes:

In a 30" bank level turn, the load factor is 1·15'g', i.e. the wings are producing
ISo/o more lift than in straight and level flight, and the pilot will feel 15o/o heavier.

At 60° bank angle the load factor is 2. The wing is required to produce lift equal
to twice the weight to maintain altitude. The pilot will feel a force of 2'g' i.e. he
or she will feel twice as heavy as normal.

At a little over 70° bank angle, the load factor is 3.

11-4 Turning The Commercial Pilot Series


When the aircraft load acts in the direction which the weight normally acts, the
load factors are positive. It is possible to have a load factor which is less than
one-for example, when the control column is pushed forward in what is called
a bunting manoeuvre and the lift is reduced to below the weight. If, for
example, lift is reduced in this manner to 50% of the weight, the load factor will
be 0·5. If the bunt is strong enough, it is possible to achieve a zero load factor for
a short period. This is the state of so-called 'weightlessness' where anything in
the aircraft (including pilot and crew) tend to float up out of the normal position
if not restrained. The same condition can be experienced for brief periods
during severe turbulence.

Negative load factors occur whenever the load on the aircraft and the lift vector
act in the reverse direction to that in which they would normally act. For
example, an aircraft has a load factor of minus I when it is rolled 'upside down'
and is flown level in the inverted position. For the aircraft to remain level when
it is inverted, the wings must generate sufficient lift to counteract the weight, but
at a 'negative' angle of attack when compared to normal upright flight. Negative
load factors may also be experienced for brief periods in severe turbulence.

Stalling in Turns
Consider an aircraft flying at a constant !AS in a level turn at successively
increased angle of bank (I 0', 20', 30', and so on). In each successive turn with a
steeper bank angle, the angle of attack must be increased to provide the
increase in lift needed for the turn. As the aircraft is maintaining a constant
speed, this means that at each successive angle of bank, the angle of attack is
closer to the stalling angle than it was previously. Eventually, an angle of bank
will be reached where, to provide the lift needed for the turn, the angle of attack
will have reached the critical angle. Any further attempt to increase the angle of
bank and rate of turn will result in the aircraft stalling (at the speed which has
been maintained). Thus the stalling speed in the turn has been increased in
concert with the increase in the angle of bank and the increase in lift needed to
sustain the turn.

PERCENTAGE
INCREASE
IN STALL
SPEED

) load factor
or

0 15' 30' 45' 60' 75'


BANK ANGLE

Fig. 11-5. Percentage increase in stalling speed versus bank angle.

Principles of Flight Turning 11-5


The increase in the lift needed for the turn results in an increase in the load factor,
and, as we saw in Chapter 8, the amount of the increase in stalling speed in a turn
(or any manoeuvre) is directly related to load factor. You will recall that:
stall speed (in the turn or manoeuvre) = basic stall speed x --lload factor.
Few light aircraft are fitted with an accelerometer to give a cockpit reading of the
load factor (or 'g') being applied at any time. However, as has been explained in
the previous section, the load factor and angle of bank in a level turn are directly
related. The artificial horizon can thus be used to give an indication of the loading
on the aircraft. The increase in stalling speed (above the straight-and-level stalling
speed in the same conFig.uration) with increased angle of bank is shown in Fig.
11-5. Note that this is the same graph as Fig. 11-4, except that, against the vertical
axis, load factor has been replaced by the increase in stalling speed (which is, of
course, the square root of the load factor).
Although strictly speaking, Fig. 11-5 applies only to level turns, it can for all
practical purposes be taken to apply to turns in a normal climb or descent-any
difference in stalling speed at a given angle of bank in climbing level or
descending flight will be very small. Note that by comparison with the straight-
and-level speed in the same conFig.uration, stalling speed is increased by:
• 7% at 30' bank angle;
• 19% at 45' bank angle;
• 41 o/o at 60'bank angle; and
• 100% at 75' bank angle.
Hence, an aircraft with a basic stalling speed of SOKIAS will stall at a little over
53KIAS at 30' bank, about 59KIAS at 45' bank, and 71KIAS in a turn with 60' bank
angle. In medium-banked turns at normal cruising speeds, the small increase in
stalling speed is usually of no concern. In steep turns, power is normally increased
to offset the increase in drag-to limit the reduction in speed which results and to
provide a bigger margin above the increased stalling speed. At altitude, provided
any 'g' limitations are observed, the aircraft can normally be flown safely in a steep
turn right to the verge of the stall where the onset of pre-stall buffet signals the
need to reduce the angle of bank and or 'g' loading. There is however, a need for
heightened awareness of stalling speed during slower-speed turns near the
ground. The main concern here, is that the aircraft is being operated with a
relatively small margin above the stalling speed and any untoward increase in
angle of bank and loading can quickly bring the wings to the stalling angle of
attack.

Gusts and turbulence can also have an effect in producing an uninvited increase
(or decrease) in angle of attack. Where angle of attack is increased for this
reason, the load factor and stalling speed are also increased, Accordingly, in
gusting, turbulent conditions it is normally recommended that higher approach
speeds be used so that a safe margin above stalling speed is maintained.

Turning Performance
Turning perfmmance is comprised of the rate of turn (i.e. the rate of change of
the aircraft heading) and the radius (i.e. 'tightness') of the turn. There are two
elements which determine the turning performance-load factor and TAS. In a
level turn, as explained previously, the load factor is always associated with a
particular angle of bank. Regardless of aircraft type, speed or weight, the load
factor in a 60' banked level turn for example, will always be 2. Hence, from the
pilot's point of view, turning performance is determined by the angle of bank and
airspeed in the turn.

11-6 Turning The Commercial Pilot Series


For a given airspeed, as the angle of bank is increased, the rate of turn
increases and the radius of turn decreases. (Fig. 11-6)
L steeper bank angle L

low bank angle


+

I same lAS I

small turning force larger turning force


large radius smaller radius
low rate of turn higher rate of turn

Fig. 11-6.

For a given bank angle, the radius of turn increases, and the rate of turn
decreases as speed is increased. (Fig. 11-7)

Fig. 11-7.

larger radius and


lower rate of turn

same
bank angle

The Standard Rate (or Rate 1) Turn

There is sometimes a requirement, particularly in instrument flying, to execute a


standard rate turn of 3° per second or 180° per minute. The angle of bank which
will achieve this depends on airspeed and an easy way of mentally calculating
the required bank angle in a light aircraft is given by dropping the last digit from
the airspeed Fig.ure and adding seven. For example, if the airspeed is 110 knots,
the angle of bank required for a rate I (or standard rate) turn is II + 7 = 18°. If
the airspeed is 130 knots, the angle of bank for a rate I turn is 13 + 7 = 20°.

The Effect of Altitude

The force used to turn an aircraft-the horizontal component of lift-is an


aerodynamic force, and as such, depends on lAS. On the other hand, the force
required to achieve any given radius and rate of turn-the centripetal force-is
measured by WV'/gr and therefore depends on V, or TAS.

Imagine an aircraft in a steady turn at low level where the conditions just
happened to be 'standard' and the lAS was equal to the TAS. The centripetal
force produced by the wings could be measured by the formula WV2/gr and
would amount to some Fig.ure, let's say I 00 units. Now consider the aircraft in
the same turn at altitude with the same angle of bank and lAS. As the lAS is the
same, the same centripetal force of 100 unit is being produced by the wings.
However, as the TAS (V) has increased with the gain in altitude, the centripetal
force (WV2/gr) required to turn on the same radius and at the same rate will also
have increased-say to II 0 units. As the centripetal force being produced by the
wings has remained at the same I 00 units, it is less than that required to achieve
the same performance at the higher altitude.

Principles of Flight Turning 11-7


Fig. 11-8.
The effect of altitude on the turn.

HIGH ALTITUDE

larger radius and


lower rate of turn

same lAS and bank angle


same horizontal component of lift

smaller radius and


higher rate of turn

Hence for a given !AS and angle of bank, the rate of turn decreases and the
radius increases as altitude is gained. Or put another way, for a given lAS, an
increase in angle of bank will be required to maintain a turn at the same radius
and rate, as altitude is gained.

The Effect of Weight

Weight affects only the maximum rate and mtmmum radius which can be
achieved by an aircraft in a turn. Apart from this, the turning performance of an
aircraft is independent of its weight. At a given lAS and load factor (angle of
bank), a high-weight aircraft will achieve the same rate and radius of turn as a
low-weight aircraft at the same lAS and angle of bank. The difference is that the
higher weight aircraft requires a higher angle of attack (at the same speed and
angle of bank) to generate the lift force necessary to overcome the weight and
provide the centripetal force for turning. If the aircraft is heavy, it will reach its
stalling angle of attack sooner (at a lower angle of bank) than when it has less
weight, and it will thus have a poorer maximum rate/minimum radius of turn
capability.

A----- L .-----
1
L

Both aircraft 1 : I Higher weight:


SAME: I : I HIGHER:
lAS I I angle of attack
angle of bank stalling speed
load factor (LNI)
rate of turn
radius of turn

w w

Fig. 11-9. The effect of weight on the turn.

11-8 Turning The Commercial Pilot Series


Steep Turns
A steep turn is one in which the bank angle is 45° or greater. In a medium turn,
(less than 45° bank), power is not normally increased to offset the increase in
drag, with the small loss in speed being accepted for the duration of the turn.
However, in a steep turn, the increase in drag at the higher angles of attack
being used, becomes significant. It is usual therefore to increase power as the
aircraft is rolled into a steep turn, to offset the potentially much greater loss of
speed.

In addition to combating the higher drag, the use of a high power setting also
helps the aircraft to turn and provides a a wider operating margin above the
stalling speed. At the higher angles of attack involved in a steep turn, a
component of thrust increasingly acts in the same direction as the lift (Fig. 11-1 0
refers). When the aircraft is banked with high power applied, this helps to
reduce the aerodynamic lift force required from the wings, resulting in a lower
angle of attack and a lowered stalling speed.

Fig. 11-10 In addition to helping overcome the drag, the use of high power in a steep turn
helps the aircraft to turn and reduces the stalling speed.

The steep turn is a high performance manoeuvre which requires good


coordination, positive control and accurate handling. In a steep turn, the aircraft
is being operated nearer to its manoeuvre limits and the margin between the
aircraft speed and the stalling speed is reduced. For these reasons, steep turns
at low altitude should be avoided where possible.

Maximum Rate and Minimum Radius Turns


Minimum Radius

Consider Fig. 11-11 which depicts an aircraft being flown in a level turn at the
verge of the stall (wings at CLmax) and at the maximum angle of bank which can
be sustained without stalling.

Fig. 11-11. L
..----~,
I I '
I I
(a) I (b)

low lAS higher lAS


15, a (CLmax) 15, a (CLmax)
Max. bank angle
l small
Max. bank angle
I large radiu
I low rate of turn
t
w w

Principles of Flight Turning 11-9


The diagram at (a) represents the maximum centripetal force which can be
generated at the relatively low speed. The aircraft will thus be turning on the
smallest radius which is possible at that speed. If the angle of bank and angle of
attack are increased in an attempt to increase the centripetal force (and thereby
reduce the radius of the turn) the aircraft will stall.

To have a smaller radius of turn, the CPF must be increased and, with the wings
already at CLmax, the only option for achieving that is to increase speed so that
the lift and angle of bank can be increased without stalling. This is shown in
diagram (b).

There is an apparent contradiction here. For, as we saw previously, an increase


in speed tends to increase (not decrease) the radius of the turn-and it was
certainly true for a fixed angle of bank. However, when a higher bank angle can
be sustained as a result of the increase in speed, the advantage gained through
the increase in CPF outweighs the 'disadvantage' (in terms of increased radius
of turn), which follows that increase in speed.

When an aircraft is flown in a level turn with the wings at the angle of attack for
CLmax, it can be shown that the radius of turn:
where VB = basic stall speed, and ~ = angle of bank.
VB'
g tan~

At any given time, VB and g can be considered as constants, hence the radius of
turn (wings on the verge of the stall) is inversely proportional to the tangent of
the angle of bank. Thus, for minimum radius the angle of bank needs to be as
high as possible, and for that to be sustained in a level turn, the speed-at least
in theory-also needs to be as high as possible.

Fig. II -12 shows a curve of the theoretical radius of turn against lAS. It indicates
that the radius of turn (turning as hard as possible) will decrease with increased
speed, with the minimum radius being achieved at infinitely high speed. In
reality, practical considerations intervene:
• The minimum radius will
be achieved at the design
manoeuvring speed VA. If I
the aircraft is flown on the I

iI
1 turn radius with wings at
verge of the stall above CLmax and with max. bank
this speed, the aircraft 'g'
limits will be exceeded. 1 practice
radius
With the restriction in
load factor (and i _;;;imposed)

consequently on angle of
bank) above VA, the
I 1- - -
radius of turn increases 1 theory
I ---
again, as indicated. 1 I
I I
• Many light aircraft do not .. t
have sufficient excess stall VB VA lAS
power available which
can be used to offset the Fig. 11-12. Min. radius versus airspeed.
drag in a sustained
minimum-radius turn at
VA. The aircraft will usually settle in the turn at a slower speed and,
therefore, at a greater radius than it is possible to achieve at the VA speed.

11-10 Turning The Commercial Pilot Series


NOTE: Design manoeuvring speed VA is explained in more detail later in this
chapter.
Maximimum Rate

Rate of turn is given by speed over radius (VIr). Hence, for maximum rate in the
turn, we need to have the highest speed and the smallest radius. It can therefore
be seen that the requirements for a maximum-rate turn are essentially the same
as for a minimum-radius turn. The aircraft must be flown on the verge of the
stall, with the highest speed and angle of bank which can be sustained. Once
again, practical considerations mean that the best speed for a maximum-rate
turn is VA. Above VA, the load factor and angle of bank must be restricted,
resulting in a lower rate of turn. Below VA, the angle of bank and load factor are
naturally restricted by the stall boundary, and the maximum rate of turn which
can be achieved reduces in concert with reducing speed.

Practical Application

The maximum rate/minimum-radius turn involves flying at the angle of attack for
CLmax which incurs a large increase in drag (mostly induced drag) particularly if
the airspeed is high. To counter the drag, it is essential that the maximum
allowable power/thrust is used. In most light aircraft, this means opening the
throttle fully.

Thus, to execute a max rate/min radius turn, the aircraft is rolled to the steepest
possible angle of bank as full power is applied and back pressure increased to
bring the wings to CLmax or just below-normally just to the onset of pre-stall
buffet. (If the speed is above VA, care must be taken not to exceed the
maximum 'g' permitted). If the speed falls off in spite of full power, it will be
necessary to reduce the angle of bank while staying on the verge of the stall,
until the angle of bank is found where the aircraft can maintain height at a
stabilized speed.

As power becomes the limiting factor in this manoeuvre, it follows that anything
which increases the drag or reduces the lift/drag ratio, must degrade the
effectiveness of the excess power being applied to the manoeuvre. Thus the use
of flap, even optimum flap, is counterproductive in a sustained max-rate/min-
radius turn although it is accepted that a temporary advantage will be gained for
a short period as the flap lowers.

If the available power is not sufficient to maintain the required manoeuvre


speed, and altitude is not a problem, it is possible to sustain the turn
performance by trading altitude for speed during the turn. If this is done, care
must be taken to keep the bank angle and the rate of descent under control.

Effect of Power
It can be seen from the foregoing paragraphs that turning performance (like the
climb) is ultimately dependent on the excess of power available over power
required for level flight.

In a level turn at a given lAS, the drag is increased by comparison with the same
speed in straight and level flight. There is thus a requirement to increase the
power/thrust if the same speed is to be maintained in the turn. If the increase in
angle of bank and load factor is continued, full power will eventually be needed
to offset the increase in drag at the given speed. Hence, the greater the excess
of power available, the greater the bank angle and the better the turning
performance which can be obtained.

Principles of Flight Turning 11-11


Light training aircraft generally have a relatively low excess of power available
over power required. Hence, anything which reduces or 'eats into' this excess of
power available will be significant in the effect it has on reducing the ability of
the aircraft to turn. From chapter I 0, we saw that excess power was reduced
by:
• increased weight;
• increased altitude; and
• flap and undercarriage extension.

These same factors will therefore reduce the turning performance of an aircraft.

The Effect of Wind


For an aircraft turning at constant angle of bank and airspeed, the radius of turn
in the airmass will also be constant i.e. the llightpath will be circular. If a wind is
blowing, the llightpath is no longer circular with reference to the ground, as the
aircraft and the airmass drift downwind from the reference point. (Fig. 11-13).

/
/
I
I constant bank angle
flightpath in wind

I
\
\ flightpath in the airmass

\.
' - Fig. 11-13. Ground track for a constant-bank-angle turn in a wind.

If the angle of bank is maintained, the ground radius of the tum will change,
becoming greatest where the groundspeed is highest (aircraft headed directly
downwind) and smallest where the ground speed is lowest (headed directly
upwind).

To maintain a constant-radius turn around a ground feature, the centripetal force


(CPF) must be adjusted to allow for the differences in groundspeed as the
aircraft flies around the turn, This means that the higher the groundspeed the
higher the CPF (and angle of bank) required, and vice-versa.

For the aircraft depicted in Fig. 11·14, the groundspeed will be least at position B,
and highest at D. Therefore the angle of bank needs to be least at B and greatest
at D to provide a constant ground-radius tum. This means that the angle of bank
will need to increase between B and C, and decrease between D and A.
However, since the effect of the wind at Cis to assist the turn, but not assist at A,
the angle of bank at C needs to be smaller than at A.

11-12 Turning The Commercial Pilot Series


c

CPF

Fig 11-14.
A constant radius turn
around a ground feature.

Climbing Turns
Rate of Climb
As for the level turn, a climbing turn requires a tilted lift vector and increased lift
to provide the centripetal force needed to turn. There is an increase in drag as
well as lift. To offset this increase in drag and maintain the climbing speed,
some of the excess power being used in the 'straight ahead climb', is absorbed
by the turn. With less excess power available for the climb, the rate of climb is
reduced. The nose attitude required to maintain the climb speed will also be
lower than in the straight climb.

If that same speed is maintained, the reduction in rate of climb in the turn is
proportional to the angle of bank used. Eventually an angle of bank will be
reached where all of the excess power is being absorbed by the turn-the rate of
climb will then be reduced to zero and the aircraft will be turning in level flight.

Hence, the ability of an aircraft to climb and to turn at the same time depends on
the amount of excess power available at the climbing speed. At high altitudes,
where the excess power available for the straight climb is reduced, the
capability of the aircraft to turn and maintain any appreciable rate of climb may
be poor.

Tendency to Overbank
Once established in a climbing turn, there is a tendency for the aircraft to
continue rolling into the turn-i.e. to overbank. This is caused by the outer wing
in the turn having a higher effective angle of attack than the inner wing. Both
wings travel through the same vertical distance in the climb, but the outer wing
travels on a wider arc and therefore through a greater horizontal distance. This
results in the difference in angle of attack as shown in Fig. 11-15.

The tendency to overbank in a climbing turn can be quite pronounced in some


light aircraft. With these aircraft, once the desired angle of bank has been
established in a climbing turn, it will be necessa1y to hold a steady aileron
pressure in the opposite direction from the turn-or to 'hold off bank' as it is
called-to prevent the bank angle frorn increasing beyond the desired angle.

Principles of Flight Turning 11-13


gain
in
altitude

horizontal distance travelled

angles exaggerated for purposes of explanation


\

I
Fig. 11-15. The tendency to overbank I
in a climbing turn.
/ /
/ /
/

/ 7
z I
I \
\

Descending Turns
Rate of Descent
In a descending turn, drag increases with bank angle for the same reason as in
level and climbing turns. If the speed for the straight descent is maintained in
the turn, the rate of descent will increase. This can be compensated for by
increasing power. However, if the turn is steep, the rate of descent can increase
relatively quickly and will require a large increase of power and possibly also a
smart reduction in angle of bank to keep it in check. For this reason, steep
descending turns near the ground should be avoided.

Tendency to Underbank

""
""
Fig. 11-16. The tendency to underbank
in a descending turn.
" " '\
" \
"'\ I\
l I
I /
/ /
~~~~~;;;;;;--/
loss
in
angles exaggerated for
altitude purposes of explanation

horizontal distance travelled

11-14 Turning The Commercial Pilot Series


In a descending turn, there is a tendency for the aircraft to roll out of the turn.
The reason is similar to that which applies to the overbanking tendency in a
climbing turn but the effect of the outer wing travelling through a greater
horizontal distance is reversed. In this case, as shown in Fig. I 1-16, it is the inner
wing which has the higher angle of attack, and which therefore generates more
lift. Again, this tendency to roll out of the turn as a result may be more
pronounced in some light aircraft than it is in others. If it is evident, it can be
easily countered by holding sufficient aileron pressure into the turn-i.e. by
'holding on' aileron.

Manoeuvring in the Vertical Plane


So far, we have covered the forces acting in a turn, i.e. with the aircraft following
a curved or circular path in the horizontal plane. It is possible for an aircraft to
follow a curved path in the vertical plane-in what is called a looping
manoeuvre.

It is instructive to consider the forces acting in a perfectly circular loop at


constant speed-although, in practice, it is a difficult manoeuvre to achieve. To
follow any circular path, whether a turn or a loop, a constant steady acceleration
must be provided toward the centre of the circle. This is not particularly difficult
in a turn, where the weight of the aircraft acts consistently at right angles to the
centripetal force provided by the horizontal component of lift. In a loop
however, the angle between the weight and the centripetal force (lift) are
constantly changing-with weight acting in the opposite direction to lift at the
bottom of the loop, and in the same direction when the aircraft is inverted at the
top.

Fig. 11-17. The forces acting c


in a loop.

--+L L~ D

W+
\~ L
)fw
~~net
D

2'g')
3'g'

finish start
A
w

Principles of Flight Turning 11-15


The forces acting at each 90° position of a constant-speed circular loop are
shown in Fig. 11-17. We have chosen a constant net acceleration toward the
centre of 2'g'. To achieve this;

• The load factor at the start of the loop (point A), must be 3'g' (2'g' net after
overcoming the weight).

• At B and D, when the aircraft is vertical, the lift acts at right angles to the
weight and the wings must produce only the required acceleration of 2'g'.

• At the top of the loop (C), the aircraft is inverted and, since weight now acts
in the same direction as lift, the acceleration from the wing is reduced to a
load factor of 1'g'-i.e. the lift plus the weight provide the required
acceleration of 2'g'. It is worth noting that at position C, the pilot would feel
a 1'g' force from the seat as if he or she were flying straight and level.

The difficulty in flying a constant-speed loop can be appreciated if the thrust and
drag forces are now considered. At the start of the loop, T=D as in straight and
level flight. If speed is to be maintained on the way up in the manoeuvre, there
must be a large increase in thrust so that by the time position B is reached, it is
sufficient to balance both the weight and the drag. (Few aircraft have this
amount of excess thrust available). At the top (C), the requirement for thrust
reduces so that once again it balances the drag.

From the inverted position at the top of the loop, there is a need to reduce thrust
since weight and thrust begin to act more and more in the same direction.
However, by the time position D is reached, even with the throttle closed and
extra drag devices deployed (e.g. airbrakes), weight is likely to exceed drag by a
significant margin and there will be nothing the 1•9 .
pilot can do to prevent increasing its speed. To
keep the loop circular at least, a large increase in 80 kts
load factor (lift) will be required particularly over
this last 90° of the manoeuvre. 2'g' 2'g'

For the majority of light aircraft, where there is


insufficient excess power available to prevent a
large reduction in speed over the first half of the
loop and, even with the throttle closed, a large gain Fig. 11-18. 140 kts
in speed on the way down, the path followed by a
loop in practice would look something like that
depicted in Fig. 11-18. In this case, the speed, load
140 kts '
4'g'

factor, and the radius of the loop will be greatest at the bottom of the loop and
smallest at the top of the manoeuvre.

The radius of a looping manoeuvre is proportional to V2/'g' (TAS/Ioad factor).


This relationship becomes important when recovery from a dive is considered-
particularly if it is at low altitude. A dive recovery can be equated with the last
quarter of the loop described in the foregoing and it is important to realise that:

• once a steep nose-down attitude is adopted, a rapid increase in speed is


inevitable even with the throttle closed;

• the radius of the manoeuvre increases with the square of the airspeed; and

• a high load factor and a lot of altitude will be required to affect the recovery.

Do not indulge in diving manoeuvres at low altitude.

11-16 Turning The Commercial Pilot Series


Manoeuvring Limitations
All aircraft have speed and load factor ('g') limitations placed upon them to
prevent structural damage or failure due to excessive aerodynamic loads. These
limitations are set out in the Pilot's Operating Manual for the aircraft type.

Speed Limitations
Speed limitations are stated in the form of a V code and, in modern light aircraft,
the appropriate speeds are marked with coloured lines or arcs on the AS!. Fig.
11-19 gives an example of such markings. The V code which applies to a light
single-engined aircraft is:
VSO-Stalling speed undercarriage down, flaps lowered, power off.
VSI-Stalling speed undercarriage and flaps up, power off.
VFE-Maximum speed-flaps extended.
VA-Design manoeuvring speed. (Varies with weight, and not usually
marked on the AS!). VA is discussed more fully in following paragraphs.
VNO-Maximum structural cruising speed for normal operations.
VNE-The never- exceed speed. Do not exceed in any circumstances.

The white arc is between VSO and VFE and denotes the flap operating speed
range. The flaps must not be lowered or the aircraft flown with flaps down
when the speed is above this range.

The green arc extends between VS I and VNO and denotes the normal-
operating speed range. All operations in this range should be safe in all normal
flying conditions.

The yellow arc extends from VNO to VNE and denotes the caution range. The
aircraft should be operated in this range only in smooth air conditions.

VNE
never-exceed speed
(maximum speed
for an operations)
vso
stalling speed,
landing gear down,
flaps lowered, power off.

VS1
stalling speed,
landing gear up*
maximum structural flops up, power oft
cruising speed *(If retractable)
(for normal operations)

9r~
~,., ~e

· orrnat operating "'""'


Fig. 11-19.
Airspeed limitation markings VFE
on a modern AS/. maximum speed, flaps extended

Principles of Flight Turning 11-17


Load Factor Limitations
Load factor (or 'g') limitations are also set out in the Pilot's Operating Manual for
the aircraft type. For most light aircraft with the flaps raised, the positive 'g' limit
will be typically in the order of +3·5 to +4·0 'g', and about minus 1·5 'g'. With
flaps extended, the positive 'g' limit will normally be reduced to about +2·0.
Some aircraft may be capable of being operated in a 'utility' categ01y where,
with further limitations placed on operating weight and CG position, the positive
'g' limits may be raised slightly.

Great care must be taken to see that the speed and load factor ('g') limitations
are not exceeded. Although most limitations have a safety factor (or buffer)
included, when the aircraft is being operated near any of the limits-and
particularly at high speed-unexpected turbulence can easily impose loads
which result in overstressing and possible airframe damage.

The V-n (or V-g) Diagram


Some aircraft have their speed and load factor limitations set out in what is
called a V-n (or V-'g') diagram. Such a diagram gives a graphical representation
of the aircraft's manoeuvre envelope and the speed and load factor limitations
can be more easily related and better understood. An example of a V-n diagram
is given in Fig. 11-20 below.
+5
I FLAPS UP I
+4

+3 I<@-- A
positive stall I I
load I I
' l<r-B
factor +2 design I
(n) manoeuvring I
or speed VIm
('g') +1 VA I
I
I
0

-1
negative stall

-2

Fig. 11-20. A- Full deflection and harsh use of controls should be progressively avoided.
A V-n diagram.
B - Caution range -operation only in smooth air conditions.

NOTES:

• A V-n diagram will be valid for one aircraft weight and configuration e.g.
flaps up.

• The upper and lower curves represent the positive and negative stall
respectively. Flight in the area to the left of these curves is not possible as
the wing will be stalled. Flight within the envelope (between the curves)
below VS 1 will only be possible for a short duration and at low load factors-
basically only for a short while during some aerobatic manoeuvres.

11-18 Turning The Commercial Pilot Series


• The design manoeuvre speed (VA) occurs where the positive stall curve and
the maximum positive 'g' line intersect. Although not usually marked on the
AS!, VA is significant. It is the speed above which it is possible to 'overload'
and overstress the aircraft and it is therefore the speed in practice at which
the best rate and smallest radius of turn can be achieved. Also of
importance to structural safety, it is the speed above which full and rapid
movement of the controls should be progressively reduced, and it is the
'reduce-to' speed if significant turbulence is encountered.

• If the aircraft is flown in the area above/below and to the right of the
envelope, there will be an increasing risk of structural damage and possible
failure.

• In some aircraft with negative 'g' limits, inverted flight may not be permitted.
In this case the negative limits given in the Pilot's Operating Handbook apply
to the negative loadings which could be encountered in severe turbulence.

Further Notes on VA

The design manoeuvre speed (VA) varies depending on weight. It occurs at a


higher speed at high weight than it does at low weight. This may appear at first
sight to be the wrong way around, but if it is remembered that VA occurs in the
V-n diagram at the junction of the positive stall and the maximum load factor or
'g' line, the reason should be clear. At high weights the stalling speed is
increased-hence the speed at which the aircraft will stall with the maximum 'g'
factor applied is increased by comparison with the stall speed at low weight and
the same 'g' factor. In a sense, the aircraft is better protected from overstress at
the higher weights, because until it reaches that higher VA, it will stall before the
maximum load factor can be applied.

Most light aircraft are not fitted with an accelerometer (or 'g' meter). As
explained earlier in this chapter, the angle of bank indication on the AI can be
used to give an indirect indication of load factor being applied in a level turn.
(Remember, at 60' bank the load factor will be 2, at just over 70' it will be 3). If
the pilot knows the VA which applies at the time, he will be able manoeuvre the
aircraft right up to the pre-stall buffet at any speed below VA and be confident
that the aircraft is not being overstressed. The Pilots Operating Handbook will
normally give the VA but it will not normally be marked on the AS!. For some
aircraft the VA may be stated for different weights. A competent pilot should be
able to remember these Fig. ures and be able to extrapolate if necessa'Y to
determine the VA with reasonable accuracy at any time during flight.

Principles of Flight Turning 11-19


Review 11

I. The horizontal component of lift which holds the aircraft in the turn is
called .......................... force.

2. State the formula used for measuring centripetal force.

3. When an aircraft is turned in level flight, the lift must be increased so that its
vertical component balances the ........................ .

4. The steeper the bank angle in a turn, the greater the ............... required from
the wings.

5. The increased angle of attack in a turn not only means that lift is increased,
but also ................ .

6. The increase of drag means that in a level turn, there will be a loss in speed
unless ....................... is increased to counteract it.

7. Load factor is the .............. being generated by the wings divided by


aircraft .................... .

8. In a 30° banked level turn, the load factor is ......... and the pilot will feel 15%
heavier. In a 60° banked level turn the load factor is ........ and the pilot will
feel ...... 'g'.

9. The stalling speed (increases/decreases) in a turn. The relationship is-stall


speed in the turn ~ basic stall speed x -J .......•...•..•.••••••.•.•••••••.•...

I 0. At any given airspeed, the radius of turn (increases/decreases) as bank


angle increases.

II. State the angle of bank needed to achieve a rate I turn if the aircraft is flying
at 130 kt.

12. An aircraft turning with a given lAS and angle of bank will have a (larger/
smaller) radius and a (higher/lower) rate of turn at higher altitude.

13. Two aircraft have the same angle of bank and lAS in a level turn, but one is
at higher weight than the other. What is the difference between their rate
and radius of turn.

14. In practice, the smallest radius/maximum rate of turn is obtained at


the ................................................. speed (VA).

15. In a constant radius turn around a ground object, the highest angle of bank
will be required when the aircraft is flying (directly downwind/crosswind/
directly into wind).

16. A climbing turn is entered from the straight climb at the same airspeed and
power setting. The rate of climb will (reduce/stay the same/increase). If all
of the excess power is absorbed by the turn at a high angle of bank, what
happens to the rate of climb?

11-20 Turning The Commercial Pilot Series


17. An aircraft tends to (roll into/roll out of) a climbing turn.

18. If the aircraft is descending and a turn is entered with power and speed
remaining constant, the rate of descent will (increase/decrease/remain the
same).

19. In a descending turn, the inner wing has a ..................... angle of attack and
the aircraft will therefore tend to (roll into/roll out of) the turn.

20. On a modern light aircraft AS!, three coloured arcs and a red line are
marked. Their significance is:
(a) the white arc denotes the ........................................... speed range;
(b) the green arc denotes the ............................................... speed range;
(c) the yellow arc denotes the ......................... range. The aircraft should be
operated in this range only in .......................... air conditions;
(d) the red line denotes the ................................ .................. .for all
operations.

21. If an aircraft is flown to the pre-stall buffet below the design manoeuvring
speed VA, the aircraft 'g' limits (will/will not) be exceeded. If flown to the
pre-stall buffet above VA, the aircraft 'g' limits are (likely/unlikely) to be
exceeded.

22. VA occurs at a (higher/lower) speed as aircraft weight decreases.

Principles of Flight Turning 11-21


11-22 Turning The Commercial Pilot Series
Propellers
Introduction

A piston engine requires a propeller to convert the power output (engine torque)
of the engine into a useful straight-line force called thrust. There are two basic
types of propeller, both with similar aerodynamics-fixed-pitch and variable-
pitch (which uses a constant-speed unit). In this chapter, we discuss the basic
principles as they apply to the fixed-pitch propeller, and then we cover the
operation of constant speed (i.e. variable pitch) propellers.

Terminology
Consider one section across a propeller blade at some distance from the hub or
the centreline of rotation (Fig. 12-1 ). The blade section is a cambered aero foil
shape and it has a leading edge, a trailing edge and a chord line just like any
other aerofoil. The cambered (lifting) side of the blade is called the blade back
and the flatter side the blade face. The angle which the chord line makes with
the plane of rotation is called the propeller blade angle. Note:
• For convenience, the chord line is normally taken to be the blade face. The
actual chord line will depend on the cross-sectional shape of the blade.
• The blade angle, as we shall soon see, varies from a large angle at the blade
root near the hub, gradually decreasing to a much smaller angle at the
propeller tip.

DIRECTION
OF FLIGHT

Fig.12-1.
Propeller terminology.

Basic Principles
Since it is an aerofoil, the propeller blade generates an aerodynamic force in the
same manner as any other aerofoil being moved through the air. In normal
operation, when the propeller blade is rotated, the acceleration of the airflow
over the fmward cambered surface causes a reduced static pressure ahead of
the blade. At the same time, when the blade has an angle of attack with the
oncoming airflow, the flow is retarded by the blade face resulting in an
increased static pressure. One of the results of these changes in pressure
around the blade is that a fmward thrust force is generated which pulls the
aircraft along.

Principles of Flight Propellers 12-1


The development of thrust can also be considered in terms of the rearward
movement of a mass of air. The changes in pressure around the rotating blades
causes air to be drawn into the propeller disc (the circular area swept out by
the blades) and accelerated in a roughly cylindrical area toward the rear (the
slipstream). This rearward acceleration of the mass of air in the slipstream
produces a forward reaction force on the blades in accordance with Newton's
Third Law. The magnitude of this forward reaction force is dependent on the
mass of the air and the acceleration it is given toward the rear. For a fixed-pitch
propeller rotating at a given rpm, the lower the air density the less the mass of air
which is accelerated rearward, and the lower the effectiveness of the propeller.

Factors Affecting Airflow Across the Blade Sections


Consider again just one of the propeller sections. If the aircraft is stationary with
the propeller rotating, the direction of the airflow approaching the blade section
will be in the plane of rotation and opposite to the direction of rotation. The
speed of the flow across that section will depend on its radius of rotation. Those
sections further out from the centre of rotation will experience a faster flow than
those closer in. And, the higher the rpm, the faster will be the flow across any
given blade section. Hence, the first two factors affecting the flow across a given
section are, the radius of the section and the speed of rotation.

Fig.12-2.
The airflow across a blade 1200 RPM 2400 RPM
section depends on propeller
rpm and the radius of the
section.

Once the aircraft begins to move forward, the airflow approaching the propeller
section will also develop a component opposite to the direction of that motion.
When this forward motion is combined with the rotational motion of the
propeller, a resultant velocity vector for a given blade section through the air can
be obtained, as shown in Fig.l2-3. The angle between this resultant velocity
vector and the plane of rotation of the propeller blade is called the helix angle
or the pitch angle or the angle of advance.
helix angle = pitch angle = angle of advance
As a result of this combined rotational and forward velocity, each propeller blade
section follows a 'corkscrew' path through the air called a helix. The relative
airflow of each blade section is directly opposite to its own particular helical
path through the air. The angle between the chord line of the blade section and
the relative airflow is its angle of attack. Note that the angle of attack plus the
helix angle (pitch angle) make up the blade angle.

When the aircraft is in flight, each propeller blade section has the same forward
velocity component. What differs, however, is the rotational component of
velocity-the further each blade section is from the propeller shaft, the faster it is
moving. If the blade angle was the same along the whole length of the propeller
then the angle of attack would continuously increase toward the tips. Under
most conditions, this would result in the outer sections being operated at an
angle of attack greater than the stalling angle.

12-2 Propellers The Commercial Pilot Series


helix
~
lJ ~\
(pitch) ./"1 '-'--angle of attack
angle :
helical (corkscrew) path '
1 relative airflow
of one blade tip ''
'' resultant velocity
velocity in :
of the propeller
the planet
of rotation 1 blade section

''
Fig. 12-3. Each propeller section follows its own helical path. '
L_ ___ ..,... __

velocity in
direction of flight

Each blade section has an angle of attack at which


it is most efficient in generating thrust. To enable
the propeller as a whole to operate efficiently at a blade angles
given airspeed and engine (propeller) rpm, then it exaggerated
for purposes of
must be designed to have the most efficient angle demonstration
of attack along its whole length, at that airspeed A
A section
and rpm. To achieve this, all propeller blades must rotational
relative
have a 'twist'-or reduction in blade angle from hub Airflow
velocity
to tip-built into them. This is known as blade
twist or helical twist.
'
Fig. 12-4 demonstrates the need for blade twist. A
B section
i1' 8 i
relative
represents a propeller section near the middle of rotational ! airflbw
the blade; B one near the tip. Although both velocity !
sections travel the same distance forward in given '
time, the higher rotational velocity toward the tip i'
''
'!:: ___________ _
sections reduces the pitch angle and thus, to
maintain the required efficient angle of attack along
A and B same!
the length of the blade, the blade angle must be forward velocity
progressively reduced toward the tip.
Fig. 12-4.

Forces Acting on a Blade Section


When an airflow is established around an aerofoil, the consequent pressure
changes generate an aerodynamic force on the aerofoil. In the case of a wing,
we resolve this total reaction into two components, one perpendicular to the
relative airflow-lift, and the other parallel to the relative airflow---drag.

Principles of Flight Propellers 12-3


For the propeller however, where the relative airflow
differs for each blade section, and where we are mainly
concerned with generating thrust and not lift, it is no
longer appropriate to talk in terms of lift and drag. Hence,
it is more convenient to resolve the total reaction force
into:

• one component perpendicular to the plane of


rotation called thrust; and
• one component in the plane of rotation called the
propeller torque force.

For our purposes, we can generally assume the direction


Fig. 12-5. The forces acting perpendicular to the plane of propeller rotation to be the
on a propeller blade.
same as the direction of flight, and therefore, for the thrust
to be considered as acting in the direction of flight.

The propeller torque is the resistance to motion in the plane of rotation.

For a wing, drag must be overcome to provide lift. For a propeller, the propeller
torque must be overcome or balanced by the engine torque for the propeller to
provide thrust. Opening the throttle increases engine power and engine torque,
causing the propeller to rotate faster-until propeller torque builds up to the point
where it comes into balance with engine torque and rpm stabilizes.

NOTE: If an aircraft with a fixed-pitch propeller is put into a dive, the relative
airflow is changed and, because of the higher forward speed, the angle of attack
is reduced over all propeller blade sections. Propeller torque is reduced and
engine rpm will increase even though the throttle may not have been moved. Be
aware that maximum rpm may be exceeded and power should be reduced as
necessary to prevent this from happening.

The RPM/Airspeed Relationship


Consider again one section of a well-designed fixed-pitch propeller blade-at say,
about 75% of the blade radius where, as we will see shortly, the thrust is produced
most effectively. As illustrated at Fig. 12-6, if the propeller rpm is constant, then
the direction of the relative airflow and the angle of attack will be determined by
the forward velocity.

As airspeed increases, the angle of attack of a fixed-pitch propeller blade at


constant rpm will decrease. At some high airspeed , the angle of attack of the
blade will be so reduced that little or no thrust will be produced. Hence, for a
given rpm, there will only be one forward velocity (true airspeed) at which
the fixed-pitch propeller will operate at its most efficient angle of attack.

The designer chooses a fixed-pitch propeller which has a best-efficiency airspeed/


rpm combination which matches the tasks for which the aircraft is designed. For
an aircraft designed to lift heavy loads off short runways and to operate at fairly
low airspeeds-e.g. for agricultural work-a fine-pitch propeller (with a small
blade angle) is most suitable. For an aircraft designed to cruise long distances at
reasonably high speeds, a coarser pitch propeller (i.e. with a larger blade angle)
will be more suitable.

12-4 Propellers The Commercial Pilot Series


different forward velocity same forward velocity
same rpm different rpm

-8t-!QW forward velOCity


high'&~gle of attack ·~thigh rpm
high.:angle of attack

at low rpm
at high forward velocity low angle of attack
low angle of attack
rpm

~
same
rpm
relative
airflow

high
rpm
low

forward velocity same


forward
velocity

Fig. 12-6. Fixed-pitch propeller: angle of attack varies with forward velocity,
and with rpm.

Effective Blade Sections


The most effective sections of the propeller blade in producing thrust are those
which lie between 60 and 90% of the radius. The greatest useful thmst is
produced at approximately 75% of the blade radius. When the blade angle of a
propeller is quoted, it usually refers to the 75% station for this reason-station
meaning position on the blade.

Near the hub, the propeller sections must be thick for structural strength, which
may interfere with effective aerodynamic design. There is also the interference
to the airflow from the engine and associated structures. Apart from this, the
main reason for the inner propeller sections being less effective is that, of
necessity, they must have a relatively high blade angle. This effectively tilts the
local aerodynamic total reaction (TR) more toward the plane of rotation as
shown in Fig. 12-7. In turn, this means that for these blade sections, less of the
TR is resolved in the forward direction as thrust.
propeller
torque TR
Fig. 12-7.
propeller
torque
TR

thrust
thrust

relative
airflow direction
of rotation of rotation

inner section outer section


large blade angle smaller blade angle

Principles of Flight Propellers 12-5


Again, as shown in Fig. 12-7, with their smaller blade angle, the TR of the outer
sections is tilted further forward resulting in a greater proportion of the TR being
resolved as thrust. However, toward the tips, the propeller blades become less
effective, because:

• Vortices are formed at the propeller tips for the same reason as wing-tip
vortices are formed. These increase the aerodynamic 'drag' and reduce the
'lift', resulting in a tilting back of the TR, with a consequent reduction in
thrust and an increase in torque in the region of the tips.

• The propeller tip is the fastest travelling part of the aircraft since it has the
highest rotational velocity compounded with the aircraft forward speed. At
tip speeds approaching the local speed of sound, there is a significant
increase in aerodynamic drag which, again, tilts the TR rearward and acts to
increase the torque/reduce the thrust produced in the area of the tip.

Propeller Performance
Propellers are unable to convert all of the power developed by the engine into
useful thrust. There are losses involved in imparting motion (both translational
and rotary) to the slipstream, and in the aerodynamic drag on the propeller.
Under ideal conditions, a propeller will convert not more than about 90o/o of the
brake horsepower delivered at the propeller shaft to useful thrust. In normal
operating conditions, the efficiency of the propeller will typically vary between 50
and 85o/o.

Propeller efficiency is defined as the ratio of thrust horsepower (what the


propeller produces) to brake horsepower (what is delivered to the propeller by
the engine):
thrust horsepower
propeller efficiency= %
brake horsepower

Power is force x distance over time, hence thrust (horse)power in the above
equation can be expressed as thrust x TAS, and propeller efficiency can also be
written as:
thrust x TAS
propeller efficiency = ~~~­
brake power

(where brake power is the output of the engine, measured at the propeller shaft)

From the second of the above equations, note that propeller efficiency is zero
under two conditions-(!) when aircraft has no forward speed (zero TAS) and
(2) when there is no thrust produced. In practice, of course, the normal range of
operation of a propeller falls within these two extremes. Fig. 12-8 shows typical
curves of efficiency versus TAS for two different fixed-pitch propellers-one of
fine pitch (small blade angles) and the other of coarse pitch (large blade angles).

100%

coarse pitch
efficiency

I for constant rpm I

Fig. 12-8. Propeller efficiency curves.


aircraft speed (TAS)

12-6 Propellers The Commercial Pilot Series


Notes:
• For both propellers, efficiency is zero when the aircraft has no forward
speed as has been previously explained. Although a high value of thrust can
be developed when the aircraft is stationary, no useful work is done and the
propeller is, by definition, inefficient until the aircraft begins to move and
gains speed as a result.
• Given that each of the curves is for a constant rpm, the peak of efficiency for
a fixed-pitch propeller is reached at one airspeed. This is a low airspeed in
the case of the fine-pitch propeller and a high airspeed in the case of the
coarse-pitch propeller.
• Once maximum efficiency is attained, any further increase in airspeed
results in a relatively rapid decline in efficiency. This is because the angle of
attack of the blades is being reduced to below the optimum, with a
consequent reduction in thrust. It is possible, but only in a dive, to reach a
speed where the blade angle of attack is reduced to the zero-thrust angle
(which occurs just before the 'zero-lift' angle for the propeller aerofoil
section). At this speed, no thrust is produced and efficiency is again
reduced to zero.

Slip
The distance which the propeller would theoretically move forward (or
advance) in one revolution when it is giving no thrust is called the experimental
mean pitch. The difference between this distance and the actual distance
moved forward, is called slip. This difference is usually expressed as a
percentage of the experimental mean pitch.

A propeller can be imagined as a type of screw-in fact, it was once commonly


called an 'airscrew'. If an ordinary screw is turned through one revolution in a
solid medium, it advances the same distance as its pitch (or distance between
between the threads). When it is producing thrust however, a propeller cannot
advance the same distance as its 'pitch'. As air is not a solid medium, when a
propeller is turned in it, some of the air 'slips' backwards and the distance that it
advances is less than its mean pitch. The amount of slip present is an indication
of propeller efficiency.

Slip is shown diagrammatically in Fig. 12-9. Experiments have shown that


maximum propeller efficiency is obtained when the slip is about 30%.

relative airflow
at experimental mean pitch
distance (zero-thrust angle of attack)
moved in
plane of
rotation

Fig. 12-9. Propeller slip.

advance at experimental :' SLIP


mean pitch

Principles of Flight Propellers 12-7


Geometric Pitch
direction of TR at zero thrust angle
Geometric Pitch is the distance
of advance at zero degrees
angle of attack. It is a fixed
quantity determined by the
construction of the propeller
and not by its performance.
Note the difference between / relative airflow
d at experimental mean pitch
geometric pitch and distance
(zero-thrust angle of attack)
experimental mean pitch as moved in
plane of
shown in Fig. 12-1 0. Geometric rotation
pitch is measured from the
chord line and reflects the blade
angle of a given section.
Experimental mean pitch will
geometric pitch
occur when the blade sections
have a small negative angle of
experimental mean pitch
attack-where the TR becomes
aligned with the propeller
torque and, therefore, no thrust Fig.12-10. Geometric and experimental pitch.
is produced.

Constant Speed Propellers


A fixed-pitch propeller operates at optimum efficiency only under one set of rpm
and airspeed conditions. Because of this limitation, an early development in
propeller technology was the two-pitch propeller-with a fine pitch setting for
take-off and low speed operation, and a coarse pitch setting for higher airspeeds.
This propeller was able to operate within two ranges of efficiency similar to
those shown by the curves in Fig. 12-8, but within each range, it operated
basically as a fixed-pitch propeller.

Subsequently, the modern-day constant-speed propeller was developed. This


type of propeller has a blade angle which automatically adjusts to any position
between two in-flight limits at the fine and coarse ends of its range. In normal
operation, the pitch of the blades is determined by a mechanism called the
constant speed unit (CSU) which governs the propeller speed at an rpm set by
the pilot. By varying the pitch of the blades in this way, the constant-speed
propeller can be operated at optimum efficiency over a much wider range of
speeds than a fixed-pitch propeller.

100% peak efficiency 'envelope'

fine coarse
pitch A A pitch
limit limit
efficiency

for constant rpm

aircraft speed (TAS)

Fig. 12-11. Between its fine and coarse pitch limits, a constant-speed propeller can be
operated at peak efficiency over a wide range of airspeeds.

12-8 Propellers The Commercial Pilot Series


Whereas with a fixed-pitch propeller, the pilot has only the throttle to control
both engine power and propeller rpm, with a constant-speed propeller there are
two controls:
• The propeller control (usually called the pitch control) which controls
propeller rpm.
• The throttle to control manifold pressure (MP) and, therefore, the power
developed by the engine.

Single engine. Throttle, propeller (pitch) and


mixture. Twin engine. Quadrant controls.

Fig. 12-12. Typical control arrangements.

Moving the pitch control alters the rpm at which the propeller will be governed.
Moving the throttle alters the amount of power which will be delivered to the
propeller at the selected rpm. Controlling the power delivered by the engine and
the thrust from the propeller is thus a matter of selecting certain combinations of
propeller rpm and manifold pressure.

The Constant Speed Unit


The constant speed unit (CSU)-sometimes called a propeller control unit (PCU)
-contains a governor which controls the rpm to that selected by the pilot. It
does this by adjusting the pitch angle of the blades so that propeller torque
remains matched (equal and opposite) to engine torque regardless of changes
to airspeed and/or throttle setting.

Changes in Throttle Setting

If, at a given rpm setting, the throttle is opened (MP increased), engine torque
will increase. As a result, the rpm will want to increase, but as soon as the CSU
senses this it will increase the blade angle (coarsen the pitch) so that propeller
torque is increased to match the increase in engine torque. The selected rpm
will be maintained because the increase in engine power has been absorbed by
the increased angle of attack of the blades. At this higher angle of attack, the
propeller will produce more thrust and there will be a gain in performance-for
example, in level flight-an increase in speed or, in a climb-an increased rate
of climb.

Similarly, if the throttle is closed, the CSU will decrease the blade angle (fine the
pitch) so that the selected rpm is maintained. With the decreased angle of
attack of the blades, the propeller will produce less thrust.

Principles of Flight Propellers 12-9


Changes in Airspeed

If the airspeed is reduced by placing the aircraft in a climb without adjusting the
throttle, propeller torque will increase (because the propeller blades have a
higher angle of attack at the lower airspeed). As a result, the rpm will want to
drop, but as soon as the CSU senses this, it will fine off the pitch to keep the
propeller torque matched to engine torque thereby maintaining the selected
rpm.

Similarly, if airspeed is increased by placing the aircraft in a dive without


adjusting the throttle, the CSU automatically coarsens the pitch so that the
selected rpm will be maintained.
propeller
torque

Fig. 12-13. When changes to power and/or airspeed occur,


the direction of the relative airflow is affected.
The CSU responds by changing the propeller blade angle
so that propeller torque remains matched with engine torque
at the rpm selected by the pilot.

The CSU can only govern propeller rpm in the manner just described, within
certain limits. If engine power is reduced to a low setting, the blade angle will
fine off until it reaches what is known as the fine-pitch stop. Thereafter, the
propeller will act as a fixed-pitch propeller, with the rpm being controlled at low
power settings by the throttle. If the throttle is opened and the rpm rise to the
figure previously selected with the pitch control, the CSU will once again take
over and begin to govern the rpm at that figure.

The design of constant-speed propellers is such that, provided they are operated
within the range of in-flight rpm/MP settings recommended in the Pilot's
Operating Handbook, they will remain working at or close to their best angle of
attack and thus with optimum efficiency for the conditions.

Operation of Constant Speed Propellers


Constant-speed propellers are normally operated on the ground with the pitch
control in the FULL FINE position. As a part of the pre-flight checks, the CSU is
normally 'exercised' before each flight on cold days, otherwise before the first
flight of the day. This is achieved by running the engine up to the rpm specified
by the manufacturer (normally about 1800 - 2000 rpm) and then cycling the pitch
control slowly from FULL FINE to COARSE and back again two or three times.
The rpm should decrease (normally not more than 500) and increase again in
concert with the movement of the pitch control. The object of exercising the
CSU in this manner is to enable warm oil to circulate through the hydraulic pitch
changing mechanism. If this is not done, cold and viscous oil may result in
sluggish operation of the unit and the CSU may not be able to control rpm
properly on the take-off and climb-out.

12-10 Propellers The Commercial Pilot Series


As part of the pre take-off checks, ensure that the pitch control is placed in
the FULL FINE position for take-off. If this is not done and an attempt is made
to take off with the pitch control in a coarse position, the required rpm and full
thrust will not be developed during the take-off run. At best, the take-off
distance will be significantly increased and there will be a risk of damaging the
engine through 'overboosting' and detonation. At worst, the aircraft may simply
not become airborne in the distance available!

At all times, operate the engine within the range of rprn/MP settings
recommended in the Pilot's Operating Handbook. Operation at a higher MP
than that recommended for the rpm selected, can lead to high cylinder head
temperatures and detonation.

Changing Power Settings

To increase power (if it is desired to operate at higher rpm):


• first increase rpm with the pitch control;
• then increase MP to the desired value with the throttle.

To decrease power (if it is desired to operate at lower rpm):


• first reduce MP with the throttle;
• then reduce rpm with the pitch control.

As altitude is changed with an unsupercharged engine, so will MP at the rate of


about I" Hg per I 000 ft. The selected rpm should remain constant, but the
throttle will have to be occasionally adjusted to maintain a given MP (opened in
a climb, closed in a descent).

During the pre-landing checks, ensure the pitch control is set to FULL FINE
so that full power will be available in the event of a go-around.

Other Modes of Operation


In addition to providing more efficient operation over a wide range of airspeeds
and engine power, with some constant-speed propellers the advantages of being
able to va1y the pitch are extended to:
• The ability to be feathered in flight to avoid 'windmilling' drag and the
possibility of further damaging an engine which has failed.
• The ability to be placed in ground-fine pitch or reverse pitch, to provide
aerodynamic braking dUling the landing run.

Wind milling

If there is a loss of engine torque to the propeller, the CSU will fine off the pitch
in an attempt to maintain the rpm selected at the time. A point is soon reached
where the relative airflow approaches the blade at a 'negative' angle of attack
which is large enough to produce a TR in the reverse direction to normal-refer
to Fig. 12-14. The torque component of this reversed TR now acts in the same
direction as engine torque would normally act and will drive (i.e. 'windmill') the
engine, even though no power is being produced. The reversed thrust
component (windmilling drag) also now acts in the same direction as aircraft
drag. This extra drag from a windmilling engine can be substantial, and is
roughly equivalent to a flat disc with the same diameter as the propeller being
placed at right angles to the relative airflow.

Principles of Flight Propellers 12-11


windmilling
drag
~ - - - ,.,,,...~:;;.;i\
'

TR

relative
airflow
Fig. 12-14. The forces acting I
on a windmilling propeller. I
I
I
I
rotation I
I

TAS

Feathering

It is clearly a big advantage to be able to prevent windmilling drag following an


engine failure. The extra drag from this cause can limit the 'engine-out' range,
degrade performance, and create control difficulties on a multi-engined aircraft.
Additionally, with windmilling torque continuing to drive a possibly damaged
engine, there is the eventual risk of seizure or fire.

Feathering involves turning the blades to any torque or 'lift' produced


the angle of attack with the oncoming by inner sections:
airflow at which no net propeller torque is
produced. In this position, the blade
airflow
sections nearer the hub will have a
'positive' angle of attack, while those <:=
nearer the tips will have a 'negative'
angle-'positive' and 'negative' lift will
cancel out and no turning moment on the is balanced by that produced
by outer sections
propeller will be generated (Fig, 12-15).
When the propeller is in the feathered Fig. 12-15. A feathered propeller.
position, it will produce the minimum
amount of drag.

NOTE: For those constant-speed propellers which do not have a feathering


capability, it is important that the pitch control is placed in the FULL COARSE
position following an engine failure. This will reduce the drag from the 'dead
engine' to the minimum, even if for some reason, it does not continue to turn
over.

Reverse Thrust

If the propeller blades are turned through the fine pitch stop to a blade angle of
about minus 20' and power is applied, reverse thrust is obtained. The blade
sections are working relatively inefficiently at a negative angle of attack. The
forces acting on the blade when operating in reverse pitch are shown in Fig. 12-
16.

As the blade is being rotated below the normal fine pitch stop to obtain the
reverse thrust angle, it must pass through an arc in which the blade angles are
small and negative. During the transition through this arc, the forces acting on the
propeller will be as shown in Fig. 12-14, i.e., it will be in the windmilling mode. If
power is applied when the propeller angle is in this arc, an overspeed is likely.
Mechanical devices are therefore used to prevent the application of power until
the propeller blades have been turned safely into the reverse pitch range.

12-12 Propellers The Commercial Pilot Series


TR

,._ - - - '-!it'C';~'f!l
reverse
thrust

Fig. 12-16. Reverse thrust. relative


airflow

TAS

Propeller Twisting Moments


Considerable stresses are placed on the propeller blades and pitch changing
mechanism in flight. The most important of these are:
• centrifugal force;
• centrifugal twisting moment; and
• aerodynamic twisting moment.

Centrifugal Force

The blades of a rotating propeller have mass and are turning on a circular path
about the propeller hub. To keep the blades turning on their circular path, a
centripetal force must be applied toward the centre of rotation. This force,
proportional to the mass of each blade and the square of its rotational velocity, is
of course, applied to the blade at the attachment point of the blade root with the
csu.
The centrifugal force on the blade is the reaction (in accordance with Newton's
Third Law) to the centripetal force required to keep it turning in its circular path.
As much as CSU attachment point has to 'pull' on the blade to keep it rotating in
a circle, the blade reacts by 'pulling' in the opposite direction with the same
force. This centrifugal force (which is acting as if to stretch the blades and pull
them out of their 'sockets') can be very strong, especially at high rpm.

Centrifugal Twisting Moment (CTM)

Consider the propeller blade in Fig. 12-17. When the propeller is rotating, the
centrifugal force which affects each part of the blade originates from the centre
of rotation. For example, that part of the blade at X on the leading edge of the
blade is subject to the force X-A, and that part at pointY is subject to force Y-8.

Each of these forces can be resolved into the following components:

• Forces in line with the span of the blade (parallel to the pitch-change axis)
which try to pull the blade out of its 'socket'. These spanwise components
on the leading and trailing edges of the blade are shown as X-C and Y-D.

• Forces at right angles to the span (and pitch-change axis) which try to pull
the blade into fine pitch. These chordwise components result in the
centrifugal tWisting moment (CTM). To visualise the action of the CTM,
consider the end view of the blade section as well as the plan view shown
in Fig. 12-1 7. It can be seen that the forces on the leading and trailing edges,

Principles of Flight Propellers 12-13


represented by X-E and Y-F B D C A
respectively, form a couple about +-:, -t. centrifugal
the pitch-change axis which
attempts to pull the blade into
'\ 1:I' twisting
:' I' moment
fine pitch.
''' ' I :
I '
(CTM)

' F -~X E)
I
In constant-speed propellers, The
' '

~
CTM places a greater demand on the F' E
pitch changing mechanism when
increasing blade angle (moving to
coarse) and less demand when PltChwchange
axis
decreasing the angle (moving to
fine). The strength of the CTM is
influenced by the shape of the
propeller-those with wide blades
and a greater chordwise distribution
of mass, will have a stronger CTM
and a greater tendency to go into fine Fig. 12-17.
pitch, than propellers with narrow Centrifugal twisting moment.
blades.

Aerodynamic Twisting Moment (ATM)

An aerodynamic twisting moment (ATM) arises whenever the aerodynamic


total reaction force (TR) does not act on a line through the pitch-change axis. In
normal operation, the centre of pressure of the blade is usually forward of the
pitch change axis resulting in an ATM which tends to coarsen the blade angle
(i.e. turn it in the opposite direction to the CTM). But, as the ATM is usually
much weaker, it only partially offsets the CTM.

When a propeller is windrnilling however, the ATM is reversed and acts in the
same direction as the CTM-i.e. both act to fine off the propeller. In a steep dive
at low power, the combined effect of the CTM and ATM may be strong enough to
prevent the CSU from moving the blades to a coarse position as speed is gained,
resulting in an engine overspeed.

ATM ATM CTW

relative relative
airflow airflow

(A) (B)

Fig. 12-18. (A) Aerodynamic twisting moment, and (B) twisting moments on a windmilling
propeller.

12-14 Propellers The Commercial Pilot Series


Asymmetric Blade Effect
When the propeller disc is operating at right angles to the direction of flight, the
relative airflow meets the downgoing blade sections at the same angle as their
counterparts on the upgoing blade. However, when the disc becomes tilted
back with respect to the direction of flight, the relative airflow approaching the
downgoing blade is different to that affecting the upgoing blade sections.

The top illustrations in Fig. 12-19 represent a propeller operating at right angles to
the direction of flight (a) and then tilted back (b). In (a) it can be seen that the
distance travelled and the angle of attack of each blade is the same. When the
disc becomes tilted back however, as in (b), the distance travelled by the
downgoing blade and its angle of attack both appear to be greater than for the
upgoing blade.

This can be confirmed if vector diagrams are drawn for each of the propeller
blades, as shown in the lower part of Fig. 12-19. You will recall that the relative
airflow for a given blade section can be determined by finding the resultant of its
rotational velocity vector (rpm) and forward velocity vector (TAS). The vector
diagrams illustrated are for an equivalent blade section of a propeller at constant
rpm and aircraft TAS. Note, therefore, that the length of the vectors representing
the distance travelled in a half revolution are the same, as is the length of the
TAS vectors, for all four diagrams.
(a) (b)
.... 0 /...- D /
~ II ~ I
\ I ' I
\ I ' I
I I ' I
direction of travel 'I
I \
/ r
''
I \ I
~ ',"
/
/ u
\,
' ... /
fu -.. '

- DOWNGOING BLADES -

rotation RAF RAF

TAS TAS

rotation RAF rotation RAF

UPGOING BLADES

Fig. 12-19. Asymmetric blade effect.

Principles of Flight Propellers 12-15


When the propeller is at right angles to the direction of flight, the vector
diagrams (on the left) clearly show that the relative airflow and angle of attack of
the upgoing and downgoing blades are equivalent. When the disc is tilted back,
it is also clear that the upgoing and downgoing blades adopt different angles
between their rotational and TAS vectors. In effect, during its 180° of rotation,
the downgoing blade advances by the amount of tilt in addition to moving
forward with the aircraft. In contrast, the upgoing blade retreats by the amount
of tilt at the same time as it is carried forward by the aircraft.

As a result of the tilt, it can be clearly seen that the relative airflow vector for the
downgoing blade is longer, and thus the airflow past this blade is faster than for
the upgoing blade. In addition, although the blade angle (angle between chord
and plane of rotation) is the same for both blades, the angle of attack of the
downgoing blade is higher. For these two reasons-faster airflow and higher
angle of attack-the downgoing blade develops more thrust when the propeller
disc is tilted back.

Asymmetric blade effect results in a shift in the thrust line toward the side of the
propeller disc with the downgoing blades. For propellers which rotate
clockwise (when viewed from the cockpit) the shift is to the right. Asymmetric
blade effect becomes significant only when the 'tilt back' of the propeller disc is
comparatively large. It is one of the factors contributing to swing on take-off in a
tail-wheel aircraft (as explained in Chapter 13) and is a determinant of 'critical
engine' in multi engine aircraft (Chapter 14).

Propeller Solidity
For efficient operation, propellers must be carefully matched to the power of the
engine so that the power which the engine is capable of producing can be
'absorbed' by the propeller and converted efficiently into thrust. The critical
factor in matching a propeller to the power of the engine is the tip velocity.
When the propeller tips approach the speed of sound, compressibility effects
significantly decrease the thrust and increase the propeller torque, reducing the
efficiency of the blades. These effects impose a limit on the propeller diameter,
and the rpm and TAS with which it can be used.

The usual method of absorbing the power of an


chord
engine is to adjust the solidity of the propeller- c
i.e. the ratio between that part of the propeller disc
which is solid at a given radius to the ' '
' '
circumference of the disc at the same radius (Fig. ' '
,' '
12-20). Solidity is measured by: radius
No. of blades x chord at radius r
circunlference at radius r
'
''
It can be seen from the diagram that solidity can '
be increased by: ''
• increasing the number of blades (e.g. to 3 or
4); or
• increasing the chord of the blades.
Fig. 12-20. Propeller solidity.
Although the second of the above options is easier
to engineer, increasing the chord decreases the
aspect ratio of the blades, making them less
efficient.

12-16 Propellers The Commercial Pilot Series


Review 12

I. A propeller converts engine torque into (thrusVhorsepower).

2. To ensure that it operates at an efficient angle of attack over its whole


length, a propeller blade is (twisted/straight).

3. The blade angle is that angle between the chord line of a propeller section
and the (plane of rotation/direction of flight).

4. The blade angle is greatest near the (hub/tip).

5. The fastest moving part of a rotating propeller is at the (hub/tip).

6. The total aerodynamic force produced by a propeller is normally resolved


into two components; one in the direction of flight called ....................... , and
the other in the plane of rotation called propeller ........................... .

7. At constant rpm, engine torque is balanced by ............................ torque.

8. If you place a fixed-pitch propeller aircraft into a dive and leave the throttle
set, engine rpm will (increase/decrease).

9. Most of the thrust produced by a propeller comes from the area of the
blades at around (50%/ 70%/ 90%) radius.

I 0. With a fixed-pitch propeller at a given rpm, there will be (one/several) true


airspeeds at which the propeller will operate at its most efficient angle of
attack.

II. For operation at high rpm and relatively low airspeeds, the most efficient
fixed-pitch propeller would have a (fine/coarse) pitch.

12. For operation at relatively high airspeeds and low rpm, the most efficient
fixed-pitch propeller would have a (fine/coarse) pitch.

13. Propeller efficiency can be measured by thrust X.......... ./brake power.

14. From Q 13, efficiency is therefore zero at ................... airspeed, and also
when the propeller is operated at the .............. - ..................... angle of attack.

15. The distance the propeller would theoretically advance in one revolution at
the zero-thrust angle of attack is called ................................................ pitch.

16. The difference between experimental mean pitch and the actual distance
the propeller advances in one revolution is called ................. .

17. Geometric pitch is the distance of advance at (zero degrees/the zero thrust)
angle of attack.

Constant-speed propellers

18. The blade angle of a constant-speed propeller can be varied in flight


between its fine and coarse pitch limits. The mechanism which governs the
angle which the blades take up is called the ......................... .................... .

Principles of Flight Propellers 12-17


19. The rpm at which the CSU will govern the propeller is determined by the
pilot using the ........................ control.

20. The power developed by the engine is controlled by the pilot varying
manifold pressure (MP) with the .................................. .

21. At a given rpm setting, if MP is increased, the CSU will act to (fine off/
coarsen) the pitch of the blades to absorb the increase in engine torque.

22. If, at constant throttle and pitch lever settings, airspeed is reduced, the CSU
will act to (fine off/coarsen) the blade angle to keep the propeller torque
balanced with engine torque (and therefore maintain constant rpm).

23. Aircraft fitted with constant-speed propellers are normally operated on the
ground with the pitch control set to (FULL FINE I COARSE).

24. In the take-off checks, ensure that the pitch control is set to ........................ .

25. When increasing power (when it is desired to operate at higher rpm), first
increase the ........................ ,then increase ........................ .

26. When decreasing power (when it is desired to reduce rpm), first


reduce ........ , then reduce ................ .

27. A windmilling propeller operates at a small (negative/positive) angle of


attack, and the propeller torque acts in the (same/opposite) direction as the
engine torque would normally act.

28. With a feathered propeller, the blade angle is turned such that no ................ .
propeller torque is produced.

29. In the reverse thrust mode, the propeller blades are turned to a relatively
large (positive/negative) angle of attack and power is (increased/reduced).

30. The centrifugal twisting moment (CTM) tends to turn the blades into (fine/
coarse) pitch.

31. The aerodnamic twisting moment (ATM) normally tends to move the blades
into (coarse/fine) pitch. However, when the propeller is windmilling, the
ATM reverses and in a dive with low power, the combined effect of the CTM
and ATM can 'lock' the propeller into (fine/coarse) pitch and the CSU may
not be able to prevent an overspeed.

32. When the propeller disc becomes tilted back with respect to the direction of
flight, the dovvngoing blade travels (faster/slower) with a (higher/lower)
angle of attack and thus develops more ..................... than the upgoing blade.
This effect is called ........................................................ effect.

33. The normal method of increasing the ability of the propeller to absorb the
power developed by the engine, is to increase ........................... . State two
methods usually employed to increase solidity.

12-18 Propellers The Commercial Pilot Series


Stability
Introduction
13
When properly trimmed in a given flight attitude, a stable aircraft will return to
that attitude after having been disturbed from it, without intervention by the pilot.
An unstable aircraft will move further away from the original attitude after a
disturbance, making it difficult to control (this is also called negative stability). A
neutrally stable aircraft will remain in the disturbed position, unless the pilot
does something about it.

In this chapter, we look at the main factors affecting the stability of an aircraft in
the air. This subject is an extremely complex one, hence our coverage of it is
necessarily limited. Toward the end of the chapter we include an allied
subject-stability and control of an aircraft on the ground.

Static and Dynamic Stability


There are two kinds of stability-static and dynamic.

Static stability refers to the initial reaction of a body after being disturbed or
displaced from a position of equilibrium. If it initially tends to return to its
original position it has positive static stability (it is statically stable). If it remains
in its displaced position it has neutral static stability. If it continues to move away
from the original position, it has negative static stability (it is statically unstable).
These conditions are simply illustrated in Fig. 13-1, which can be taken to relate
to a ball bearing resting on three different smooth surfaces.


tends to move back to
original position
stays in disturbed position
moves further away
from original position
I,.-- \
'- I

positive stability neutral stability negative stability

,.--
original position { ) disturbed position
"
Fig. 13-1. Static stability.

An aircraft will have positive static stability in pitch if, for example, once the nose
attitude is disturbed up or down by a gust, it initially moves of its own accord
back toward the original position.

Dynamic stability relates to the subsequent motion of the disturbed body, once
the static stability reaction has taken place. An aircraft will have dynamic
stability if it eventually returns of its own accord to its original position of
equilibrium, after a disturbance. It will have neutral dynamic stability if it
continues to oscillate evenly about the original attitude, and it will be
dynamically unstable if it continues to diverge from the original attitude.

Principles of Flight Stability 13-1


original disturbance returns to original attitude of own accord
attitude

~~ positive
static positive dynamic stability
stability
!
"
,"'
,' oscillates evenly about original attitude
-----z '

neutral dynamic stability

diverges from ·
, original attitude 11~
P/ dynamically unstable
Fig. 13-2. Stability of an aircraft in the pitching plane. {negative dynamic stability)

Static and dynamic stability for an aircraft in the pitching plane are illustrated in
the above diagram. Although we have shown this stability in the pitching plane
only, the same considerations apply in roll and yaw and this will be covered
shortly. An aircraft which returns to its original trimmed attitude unassisted by
the pilot, as in the first example above, is said to be inherently stable. Most
aircraft are designed to have this in-built stability.

An aircraft must have static stability before it can be dynamically stable.


However, note how, in the third example above, an aircraft can be statically
stable, but be unstable dynamically. An aeroplane with this characteristic can be
dangerous, if not impossible to fly, hence designers make sure their aeroplanes
do not exhibit it. In the following paragraphs, we deal with the stability of an
aeroplane in the three planes of movement (in pitch, yaw and roll). We are
mainly concerned with static stability in these three planes-i.e. the initial
tendency of the aeroplane to return to its original attitude. We do not consider
in any great detail the way in which it returns to equilibrium (i.e. its dynamic
stability) as this is a complex subject and beyond the scope of this manual. We
can take it however, that the aeroplane will be designed to have positive
dynamic stability as shown above.

Stability and Controllability


Stability is the inherent ability of the aircraft to return to its original attitude after
being disturbed, without any action being taken by the pilot.

Controllability refers to the ease with which a pilot can manoeuvre the aircraft
and change its attitude by using the control surfaces.

In the design of most aircraft, there has to be a compromise between stability


and controllability. An aircraft which has a high degree of stability is resistant to
changes of attitude and manoeuvre thereby reducing its controllabilty. For a
pilot used perhaps to less stable aircraft, a 'high-stability' aircraft will feel
'sluggish' on the controls and require heavy stick forces to manoeuvre.

13-2 Stability The Commercial Pilot Series


On the other hand, an aircraft with little or poor stability will require constant
control inputs from the pilot to keep it from diverging from the selected attitude.
While the stick forces are light and the aircraft is readily manoeuvred, this type of
aircraft is difficult and tiring to fly.

Most aircraft are designed to have a sufficient degree of static stability so that,
when trimmed in the required attitude, they will virtually fly 'hands-off-i.e.
require guidance from the pilot rather than constant control inputs. At the same
time, this degree of stability should leave the aircraft with sufficient
manoeuvrability to suit its role.

Longitudinal Stability
Longitudinal stability is in the pitching plane about the lateral axis. To be
longitudinally stable, an aircraft must have an inherent tendency to return to the
same pitch attitude after a disturbance. Longitudinal stability is the most
important static stability mode because disturbances to the pitch attitude affect
the angle of attack of the wings, resulting in changes to the lift being generated.
The aircraft must be designed in such a way that the forces generated by these
uninvited changes in pitch angle/angle of attack act so as to return the aircraft to
the original pitch attitude.

In conventional aircraft, the tailplane (or horizontal stabilizer) is the principal


means of obtaining longitudinal stability. If, as often happens, turbulence (a
gust, swirl or eddy) causes the pitch attitude to change, it is the change in angle
of attack of the tail plane which produces the main restoring (or stabilizing)
force.

The action of the tailplane in maintaining stability is shown in Fig. 13-3. In this
example, the aircraft has been disturbed nose up from its trimmed attitude. As
the disturbance is taking place, the aircraft, because of its momentum, will
initially tend to continue on its original flight path and the angle of attack of both
the wings and the tailplane will be increased. The resulting increase in tailplane
lift (or decrease in negative lift) produces a nose down pitching moment which
returns the aircraft to its original trimmed position.

~--0
constant pitch attitude nose-up disturbance: tail plane lift increased restoring moment
with tailplane providing angle of attack of wings to provide nose down disappears when
(say) a small amount of and tailplane increased restoring moment original attitude regained
lift to balance main forces
Fig. 13-3. Longitudinal stability following an uninvited nose-up pitch.

Should an uninvited nose-down pitch occur, the tailplane operates in the reverse
sense to that depicted in Fig. 13-3 to provide the restoring moment and return
the aircraft to the trimmed attitude and equilibrium.

Wing Pitching Moment

Consider the distribution of the lifting forces and weight shown in Fig. 13-4. In
diagram (a), the centre of pressure (CP) of the wing is positioned behind the
centre of gravity (CG). If the aircraft is displaced nose-up, wing lift increases
which tends to increase the nose-down pitch about the CG, i.e. the wing pitching
moment from lift tends toward stability. However, for a cambered wing section,
the CP also moves forward with increased angle of attack, and this tends to

Principles of Flight Stability 13-3


reduce the stabilizing effect of the increase of lift. In some circumstances, the
forward movement of the CP can be large enough to cause an unstable pitching
moment.

increased probable
wing lift nose-down
increased tailplane lift; (stable)
nose~down pitch pitch

(a) CP behind CG nose-up


displacement

increased tailplane lift; increased


wing lift
t nose-up
(unstable)
nose-down pitch
pitch

(b) CP ahead of CG
nose-up
displacement

w
Fig. 13-4. Wing pitching moment following a nose-up disturbance.

In diagram (b), the CP is shown as acting ahead of the CG (although not the
usual arrangement of forces, it does occur with some aircraft). In this case, the
wing pitching moment following a disturbance is clearly unstable. That is, if the
nose is displaced upward, for example, the increase in the wing pitching
moment (through the increase in wing lift and movement of the CP) tends to
increase the displacement in pitch.

There will also be a similar (but smaller) pitching moment for the fuselage,
which generally tends to be unstable, and which the designer must take into
account.

The tailplane must be designed to counter any unstable influence of the wing
and fuselage pitching moments. For positive longitudinal stability, the tailplane
restoring moment must remain greater than any unstable wing (plus fuselage)
pitching moment. Disregarding the small effect of the fuselage pitching
moment, for this to occur in diagram (b):

tail plane lift x y must be greater than wing lift x z.

13-4 Stability The Commercial Pilot Series


factors Affecting the Degree of Longitudinal Stability
The main factors affecting the degree of longitudinal stability are the design of
the tailpliu1e, and the position of the centre of gravity.

Tailplane Design

The restoring moment and the effectiveness of the tailplane in maintaining


longitudinal stability are in large part determined by its design. Features which
affect the restoring moment include; the distance between the tailplane CP to
the aircraft CG (tailplane moment arm), tailplane area and shape, and the effect
of downwash from the wing.

Longitudinal Dihedral

In many aircraft, the tailplane is set at a lower angle angle of incidence than the
wings. For example, as shown in Fig. 13-5, the wing may be set at an angle of 3'
incidence with respect to the fuselage datum line whereas the tailplane may be
set at a lower angle of (say) 1'. This difference is called longitudinal dihedral.

tail plane wing


fuselage
datum

Fig. 13-5. Longitudinal dihedral.

The main reason for longitudinal dihedral is to enable the tailplane to operate
efficiently in the cruise. With aircraft of conventional configuration, the tailplane
is influenced by the downwash behind the wing. To trim out the pitching
moments in normal cruise, most aircraft require the tailplane to produce a small
downward force (nose-up pitching moment). To provide for this, after taking
into account the downwash from the wing, the tailplane must usually be set at a
slightly lower angle of incidence than the wings. (Note the difference between
angle of incidence and angle of attack).

Longitudinal dihedral has no effect on the basic stability of the aircraft. As the
coefficient of lift cmves in the normal operating range for both the wing and
tailplane may be considered as straight lines, the variation in lift per degree
change in angle of attack does not depend on the initial incidence settings or on
their difference. If, for example, the aircraft represented in Fig. 13-5 became
displaced 2' nose-up, increasing the angle of attack of the wing from 3' to 5', and
the angle of attack of the tail plane from 1' to 3', the lift from the tail plane will
increase in the same proportion as the lift from the wings. The main criteria for
stability is that the tailplane restoring moment remains greater than any unstable
moment from the wings.

Position of the CG

The pilot, of course, has no control over the design features which affect
longitudinal stability. However, longitudinal stability-and controllability in pitch
-is determined to a large extent by the position of the CG, and this is something
which the pilot can influence through control of the loading of the aircraft.

Principles of Flight Stability 13-5


Forward CG Position

The further forward the CG, the longer the moment arm becomes for the
tailplane. A given tailplane force therefore creates a more effective restoring
moment and the aircraft becomes more stable in pitch.

tail plane I FORWARD CG I tail plane I AFTCG I


lift lift

long arm shorter arm


high restoring moment- stable lower restoring moment -less stable

Fig. 13-6. A forward CG position -greater longitudinal stability.

At the same time however, the further forward the CG, the less effective the
elevator becomes in changing the pitch attitude. This is because any given
elevator deflection must now overcome a greater tailplane restoring moment-
i.e. stability opposes manoeuvre. If flown with the CG further fmward than
normal, the aircraft feels 'nose-heavy' and higher stick forces are required to
manoeuvre it in pitch.

Aft CG Position

The further aft the CG, the shorter the tailplane moment arm and the less stable
the aircraft becomes in pitch. At the same time, the elevator becomes more
effective and a smaller elevator deflection/lower stick force is required to
achieve the same change of nose attitude in pitch. When flown with an aft CG
position, the aircraft feels 'tail heavy' and becomes more sensitive to elevator
movement.

CG Limits

Limits are laid down for the range within which the CG must lie for safe flight. It
is the responsibility of the pilot to ensure that the aircraft is loaded so that the CG
position falls within these limits.

If the aircraft is flown with its CG behind the aft limit, it will have insufficient
longitudinal stability and will be difficult to control. It may even be impossible to
regain control if the aircraft stalls and enters a spin.

If flown with the CG forward of the fmward limit, the aircraft will Jack
manoeuvrability; elevator stick forces may become excessive and, with the lack
of elevator effectiveness at slow speeds, it may not be possible to flare the
aircraft properly for landing.

Always ensure the aircraft is loaded within its CG limits.

Directional Stability
Directional stability is in the yawing plane about the normal axis. It may be
considered as the inherent ability of an aircraft to 'weathercock' so that the nose
remains pointed into the oncoming airflow.

13-6 Stability The Commercial Pilot Series


Directional stability relies on the aircraft having a greater amount of 'keel' (or
side) surface behind the CG than ahead of it. In a conventional aircraft the
major part of this keel surface, and therefore of directional stability, is provided
by the fin (or vertical stabilizer), but it is worth noting that all keel surfaces, and
particularly those of the fuselage, play a part.

The fin is a symmetrical aerofoil which, when the aircraft is aligned with the
oncoming airflow, has a zero angle of attack and accordingly produces no net
side force. If the aircraft is disturbed in yaw by turbulence, or for some other
reason, its momentum will cause a skid to develop. The fin now develops an
angle of attack and a restoring moment which returns the fuselage back to
alignment with the airflow.

fin side
force

restoring moment

Fig. 13-7. Directional stability following an uninvited yaw.

The factors affecting directional stability are similar to those mentioned for
longitudinal stability. The greater the fin area and keel surface behind the CG,
and the greater the moment arm, the greater the directional stability of the
aircraft. Thus a forward CG favours directional stability more than an aft CG as it
gives a longer moment arm for the vertical stabilizer.

Design measures for improving directional stability include the fitting of dorsal
and ventral fins (to increase the fin area), and sweeping back the fin (which
moves its CP rearward and increases the moment arm). A swept-back fin also
stalls at a greater angle of attack, allowing greater side-slip angles to be attained
before the control surface stalls.

Lateral Stability
Lateral stability is the inherent ability of an aircraft to recover from a disturbance
in the lateral plane (i.e. in roll about the longitudinal axis) without any control
input by the pilot. L

A disturbance in roil will cause one wing to drop and the I


I
other to rise. As we have seen previously, when the I
aircraft is banked, the lift vector is inclined-and the I
resultant of weight and the inclined lift vector generate a I
force which causes the aircraft to slip sideways toward the I
lower wing. I~
With side-slip, the relative airflow approaching the aircraft / '.:::::::
/ lateral
has a lateral, or sideways component. This lateral component
component of the relative airflow is then used in a various 1 ofRAF
ways to produce a rolling moment to restore the aircraft to w
its original wings-level position. This may sometimes be
Fig. 13-8 A roll disturbance
referred to as keel effect. causes a sideslip,

Principles of Flight Stability 13-7


The use of. dibedral.in the construction. of the.wings (and tailplane) is the usual
means .of pmviding lateral stability, but a number of other factors also make a
contribution. These are discussed in the following paragraphs.

Factors Affecting Lateral Stability


Dihedral

As shown in Fig. 13-9, Dihedral is the upward inclination of the wingtips with
respect to the wingroots.

Fig. 13-9. Dihedral.

_ _Qi!J§\!.r_aj_ ___ _

As the aircraft develops side-slip after an


uninvited roll, the lower wing, because of
L
dihedral, meets the relative airflow at a
higher angle of attack than the upper
wing. The lower wing therefore produces
more lift than the upper and a rolling
moment is produced which tends to
return the aircraft to its original wings-
level position with no side-slip.
sideslip component
A negative dihedral, known as anhedral, of relative airflow
has a destabilizing effect, i.e., it tends to
accentuate any uninvited roll. Anhedral is
used typically on large high-wing aircraft Fig. 13-10. Dihedral corrects
to moderate and prevent the natural uninvited rolling
laterai stability of this configuration from
becoming too strong (see below).
Shielding

Once the aircraft begins to side-slip, the trailing (upgoing) wing becomes
shielded to some extent by the fuselage, which contributes to dihedral effect.

Wing Position

A high-wing design confers greater lateral stability than a low-wing design. As


shown in Fig. 13-11, when uninvited roll occurs, the centre of pressure (CP) of
the wing becomes laterally displaced with respect to the centre of gravity (CG).
In addition, with side-slip and shielding of the upper wing, the CP tends to move
toward the lower wing, increasing this displacement. The couple which is set
up laterally between the vertical component of lift and the weight tends to
restore the aircraft to the wings-level attitude. Low-wing aircraft also develop
the same lift/weight restoring couple after an uninvited roll, but as the CP and CG
are much closer vertically, the restoring moment is much smaller.

13-8 Stability The Commercial Pilot Series


Fig. 13-11.
The stabilizing effect of a
high wing/low centre of gravity.

w
Keel Surface/Fin

Where the geometric centre of the keel surface/fin area is above the CG, the
side-slip 'drag' force will tend to roll the aircraft away from the direction of the
slip-i.e. provide a stabilizing influence. Tall fin designs, particularly when
combined with a high T-tail and a low CG, make a positive contribution to lateral
stability.

keel surface~ rolling


'drag' ' \~ffect

Fig. 13-12. High keel surfaces and a low CG


provide a lateral restoring moment.

component
ofRAF
Sweepback

Lateral stability is increased if the wings have sweepback. As the side-slip


develops after an uninvited roll, the lower swept-back wing presents more of its
span to the oncoming airflow, and the length of the effective chord is decreased
(thickness/chord ratio effectively increased). As shown in Fig. 13-13, these
factors combine to give an effective increase in the aspect ratio of the lower
wing which, as a result, generates more lift and provides a restoring moment to
any uninvited roll.

effective
span relative airflow

effective relative airflow


span

______L__ _
LOWER WING - MORE LIFT

Fig. 13-13.
Sweepback counters univited roll.

Principles of Flight Stability 13-9


Lateral and Directional Stability Considered Together
For lateral stability, it is essential to have the side-slip which the disturbance in
roll causes. This side-slip provides the sideways component of relative airflow
which is necessary for the dihedral and other lateral stability features to work
and provide a restoring moment in roll.

However, whenever slip (or skid) is introduced, the aircraft's directional stability
is also brought into play. Hence, when the aircraft side-slips after a disturbance
in roll, its directional stability will cause it to yaw in the direction of the slip and,
through the further effect of yaw, make it want to continue to roll in the direction
of the disturbance.

Therefore, after a disturbance in roll, the lateral and directional stability


characteristics of an aircraft are in conflict in that:
• lateral stability (dihedral effect) wants to return the aircraft to wings-level;
but
• the directional stability (weathercock effect) wants to make the aircraft roll
further.
The designer has to ensure that lateral and directional stability are correctly
matched-that neither predominates too much.

Spiral Instability

If an aircraft has strong directional, but weak lateral stability-e.g. has a large
vertical stabilizer but no dihedral-once disturbed in roll it will tend to continue
to roll in the same direction. The increased angle of bank leads to more side-
slip, which leads to more yaw and more roll, and the nose begins to drop, and so
on. This is called spiral instability which, if the pilot does nothing to intervene,
continues until the aircraft is in a steep spiral dive.

The balance between lateral and directional instability is such that in most
aircraft, there is a slight tendency toward spiral instability. That is, continued
application of a yawing force (rudder, or asymmetric power) leads to the aircraft
entering a spiral dive if the pilot does nothing about it. This degree of spiral
instability is unimportant, as it is easily controlled and is preferable to having too
much lateral stability, which we now consider.

Dutch Roll

When the lateral stability of an aircraft is too strong by comparison with its
directional stability, a form of dynamic (oscillat01y) instability called Dutch roll
can occur. Dutch roll is characterised by a combined rolling and yawing
motion-a wallowing motion-which can continue for some time after an
aircraft has been disturbed from a position of equilibrium.

A simplified explanation of what is a complicated aerodynamic process, is as


follows:
• The aircraft is disturbed in yaw. Because the restoring moment from the fin
and keel surfaces is not strong (i.e. weak directional stability) the recove1y
from the yaw disturbance is slow and poorly damped. The aircraft begins to
oscillate in yaw.
• The yawing motion generates a rolling motion through dihedral effect. An
oscillat01y motion is soon set up where the aircraft is both rolling and
yawing simultaneously.

13-10 Stability The Commercial Pilot Series


• The yawing and rolling oscillations are of the same frequency. However
because of the much stronger lateral stability, the recove1y from each
excursion in roll is faster than the recove1y in yaw. The roll and yaw
oscillations thus get out of phase and tend to 'feed' off one another. At
times therefore, the aircraft will be rolling left while it is still yawing right,
and vice versa. This is the wallowing motion which characterises the Dutch
roll.

In some cases, the yawing motion in Dutch roll is much more pronounced than
the oscillation in roll. In these cases the motion is often referred to as snaking.

Stability and Control on the Ground


If an aircraft is not to tip over on the ground, the CG must remain within the area
bounded by the three wheels. The further the CG is from any of the boundaries
of this area (indicated by the dashed lines in Fig. I 3- I 4) the less likely the aircraft
will be to tip over in that direction.

nose~whee! aircraft tail-wheel aircraft

Fig. 13-14. For stability, the CG must remain within the area bounded by the wheels.

A low CG and widely spaced wheels reduce the tendency for the aircraft to tip
over on the ground, e.g. in strong winds, when turning or, in the case of
tailwheel aircraft, when brakes are applied and/or high power is used.

High keel surfaces and dihedral make the aircraft more susceptible to the
destabilizing effects of a crosswind.

Ground Roll Stability


~,

With its CG placement ahead of the main ''


wheels, the nosewheel (or tricycle) ''
configuration is inherently stable-that is, a
nosewheel steering '- '
(or something) \
nosewheel aircraft tends to maintain a straight must apply a side -
track when travelling over the ground (during force to turn
taxiing, take-off or in the landing roll-out). If
the aircraft is displaced from its straight
course, the centrifugal reaction force, acting
through the CG and pivoting about the main
wheels, tends to straighten the aircraft up. To
turn a nosewheel aircraft on the ground, a
positive force must therefore be applied at all nosewheel aircraft
times through nosewheel steering, differential
braking, asymmetric power or a combination
thereof. Fig. 13-15. The nosewheel
configuration is inherently stable.

Principles of Flight Stability 13-11


With the CG placement behind the main wheels, the tailwheel configuration is
inherently unstable. If a tailwheel aircraft is displaced from a straight course on
the ground, the centrifugal reaction force tends to increase the rate of turn. This
means that once the aircraft is made to enter a turn, a positive force must be
applied-e.g. through a tailskid, differential braking, steerable tailwheel,
slipstream past the rudder-to control the rate of turn and prevent it from
increasing too much.
,.,
''
''\
\

Fig. 13-16. The tailwheel configuration


is inherently unstable.
tailwheel aircraft

If a tailwheel aircraft is allowed to swing (or turn) about the main wheels at too
high a rate, it is possible that the ability to control the turn with rudder and/or
differential braking will be lost. When this occurs, the aircraft will spin rapidly
and uncontrollably about the main wheels in what is called a groundloop. To
avoid the possibility of a groundloop in tailwheel aircraft-and particularly with
those aircraft fitted with castoring tailwheels-care must be taken to prevent
large swings from developing during the landing roll-out.

Control on the Ground


Directional control on the ground is achieved by the use of rudder, nosewheel
steering (which may be connected to the rudder pedals), power and brakes.
Slipstream over the rudder increases its effectiveness, but in the directionally
stable nose wheel aircraft, the use of rudder alone is usually insufficient to give
good directional control on the ground. In contrast, many taildraggers can be
controlled mainly with the use of rudder alone.
/
/
RIGHJ'TURN
Fig. 13-17. The rudder pedals, I
I
the nosewheel and the rudder I
in a turn to the right.
2. nosewheel turns
to right

aerodynamic force
generated by slipstr am
moves tail left
nose right

3. rudder moves right

Sharp turns at high speed should be avoided-a high CG, narrow wheelbase and
an unfavourable wind may combine to tip the aircraft onto the outer wingtip.
Any wind tends to weathercock the aircraft into wind, so care is needed when
taxiing in crosswinds or tailwinds.

13-12 Stability The Commercial Pilot Series


Asymmetric Blade Effect

As explained in the previous chapter, when the axis of rotation of the propeller
becomes tilted back with respect to the direction of travel, the downgoing
blades in the propeller disc develop more thrust than the upgoing blades. As a
result, the line of thrust is shifted toward the downgoing-blade side of the disc
and a yawing moment is produced.

This asymmetric blade effect is significant during the initial part of the take-off
run in a tailwheel aircraft, when it is travelling across the ground in a tail-down
attitude. If the propeller rotates clockwise when viewed from the cockpit, the
downgoing blade side will be to the right and the aircraft will tend to swing to
the left. When the tail is raised later in the take-off run, the axis of rotation
becomes more nearly aligned with the direction of travel and asymmetric blade
effect is reduced.

Summary of Effects

As shown in Fig. 13-25, all of the foregoing causes of swing on take-off act in the
same direction. For engines which rotate clockwise when viewed from the rear
the tendency to swing on take-off is to the left, requiring right rudder to
counteract it. swing to left

~
Tailwheel aircraft exhibit the
greatest tendency to swing on asymmetric blade effect

take-off, with the tendency being gyroscopic effect


most pronounced during the early
part of the take-off when the
tailwheel is still on the ground. In
some aircraft, the increased
tendency to swing due to NOTE: For clockwise
rotation of engine when
gyroscopic effect is quite viewed from cockpit
noticeable as the tail lifts,
particularly if this is done quickly.
As the take-off proceeds from this
point, the tendency to swing
reduces. Fig. 13-25. Summary of the four reasons for the
tendency to swing on take-off.

Nosewheel aircraft exhibit much less tendency to swing on take-off. As the


aircraft is more nearly in the flying attitude and the axis of rotation of the
propeller is more closely aligned with the direction of travel, asymmetric blade
effect and gyroscopic effect are minimal. The main reasons for a nosewheel
aircraft to swing are slipstream effect and torque reaction but, as this type of
undercarriage configuration is inherently stable, any tendency to swing for these
reasons is less apparent and easier to control than in a tailwheel undercarriage
aircraft.

Crosswind Take-offs and Landings


Crosswind take-off

The main difficulty presented by a take-off in a strong crosswind is that of


keeping the aircraft tracking straight down the runway. A crosswind will tend to
weathercock the aircraft away from the runway direction and, depending on its
direction, this will add or subtract to already-present tendency to swing for the
reasons just described.

13-16 Stability The Commercial Pilot Series


yaw to left

Fig. 13-22. Slipstream effect.

aerodynamic force
Torque Reaction

With engine torque rotating the propeller clockwise, the reaction to this torque
tends to rotate or roll the engine (and the aircraft) in an anti-clockwise direction.
The aircraft is prevented from rolling when it is on the ground by the main
wheels, but as a result of torque reaction, more of the aircraft weight is
supported by the left main wheel than the right wheel. The left main wheel
therefore has a greater rolling resistance than the right and, because of this, the
aircraft will tend to swing to the left until the wings take the weight off the main
wheels.

engine torque

Fig. 13-23.
Torque reaction effect. left whee
increased --"fi~~-d!£,-­
resistance:
yaw to left

Gyroscopic Effect

At high rpm, the rotating propeller (together with the rotating masses of the
engine) have the same properties as the rotor of a gyroscope. In a tailwheel
aircraft, as the tail is lifted off the ground during take-off, this is the equivalent of
applying a fmward-tilting force at the top of the propeller disc. With gyroscopic
precession, this force is felt 90' further on in the direction of rotation-i.e. with
clockwise rotation, it appears as a fmward-tilting force on the right hand side of
the propeller disc. Hence, as the tail is coming up, gyroscopic precession is felt
as a yaw to the left.

1. tail rises 2. force (in effect)


applied here""''""~~~~>'-,

r ....--~

4.
causing
swing
3. is precessed thru 90" to here to left

Fig. 13-24. Gyroscopic effect


as the tail rises.

Principles of Flight Stability 13-15


To avoid a crosswind from behind lifting the upwind wing, lower its aileron
so that the wind cannot get under it, by moving the control wheel out-of-wind. A
quartering tailwind from behind the aircraft and to one side is the most difficult
and hazardous taxiing condition. Hold the control wheel forward and out-of-
wind, and maintain directional control with the rudder pedals, using differential
toe brakes when necessa1y. Avoid any sudden braking or sudden power
increases.
left

o•
quartering
tailwind

DOWN UP P••

left
quartering
tailwind
I '
down

Fig. 13-20. Taxiing in a left-quartering tailwind.

NOTE: The propwash or jet blast from


another aircraft can produce the same
effect as a wind. Always be cautious if
you have unavoidably to taxi or park
your aircraft close behind another
aircraft, especially at an angle.
when wind is from i hemisphere
control wheel forward and out-of-wind
Fig. 13-21
Summary of the use
of controls when taxiing
in windy conditions.

Swing on Take-off
There is a tendency for single-engine propeller driven aircraft-particularly those
with a tailwheel undercarriage-to 'swing' to one side on take-off. The causes of
swing on take-off are:

• slipstream effect;
• torque reaction;

• gyroscopic effect;
• asymmetric blade effect; and
• crosswind (weathercocking tendency as already explained).

Slipstream Effect

The rotation of the propeller imparts a 'corkscrew' motion to the slipstream as it


travels back past the fuselage. For propellers rotating clockwise when viewed
from the cockpit (the more common direction of rotation), the slipstream meets
the fin and rudder at an angle which generates a yaw to the left. This yawing
effect of slipstream is most pronounced at high power/slow speeds, e.g. during
take-off. Some aircraft have an offset or biased fin (i.e. with a small built-in
angle of incidence) to help counter slipstream effect.

13-14 Stability The Commercial Pilot Series


Speed is controlled on the ground by the use of power and brakes. Power
applied with the throttle is the normal means of accelerating the aircraft and,
once moving, power can usually be reduced. Air resistance, ground friction and
wheelbrakes are the means of slowing the aircraft. It is good ainnanship not
to use power against the brakes.

Hard braking may cause a taildragger to tip over on its nose, and any braking
tends to destabilize it directionally. Although the same hazards do not exist with
the nosewheel (or tricycle) undercarriage aircraft, it is a good policy to avoid
especially hard or harsh braking on any aircraft and to demonstrate plenty of
anticipation of the need to slow down.

Allowing for Wind Effect when Taxiing

When taxiing, the controls should be held in a position to avoid either the tail or
a wing being lifted by a strong wind (or a strong gust).

When taxiing into a strong headwind in a nosewheel aircraft hold the control
wheel so that the elevator is neutral or back. Ideally, the weight carried by the
nosewheel will neither be too little (causing steering difficulty) nor too great. In
a taildragger, hold the control column back to keep the tailwheel firmly on the
ground.

When taxiing with a strong tailwind, hold the control wheel fmward to move
the elevator down. This stops the wind from lifting the tail from behind.
tailwind

~
~
...

elevator forward

Fig. 13-18. Taxiing into a strong headwind and.......................... with a strong tailwind.

A crosswind will try to weathercock the aircraft into-wind because most of the
keel surface is behind the main wheels. This weathercocking tendency is
greater in tailwheel aircraft than in those fitted with a nosewheel.

The rudder pedals, especially if nosewheel steering is fitted, should provide


adequate directional control to steer a straight path even in a strong crosswind,
but, if not, use differential braking to assist. The weathercocking tendency due
to a crosswind also makes it easier to turn the aircraft upwind, and harder to
turn it downwind.

To avoid a crosswind from ahead lifting the upwind wing, raise its aileron by
moving the control wheel into wind. This also applies for a crosswind directly
from abeam- control wheel into wind.

quartering
left ' left
headwind quartering
--~---tr~------ headwind

UP DOWN

1
up
Fig. 13-19. Taxiing into a left-quartering headwind.

Principles of Flight Stability 13-13


Prevent any tendency to swing with rudder. With all aircraft types, but
particularly those with a tailwheel, 'nip any tendency to swing in the bud' with
small and timely rudder corrections, rather than have to apply large amounts of
rudder to make a 'late' correction for a swing which has been allowed to
develop.

Use 'into-wind' aileron to prevent any tendency for that wing to lift. The 'down'
aileron may also provide some drag to help counteract weathercocking yaw.
The amount of aileron used should be progressively reduced during the take-off
run-to zero at lift-off.
crosswind

Fig. 13-26.
Ill
Crosswind take-off. aileron 'into-wind'

Crosswind Landing

Once on the ground, the problems with keeping straight in a crosswind landing
are similar to those of the crosswind take-off. Use rudder to prevent any swing
and apply aileron into-wind. Again, the handling difficulties are greater for the
tailwheel aircraft than for the nosewheel type.

The main challenge in a crosswind landing comes at the point of touchdown


where, ideally, the aircraft should be headed along the runway centreline and
not have any cross-runway drift. There are two techniques for achieving this:

• On approach to land and throughout the flare, the


aircraft is tracked down the extended runway centreline
with the appropriate drift angle applied. (The aircraft is
flown somewhat crabwise with the coordination ball
centred and the wings level except where adjustments
to tracking are required). Just prior to touchdown, the
aircraft is 'yawed straight' with rudder so that it
becomes aligned with the runway heading. Fig. 13-27
refers. The timing of the yaw is critical. If the aircraft is
yawed straight but does not immediately touch down, it
will begin to drift toward the downwind side of the
runway. If touchdown occurs before the yaw is
complete, the wheels will not be properly aligned with
the direction of travel. In both of these cases, the
landing will impose some side strain on the landing gear
and some initial difficulty with directional control may
be experienced.

• A second crosswind landing method is called the 'wing-


down technique'. During the latter part of the approach,
Fig. 13-27. Crosswind landing.
the fore/aft axis of the aircraft is aligned with the runway
centreline with rudder and, at the same time, is
prevented from drifting to one side or the other by adjusting the bank angle
into the crosswind. (The aircraft is being flown slightly uncoordinated with
side-slip-although it may sound complicated, it is a relatively simple
procedure to fly.) Fig. 13-28 refers. Using this technique, the aircraft is
flared with bank applied and can then either be rolled level just prior to
touchdown, or landed gently on the upwind mainwheel first.

Pnnc1ples of Flight Stability 13-17


The strongest crosswind that an aircraft can handle is limited by rudder
effectiveness and a maximum demonstrated crosswind component is specified
in the Pilot's Operating Handbook.

Fig. 13-28. Crosswind landing:


the 'wing down' technique.

Ground Effect
Ground effect is the so-called 'cushioning' benefit which is obtained when an
aircraft is flown at ve1y low level above a smooth surface. Ground effect
becomes noticeable at a height above the surface of less than one wingspan and
increases in effect the closer the wing is to the surface (which can be land or
water).

D a. OUT OF GROUND EFFECT


.---~TR

b. IN GROUND EFFECT
drag reduced

,--.~.TR
lift ~
improved
vorticity reduced

~~~~~~~~~~~~~~~i~downwash restricted

same
effective

Fig. 13-29. Flying in ground effect.

13-18 Stability The Commercial Pilot Series


Review 13

I. Static stability refers to the (initial/subsequent) tendency of an aircraft to


return to a position of equilibrium after a disturbance.

2. An aircraft which has static stability but then continues to diverge from its
original attitude after a disturbance is ................................. unstable.

3. The ................................... provides the main means of obtaining longitudinal


stability in an aircraft.

4. For longitudinal stability, the tailplane restoring moment must be (less/


greater) than any unstable wing (and fuselage) pitching moment.

5. The further forward the CG, the (less/more) stable the aircraft in pitch.

6. If flown with the CG behind the aft limit, an aircraft will have (sufficient/
insufficient) longitudinal stability; will be (easier/more difficult) to fly: and
flight safety (may/may not) be jeopardised.

7. If the fin and keel surface area behind the CG is increased, directional
stability will be (reduced/improved).

8. With an aft CG position, directional stability will be (reduced/improved).

9. Lateral stability relies on the ................................ which develops after a


disturbance in roll.

I 0. Which of the following factors contribute the most to lateral stability;


(anhedral/dihedral): (shielding/no shielding); (high wing/low wing); (high/
low) fin and keel surface; (straight/swept) wings.

II. An aircraft which has strong directional, but weak lateral stability, will tend
towards ............................. instability.

12. An aircraft which has strong lateral stability by comparison with its
directional stability, will tend toward a form of oscillatory dynamic instability
called ......................................... .

13. The nosewheel (tricycle) undercarriage configuration is inherently (stable/


unstable) in the ground roll. In contrast, the tailwheel configuration is
inherently (stable/unstable).

14. List the five possible causes of a swing on take-off in a tailwheel aircraft.

13-20 Stability The Commercial Pilot Series


Ground effect arises through the reaction of the vortices behind the wing with a
close-proximity surface. The ability of the vortices to flow freely becomes
restricted and, as a result, downwash behind the wing is reduced. With a
reduced average downwash angle when flying 'in ground effect' than when out
of it, the Total Reaction vector of the wing is tilted further forward which, in turn,
means:
• More lift is produced at the same geometric angle of attack or, alternatively,
the same lift can be produced at a reduced geometric angle of attack.
• There is a reduction in induced drag and, therefore, of total drag.

In short, the wing is more efficient 'in ground effect' than out of it. The
aerodynamic process involved is the same as would occur if the aspect ratio of
the wing was suddenly increased-a reduction in the downwash angle leading
to a reduction in induced drag and an improvement in lift. Fig. 4- I I also refers.

Ground Effect on Take-off and Landing

As the aircraft climbs out of ground effect on take-off, lift will decrease and
induced drag will increase. A slight 'sagging' in take-off performance will be
noticeable, but this does not normally present a problem unless the aircraft is
being operated at very high weight from a short strip.

On final approach to land, the aircraft will enter ground effect at about one wing-
span height. Ground effect generally tends to cause the aircraft to 'float' briefly
before touchdown. To prevent this floating in ground effect from becoming
excessive, avoid having excess speed at the beginning of the flare and ensure
that power is properly reduced.

Principles of Flight Stability 13-19


Asymmetric Flight
Introduction

In normal flight in a multi-engine aircraft, the lines total thrust


of total thrust and total drag act through, and are
balanced about, the normal axis (Fig. 14-1 refers).
With all engines at the same power setting, there equal thrust
are no residual yawing moments resulting from from engines
the offset engine thrust lines and, in steady flight,
total thrust will equal total drag.

normal axis
through CG
Fig. 14-1. Symmetrical balance of thrust and
drag forces about normal axis.

Unless all of its engines share the same thrust line, any multi engine aircraft
which suffers a failure of one of its power-plants has a potential asymmetric
flight problem in that:

• the line of total thrust will be offset from the yawing moment
normal axis, causing a yawing moment
toward the failed engine; ~
thrust
• the line of total drag moves toward the failed
engine, which adds to the asymmetric thrust
yawing moment; and
• there is a reduction of total thrust available
which leads to a deterioration in failed
engine
performance.

Fig. 14-2. With one engine failed, thrust and


drag act at a distance from the normal axis,
causing a yawing moment toward the failed
engine.
total drag

Following an engine failure, the immediate concern is to control the yaw and
any roll which results and maintain a safe flightpath. Then, with the appropriate
drills completed and the failed engine 'secured', the aim is to fly the aircraft in
such a way as to achieve optimum performance with the asymmetric thrust and
reduced power available. In this chapter, we consider the aerodynamic factors
involved in asymmetric flight in a twin-engine propeller-driven aircraft.

Principles of Flight Asymmetric Flight 14-1


Yawing Moment
The strength of the yawing moment due to the asymmetry of thrust and drag
about the CG is proportional to:

The thrust from the live engine. The higher the thrust, the greater the yawing
moment.

The distance of the thrust-line from the CG. Increased distance increases the
arm through which the thrust from the live engine acts and therefore increases
the yawing moment.

Directional stability. To obtain directional stability, the aircraft is constructed


with the greater proportion of fin and rudder surface area behind the CG than
ahead of it. This provides a 'weathercocking' effect which acts to keep the
fuselage aligned with the direction of flight. The weathercocking moment resists
the asymmetric thrust yawing moment, and it increases with airspeed.
Directional stability is reduced with rearward movement of the CG, and
weathercocking moment is therefore less effective when the aircraft has an aft
CG!oading.

The rate of thrust decay. If the engine failure is gradual, the onset of the
yawing moment will also be gradual.

The drag of the failed engine. The drag of a failed engine will be higher than
one which is operating normally-particularly if the propeller is windmilling.
The greater the increase in drag on the 'dead engine' side, the stronger the
yawing moment toward that side.

The total yawing moment following an engine failure can be very strong-
particularly at high power and low airspeeds. The potential handling difficulty is
therefore worst under these conditions (e.g. during take-off and initial climb)
than in any other phase of flight.

Rolling Moment
Failure of an engine also causes a rolling moment to develop, mainly through:

• The further effect of yaw-as has been explained at Chapter 7 (Fig. 7-12
provides a summa1y).
• The loss of slipstream over the wing, resulting in a loss of lift on the side of
the failed engine.
Note that the roll is also in the direction of the failed engine.

Other Factors
Other factors which affect the yawing and rolling moments to a degree
depending on which engine has failed are:

Propeller Torque Reaction

The action of the engine in turning the propeller produces a reaction which
tends to roll the aircraft in the opposite direction about the engine. This is called
propeller torque reaction.

14-2 Asymmetric Flight The Commercial Pilot Series


Most piston-engined aircraft have powerplants which rotate clockwise when
viewed from behind (those with engines of United States origin). With these
aircraft, propeller torque reaction therefore tends to roll the aircraft anti-
clockwise around the live engine. In this case, as shown in Fig. 14-3, propeller
torque reaction after failure of the right engine works in opposition to the roll
generated by the failure of the engine-i.e. it aids in 'lifting' the failed engine side
and reduces the amount of aileron needed to correct it. On the other hand,
propeller torque reaction acts in conjunction with the rolling moment caused by
a failed left engine, requiring more aileron to correct for it.

torque reaction - away from failed engine torque reaction - toward failed engine

~---fbi==='
~

@t~gine
propeller rotation failed
=c =,~C=;==" failed propeller rotation

Fig. 14-3. Propeller torque reaction.

Asymmetric Blade Effect

Asymmetric blade effect has already been explained in Chapter 12. The
explanation is repeated here, with an alternative diagrammatic presentation.

In the normal cruise and at low angles of attack, the thrust line can be taken for
all practical purposes to coincide with the direction of flight. Under these
conditions, as shown in Fig. 14-4, the upgoing and downgoing propeller blades
have the same angle of attack (a) and relative airflow. Thrust is therefore
produced evenly from either side of the propeller disc and the line of thrust from
each engine can be taken as acting through the propeller hub.

forward velocity v

relative airflow
upgoing
upgoing blades rotational velocity r

')l axis of
'
direction of flight rotation

downgoing blades

downgoing

upgoing blades

downgoing blades
downgoing
(faster, higher a)

Fig. 14-4. Asymmetric blade effect. v

Principles of Flight Asymmetric Flight 14-3


At high angles of attack however, the thrust line becomes tilted back with
respect to the direction of flight. As can be seen from the lower diagram in Fig.
14-4, the geometry of the rotational and forward velocity vectors is changed,
such that the downgoing blades now have a higher angle of attack than the
upgoing (even though the blade angles have remained the same). In addition,
there is a faster relative airflow past the downgoing blades. This is shown by the
longer relative airflow vector in the diagram. (In effect, when the propeller disc
is tilted, the downgoing blades advance by the amount of tilt in one-half
revolution; the upgoing blades retreat by the same amount).

The net result is, that at high angles of attack, the downgoing blades produce
more thrust than the upgoing, and the thrust line of each engine becomes
displaced toward the downgoing-blade side of the propeller disc as shown in
Fig. 14-5. greatest yawing moment
with critical engine failed

In normal flight at high angles of


attack, the displacement of the thrust
lines is of no consequence.
However, it can be seen from the
diagram, that failure of the left engine
would result in a significantly larger
yawing moment than a failure of the
right engine. The engine with the
smallest moment arm is therefore
the worst to have fail and it is
referred to as the critical engine.
left engine right engine
arm arm
NOTE. The diagram is for an aircraft
with engines that rotate clockwise
when viewed from the rear, in which
case the left engine becomes the
critical engine. Fig. 14-5. Displaced thrust lines owing to
asymmetric blade effect.

Immediate Actions
When an engine failure occurs, the immediate actions required of the pilot to
retain control of the aircraft are:
• prevent further yaw with rudder; and
• prevent unwanted roll with aileron.

Identification of the Failed Engine

In most cases of sudden engine failure in a multi-engine aircraft, it is important


that the failed engine is quickly identified and the appropriate actions taken in
accordance with the Pilot's Operating Handbook without delay. If the fault
cannot be rectified, these actions will include feathering the propeller on the
failed engine, or, if it is of the non-feathering type, placing the propeller control in
FULL COARSE.

It is stressed that the correct identification of the failed engine is important.


Accidents have occurred in the past through pilots mis-identifying the problem
engine, resulting in the feathering and shut-down of a perfectly serviceable
engine.

14-4 Asymmetric Flight The Commercial Pilot Series


The most immediate indication of which engine has failed in a twin-engined
aircraft, is the direction in which the aircraft yaws when the failure occurs. The
aircraft will want to yaw toward the failed engine. Most pilots will however, use
rudder instinctively to prevent this yaw, in which case rudder pressure becomes
the indication of the failed engine-i.e. rudder pressure is needed away from the
failed engine. You may find it easier to remember this by using one of the
following sayings:
''working leg-working engine" or
"lazy leg-lazy engine"
The identification of the failed engine must then be confirmed through closing its
throttle (if yaw results, the wrong engine has been selected) and engine
instrument indications. (For aircraft with more than two wing-mounted engines,
yaw and control pressure will indicate only the side of the failure and proper
identification can only come from the engine instruments/warning devices).

Modes of Constant-heading Asymmetric Flight


Provided sufficient speed is maintained, an aircraft can be flown on a constant
heading under asymmetric thrust conditions with various combinations of
rudder (i.e. side-slip) and bank applied. We discuss these modes of asymmetric
flight under three headings:
• all rudder;
• all bank; and
• a combination of rudder and bank.

All Rudder

In this mode, rudder is used to prevent the yaw from the asymmetric thrust and
the wings are held level with the use of aileron. In a colloquial sense, the aircraft
is crabbing along slightly sideways with the nose pointed a few degrees away
from the direction of flight.

In using rudder alone to prevent the yaw from asymmetric thrust, a lateral rudder
side force is created which, if left unbalanced, would tend to push the aircraft
sideways toward the failed engine. The counter to this rudder side force comes
from the inherent directional stability of the aircraft. When just sufficient rudder
is used to prevent any yaw (with the wings held level), the aircraft adopts a small
side-slip angle and a 'weathercock' side-slip force is generated which opposes
the rudder side force. When these forces are in balance, the aircraft will
maintain a constant heading wings-level with the rudder yawing moment equal
and opposite to the thrust yawing moment plus the weathercock yawing
moment-see Fig. 14-6.

Note that the aircraft is constantly side-slipping at a small angle and thus the
heading and direction of flight do not coincide. Because of this side-slip, the
dihedral of the wings generates a rolling moment which must be constantly
countered with aileron applied toward the failed engine. The forces are
balanced and, in spite of the side-slip angle, with the weight and lift vertically
aligned, no slip is indicated on the coordination ball. The control forces
generated by the constant deflection of rudder and aileron can be trimmed out
with the trim controls and the flight instrument indications are:
• wings level;
• ball centred;
• apparent drift toward the failed engine.

Principles of Flight Asymmetric Flight 14-5


TOP VIEW REAR VIEW ALONG FLIGHTPATH

thrust yawing weathercock


moment yawing moment
---------..aileron rolling

\r
I
direction of fhght

\ /Sideslip angle
dihedral rolling~
moment
moment

I lift (l)

I
I
I
I
:)D

side-slip
force

weight (W)

I
I
I
I
rudder yawing moment - - ~ - ~o skid indicated
I
I

Fig. 14-6. Asymmetric balance of forces with wings level.

Although this mode of asymmetric flight is the easiest to fly accurately because
of the straightforward instrument indications, it is not particularly efficient owing
to the extra drag generated by the side-slip angle of the fuselage and the
constant deflection of the aileron surfaces.

All Bank

The arrangement of forces in the 'all bank' mode of asymmetric flight is shown
below in Fig. 14-7.

~
dihedral rolling - - - - - - - - . . . . .
moment
thrust yawing . . .
moment d!rectJon of flight ~ aileron rolling
moment
I
l

weight
side force
weight side-slip
side force force

w
\
\
slip indicated
weathercock yawing \ - ·

~
moment

Fig. 14-7 The 'all bank' mode.


--u: - ....\:;7'")

\
toward live
engine

14-6 Asymmetric Flight The Commercial Pilot Series


In this mode, the asymmetric thrust yawing moment is counteracted without the
use of rudder-i.e. with rudder remaining in its normal central position. The
aircraft is banked and is constantly 'side-slipped toward the live engine. The
resulting weathercocking yawing moment is used to counteract the thrust
yawing moment.

As can be seen from the diagram, this mode involves a relatively large side-slip
angle and a high angle of bank (perhaps as much as 15°). It is difficult and
uncomfortable to fly and very inefficient. The tilting of the lift vector requires the
aircraft to be flown at a higher angle of attack to counteract the weight in level
flight and this, coupled with the side-slip angle, results in higher drag than is
necessa1y.

The all bank mode is seldom used but is shown here to demonstrate the
principles involved. Note that by banking the aircraft toward the live engine, a
component of weight acts in the lateral plane and can be used to offset side-slip
(or rudder) side force. Another way of looking at it is that the tilted lift vector
provides the side force to balance the rudder side force. The turn which would
normally result from this sideways component of lift is being prevented in this
case (when the appropriate angle of bank is held) by the opposite turning effect
of the asymmetric thrust. Note also that although the forces are balanced, a
large amount of slip is indicated toward the live engine. With no unbalanced
slip or skid the coordination ball merely slides to the bottom of the tube to
indicate the direction in which weight is acting.

Combined Rudder and Bank

The more normal method of controlling the aircraft on a constant heading in


asymmetric flight is to use a combination of rudder and bank. The amounts of
rudder and bank can be varied between the two extremes represented by Fig.
14-6 (all rudder, no bank) and 14-7 (all bank, no rudder). The ideal arrangement
is, with the longitudinal axis of the aircraft aligned with the direction of flight, to
have sufficient rudder applied to counteract the thrust yawing moment, together
with a small amount of bank toward the live engine to balance out the rudder
side force. The arrangement of forces in this particular combination is shown in
Fig. 14-8 below.
thrust yawing moment
.----------.-..
T
direction of flight

w
I
I
I slight slip indicated
--~-- toward
~ 4-"' live engine
rudder yawing moment

Fig. 14-8. Ideal balance of forces using combined rudder and bank in asymmetric flight.

Principles of Flight Asymmetric Flight 14-7


The arrangement in Fig. 14-8 would provide the optimum performance since,
with zero side-slip angle, the extra drag is kept to a minimum. In attempting to
fly this mode however, the bank angle should be contained to between 5 and
10°, othe1wise the gains made by having zero side-slip will be lost in having to fly
at a higher angle of attack to maintain the vertical component of lift. Rather than
increase bank with zero side-slip, it is better to use more rudder and accept a
small side-slip angle.

In practice, a mode somewhere between the 'rudder and bank' and the 'all
rudder' modes is usually adopted. To get there, the aircraft is initially flown in
the 'all rudder' mode (Fig. 14-6) with the wings level, and then about 5o of bank
is applied toward the live engine and the amount of rudder reduced so that a
constant heading is maintained. In this 'hybrid' mode, the aircraft is flown with:
• a small side-slip angle, with the nose slightly offset toward the live engine
(and therefore a small amount of apparent drift);
• a small amount of bank (around 5°) toward the live engine; and
• with a small amount of slip indicated toward the live engine.

Minimum Asymmetric Control Speeds


If an asymmetric' aircraft is flown at decreasing lAS with the live engine at full
power, a point will be reached where full rudder is required to prevent yaw.
This point is the minimum asymmetric control speed, below which control in
yaw cannot be properly maintained by normal means 2 •

VMCA (standing for the minimum control speed airborne) is the minimum control
speed following a sudden failure of the critical engine after take-off at which an
average pilot will be able to maintain directional control with full rudder and no
more than 5o bank applied. To obtain airworthiness certification, VMCA must be
demonstrated under a very specific set of circumstances' to be not greater than
1·13 times the level stalling speed in the same configuration.

VMCG (standing for the minimum control speed ground) is the minimum control
speed at which, after failure of the critical engine during the take-off run, it is
possible to maintain directional control with rudder control alone. This speed
assumes the take-off will be continued. It is not applicable to light twin aircraft.

For twin-engine aircraft, VMCA (or VMc, standing for minimum control speed in US
terminology) will probably be stated in the aircraft flight manual and marked on
the airspeed indicator as a radial red line at the lower end of the scale. This
minimum speed should be treated with a great deal of caution and, when flying
asymmetric, a good policy would be to treat it as one would a stalling speed and
keep a respectable margin above it so that control can be assured.

NOTES:
1. We are concerned here with light twin-engine aircraft. On some low
powered types, and at altitude, the VMCA may be below the stalling speed in the
configuration being flown.
2. Other means of maintaining directional control are by reducing power on
the live engine, or by using an inordinate amount of bank toward the live engine.
Both of these measures will normally result in a loss of altitude.
3. These conditions include: critical engine windmilling; full power on the live
engine; flaps at the take-off setting; undercarriage retracted; CG at the aft limit; at
MCTOW; and under sea level !SA conditions.

14-8 Asymmetric Flight The Commercial Pilot Series


Review 14
The following questions refer to a twin-engine propeller driven aircraft (wing-
mounted engines).
1. With both engines operating normally in flight, thrust and drag forces will be
balanced about the ....................... axis.
2. If one engine fails, the lines of thrust and total drag become offset from the
normal axis, causing an asymmetric yawing moment toward the (live/failed)
engine.
3. Following from Q 2, the asymmetric yawing moment will be strongest
when:
(a) thrust from the live engine is (high/low);
(b) the distance between the thrust line and CG is (long/short);
(c) the drag from the failed engine is (low/high); and
(d) the aircraft has a low weathercocking moment, i.e., when the CG
position is (forward/aft) and the airspeed is (high/low).
4. Engine failure also causes a rolling moment toward the (failed/live) engine,
through the further effect of ................. and the loss of ......................... over
the wing.
5. For engines which rotate clockwise when viewed from the rear, propeller
torque reaction after an engine failure, tends to roll the aircraft to the (left/
right).
6. From Q 5, which is the worst engine to have fail from the point of view of
propeller torque reaction?
7. At high angles of attack, asymmetric blade effect results in a shift of the
thrust line of the engines (toward/away from) the downgoing side of the
propeller disc. Therefore, for the usual direction of engine rotation, the
thrust lines are moved to the (right/left) making the (right/left) engine the
critical engine to have fail.
8. In preventing yaw after engine failure, rudder pressure is needed (toward/
away from) the live engine. State two other means of identifying (or
confirming the identification ot) which engine has failed.
9. In the 'all-rudder' method of maintaining a constant heading in asymmetric
flight, the wings are held level with aileron and rudder is applied to prevent
yaw. The balance ball is (centred/offset) and there is apparent drift toward
the (live/failed) engine.
10. In the 'all bank' method the rudder is central. The asymmetric yaw from the
live engine is prevented by banking and side-slipping (toward/away from) it.
11. In practice, a mode between 'all rudder' and 'all bank' is flown. In this
mode:
(a) the nose will be offset slightly (toward/away from) the live engine, and
there will be a small amount of apparent drift;
(b) a small amount of bank is applied (toward/away from) the live
engine-not more than about ... o; and
(c) a small amount of slip will be indicated toward the (live/failed) engine.

Principles of Flight Asymmetric Flight 14-9


14-10 Asymmetric Flight The Commercial Pilot Series
Range and Endurance
Introduction

In this chapter, we cover the subject of flying for maximum range in two parts.
In the first, we discuss the theory of range flying, In the second, we cover the
practical aspects of planning a flight at maximum range using the information-
tables and charts-which will normally be available for this in the Aircraft Flight
Manual.

Both the theoretical and practical aspects of flying for maximum endurance, are
then discussed toward the end of the chapter.

Range Flying -Theory


Maximum range in level flight is achieved when the greatest distance is covered
for the amount of fuel used. Specific Air Range (SAR) is defined as the air
distance flown per unit quantity of fuel. For US-sourced aircraft, SAR will
normally be measured in air nautical miles per US gallon (anm/USG). Thus:

air distance flown


specific air range (SAR) = fuel used

Dividing both top and bottom lines of this equation by time, we get:

air distance flown time


specific air range (SAR) = =--'--''-'-'-'=-----'-'-'-'-' X
time fuel used

TAS
= GFC
(equation 1)

(where GFC =gross fuel consumption, or fuel flow, in units of fuel (e.g. USG) used per hour)

Hence, for the greatest specific air range, we need to be flying at the highest TAS
for the lowest gross fuel consumption. Developing this theme a little further, for
a piston engine, specific fuel consumption (SFC) is defined as the GFC (or
fuel flow) per unit of power produced, i.e:

GFC
specific fuel consumption (SFC) = power

or GFC = SFC x power


substituting in equation I above:
TAS 1
SAR = power X (equation 2)
SFC

From equation 2, it can be seen that to achieve maximum specific air range
(SAR), the aircraft must be flown for the maximum product of airframe
efficiency (TAS/power) and engine efficiency (1/SFC).

Principles of Flight Range and Endurance 15-1


Aitirame Considerations: Piston-engine Aircraft Range
Maximum airframe efficiency and the best range speed (maximum TAS/power
ratio) occurs at the speed for minimum drag/maximum UD ratio. As shown
back in Fig. 8-10, while this point lies at the bottom of the drag (or thrust) curve,
on the power required curve, it occurs where the line drawn from the oligin is
tangential to the curve. As can be seen from Fig. 15-1 below, this is the point at
which the ratio of TAS obtained for power applied is the highest. At any other
power setting the slope of the line drawn from the origin will be steeper and the
ratio of TAS/power will be poorer. However, note from the shape of the curve,
that at speeds slightly above or below the best range speed, there is little change
in the slope of the line and therefore ofTAS/power ratio.

power
required for
level flight

Fig. 15-1. Best range speed is the


speed for Min. drag/best UD ratio.
This is the speed which gives the best
ratio of TAS/power required.

0 t TAS
TAS (still air)
for maximum range

Effect of Altitude

At a given aircraft weight, the lAS for Min. drag/max. UD ratio and therefore, best
range speed, remains constant with altitude. As altitude is increased, both the
TAS and the power required at the best range speed increase in the same
proportion. The TAS/power ratio remains unchanged and therefore, from the
airframe point of view, altitude has no effect on the best range lAS. ('Best
range' TAS increases with altitude.)

Effect of Weight

An increase in weight means that the angle of attack for best UD ratio is reached
at a higher !AS. The best speed for range is therefore increased. While this
increase in !AS means that the TAS will be higher, drag has also increased, and
the power required has increased out of proportion to the gain in TAS
(remember power = drag X TAS). Range is therefore reduced if weight is
increased.

Effect of Wind Velocity


So far, we have confined our consideration to specific air range (SAR) which is
the range which is achieved in the air mass. To determine the actual range
which will be flown across the ground, we have to take account of the effect of
wind velocity. Whereas the SAR is dependent upon TAS/power, the specific
ground range (SGR)-the ground distance flown per unit quantity of fuel-
is dependent on groundspeed/power. A headwind component will decrease
the range achieved, while a tailwind component will increase it.

15-2 Range and Endurance The Commercial Pilot Series


To obtain maximum ground range, we must fly at the speed which provides the
highest ratio of groundspeed/power (i.e. the highest groundspeed for the least
amount of power being used). The optimum speed in headwind/tailwind
conditions can be found by locating groundspeed on the PR!I'AS graph (in the
manner shown in Fig. 15-2) and finding the speed (TAS) to fly below the
redrawn tangent to the PR curve.

power
required
(PR) \
\

' -~"'.v·
< ••·,
r i'
-
- -.............../!
.....•' I
I
I
'''
~
- - -..................····· I :
-- _..,... ,/ .J.. I
-- __ . . ...·· v v TAS (GIS nil wind)
0 50 100 I
.,
I
GIS 50 kts tailwind
0 50 100 150

GIS 50 kts headwind


0 50 100 150

Fig. 15-2. The effect of wind velocity on best range speed.

Note how the optimum speed for range increases in a headwind, decreases in a
tailwind. In practice however, unless the headwind component exceeds 25% of
TAS, or a tailwind component exceeds 33% of TAS, adjustments are not usually
made.

Summary

From the airframe efficiency point of view, maximum range is achieved by flying
at the minimum drag speed. Altitude has no effect on the best lAS to fly, but if
weight is increased, range is reduced and the best speed to fly is increased.
Range reduces in a headwind, increases in a tailwind. In strong headwinds, the
optimum speed for range is increased, and in strong tailwinds, the optimum
speed is reduced.
Recommended Range Speed

In practice, if the Aircraft Operating Manual gives a recommended range speed,


it will be about I 0% higher than the minimum drag speed. The reason for this is
that when flying at the minimum drag speed, manoeuvring or turbulence can
place the aircraft too far into the 'wrong side of the drag curve' and require
additional power to be applied to regain speed, resulting in inefficiency. It is also
generally more comfortable and easier to fly accurately at the higher speed. As
indicated previously, the TAS/power ratio does not change significantly with
small changes of speed about the min. drag speed, hence the reduction in range
through flying at the higher speed is insignificant.

Principles of Flight Range and Endurance 15-3


Engine Considerations
As stated previously, for best range, the aircraft should be flown for the
maximum product of airframe efficiency (TAS/power) and engine efficiency (1/
SFC). As we have also just seen, airframe considerations determine that the
aircraft should be flown at a recommended range speed (RRS) which is about
10% higher than the minimum drag speed. To maintain that speed, a certain
amount of power must be used. lf maximum range is to be achieved, the
engine must be operated in such a way that this power required is produced
most efficiently, i.e. with the minimum specific fuel consumption (SFC).

Factors Affecting SFC


RPM and Manifold Pressure (MAP)

To obtain the power required, a number of combinations of rpm/manifold


pressure (MAP) may be used. The lowest SFC is obtained by using the lowest
rpm with the highest MAP (within allowable limits).
• Low rpm. Use of low rpm reduces friction losses and improves volumetric
efficiency. There will normally be a limit to the minimum usable, because
richer mixtures are required at very low rpm to prevent rough running, and
some engine-driven services (generator/alternator) may not operate
properly.
• High MAP. Maximum MAP for the rpm being used is limited by the cylinder
pressures above which a rich mixture must be used for cooling and the
prevention of detonation.

Mixture Strength

Lean mixtures, and the power settings which permit them, are essential to
achieving low SFC.

Altitude

The power required is produced more efficiently if the aircraft is at full throttle
height (FTH) for the setting being used. Basically, the reason for this is that the
engine 'breathes' better and the power losses through friction in the induction
and exhaust systems are reduced. Altitude also gives the advantage of colder
intake air which increases the temperature rise within the engine and improves
its thermal efficiency.

Temperature

Cold air at a given altitude improves SFC, since the power available can be
achieved at lower rpm, and the power required by the airframe reduces (TAS
reduced).

Carburettor Air Intake

Where the application of carburettor heat is necessary to prevent ice formation,


the SFC will deteriorate. High carburettor intake temperatures reduce the
density of the air intake, giving a richer mixture and reducing the thermal
efficiency. Ram air is also not generally available with carburettor heat selected,
and this lowers the FTH. For the same reason, the air filter (if fitted) should be
'out'.

15-4 Range and Endurance The Commercial Pilot Series


Summary

For piston-engined aircraft, the lowest SFC and best range is achieved by having:
• not more than maximum weak-mixture MAP set;
• minimum rpm;
• the engine(s) properly leaned;
• supercharger in low gear (if applicable);
• the aircraft at FTH; and
• carburettor air cold and air filter 'out'.

Note that when flying in lower temperatures, SFC is enhanced and specific air
range is improved.

Flying for Range - Practical Application


For many modern light aircraft, power required curves or a recommended range
speed are not provided in the Pilot's Operating Handbook (POH). To determine
the maximum range under prevailing conditions, it is necessary to use the
performance tables or graphs provided to:

• extract the fuel flows at various altitudes and true airspeeds; then
• apply head or tailwind components to find the corresponding
groundspeeds, which are then used with the fuel flows to find the best
specific ground range (SGR).

For some aircraft, theTAS and fuel flows obtained at various power settings and
pressure altitudes, will be given in a cruise performance table. With other
aircraft, these figures will have to be extracted from a graph like that shown at
Fig. 15-3 overleaf, which is representative of the cruise performance graph for a
relatively high-performance single-engine aircraft.

Note, with reference to Fig. 15-3:


• Fuel flow figures are given (in the boxes) for 81, 75, 65 and 55% maximum
continuous power (MCP). With US-sourced aircraft, the GPH (gallons per
hour) fuel flows stated will be in US gallons per hour.
• Two sets of fuel flows are stated-for 'best power' and 'best economy'.
These refer to two different leaning techniques which will be described
elsewhere in the POH. As you would expect, the technique used for range
flying is 'best economy' as this provides a leaner mixture and lower fuel
flows.
• The various power settings (MAP and rpm) to achieve the desired
percentage power at different altitudes are given in a separate table in the
POH. As has been explained previously, the best power setting to use is that
with the lowest rpm/highest MAP within the of recommended range of
power settings.

Example I. Using the graph in Fig. 17-3, analyse the SAR and SGR at 55%, 65%
and 75% MCP, for a gross weight of I ,633 kg, under !SA conditions, and using
best economy mixture with headwind components of:
sea level .............. 0 kt, 5,000 ft ................. 0 kt
I 0,000 ft ............... 5 kt HEAD 15,000 ft ............... 20 kt HEAD

Principles of Flight Range and Endurance 15-5


OUTSIDE AIR TEMP. - 'C TRUE AIRSPEED -KNOTS

Fig. 15-3. Sample cruise performance table. (Not to be used operationally.)

1. Extract the TAS and fuel flow from the cruise performance table for each
altitude and power setting. (It will be best to make up a table for each power
setting as shown below. As in our example standard conditions apply, the TAS
at each altitude is found by following the 'Std Temp' line up until it intersects the
required pressure altitude line, then across to the appropriate 'best economy'
dashed line, then vertically down to read off the TAS which will be achieved.
The fuel flow at the different percentage power settings is given in the 'best
economy' box at the top right).

TAS
SAR= = anm/USG: and
fuel flow
GS
SGR= = gnm/USG.
fuel flow

55% MCP

Altitude TAS Fuel Flow SAR (anm/ Wind GS SGR (gnm/


USG) Component USG)
SL 122 13·8 8·84 0 122 8·84

5,000 126 13·8 9·13 0 126 9·13

10,000 130 13·8 9·42 -5 125 9·06

15,000 135 13·8 9·78 -20 115 8·33

15-6 Range and Endurance The Commercial Pilot Series


2. Work out the SAR and SGR as follows and list in the table:
Note:
• In still air, best SAR with 55% MCP is obtained at 15,000 ft (9·78 anm/USG).
• In the given wind conditions, best SGR at 55% MCP is obtained at 5,000 ft
(9·13 gnm/USG).

65% MCP

Altitude TAS Fuel Flow SAR (anm/ Wind GS SGR (gnm/


USG) Component USG)
SL 138 15·7 8·79 0 138 8·79

5,000 143 15·7 9·11 0 143 9·11

10,000 148 15·7 9·42 -5 143 9-11

15,000 !53 15·7 9·74 -20 133 8·47

Note:
• In still air, best SAR with 65% MCP is obtained at 15,000 ft (9·74 anm/USG).
• In the given wind conditions, best SGR at 65% MCP is obtained at either
5,000 or I 0,000 ft (9·11 gnm/USG).

75%MCP
Altitude TAS Fuel Flow SAR (anm/ Wind GS SGR (gnm/
USG) Component USG)
SL 146 17·4 8·39 0 146 8·39

5,000 !53 17'4 8·55 0 !53 8·55

10,000 160 17'4 9·19 -5 !55 8·90

15,000 166 17'4 9·54 -20 146 8·39

Note:
• In still air, best SAR with 75% MCP is obtained at 15,000 ft (9·54 anm/USG).
• In the given wind conditions, best SGR at 75% MCP is obtained at I 0,000 ft
(8·90 gnm/USG).
By inspection of all three of the tables we have assembled:
• For this aircraft, the best SAR at any particular level is with 55% MCP set.
Hence, if as is likely, the aircraft is limited to 10,000 ft, the best SAR
obtainable is 9·42 anm/USG with power set for 55% and leaned for best
economy.
• In this example, with the given wind components, the best SGR (9·13 gnm/
USG) will be achieved with the same power setting and leaning procedure,
but at 5, 000 ft.
If you are committed to plan a flight which is likely to be at or close to the
aircraft's maximum range, there will be no alternative to conducting an analysis
along the lines described above. However, if light winds are forecast, only the
higher altitudes and lower cruise power settings need be considered and
compared. Some aircraft Operating Handbooks may contain figures for 45%
MCP as well as 55% MCP in which case, the SGR (gnm/USG) should be
compared for both settings.

Principles of Flight Range and Endurance 15-7


If there are stronger headwinds aloft, the analysis will need to include the lower
altitudes to determine whether, as in Example I above, an advantage is gained
by remaining at a lower altitude. In yet stronger headwinds, as demonstrated in
Example 2 below, a better SGR may be obtained at a higher TAS and higher
power setting.

Example 2 Consider the effect on the SGR for the aircraft in Example I, if the
headwind component at 5,000 ft was -50 kt.

Power TAS Fuel Flow Wind GS SGR (gnrn/


Component USG)
55% 126 13·8 -50 76 5·50

65% 143 15·7 -50 93 5·92

Obviously in these stronger headwind conditions, the best SGR for this aircraft at
5,000 ft is obtained at the higher speed with 65o/o power set, rather than the 55o/o
MCP which gives the best range in lighter headwind conditions.

Summary

To achieve best range:


• Fly at the power setting/altitude(TAS at which an analysis indicates the best
gnm/USG is obtained under the prevailing conditions. (Ideally, the aircraft
will be at 'full throttle height' at this power setting and altitude, but this is not
always possible).
• Use the highest MAP/lowest rpm combination within the recommended
range to achieve the desired percentage power.
• Use lean mixture (or 'best economy' leaning procedure).
• If possible, avoid continued use of carburettor heat, and have 'ram air'
selected.
• Minimize drag-flap and undercarriage up, cowl flaps closed.
• Lowest possible gross weight.

Flying for Endurance


Flying for maximum endurance involves remaining in the air for the greatest
amount of time for the least amount of fuel used. Note that it is time in the air
which counts in flying for endurance-not distance, as in flying for range. There
are few occasions when pilots are faced with having to fly for endurance but it
can occur, perhaps for air traffic control reasons, or when waiting for the arrival
of a forecast improvement in aerodrome weather.

The speed for maximum endurance is the TAS which coincides with the bottom
of the power required curve as shown in Fig. 15-4. (Note the difference between
this position on the CU!Ve and that shown for best range speed in Fig. 15-1.) This
is the speed at which the minimum amount of power (and the lowest fuel
consumption) is required to maintain level flight.

15-8 Range and Endurance The Commercial Pilot Series


power
required for
level flight

Fig. 15-4. Maximum endurance


(piston engine) is obtained at the
minimum power speed. minimum
power
I
I
v
0 t TAS
TAS for
maximum endurance

Effect of Altitude

The minimum power speed will coincide with a given lAS for a particular
weight. At higher altitude, drag will be the same at that lAS but the TAS will be
increased. Since power = drag X TAS, power required and gross fuel
consumption (GFC) to maintain the endurance speed increases with altitude.
Therefore, best endurance is achieved at the lowest safe altitude.

Effect of Weight

Minimum power speed increases with increased weight, and the drag at the
higher speed also increases. Therefore, power required and GFC both increase
because of the higher drag and TAS, and endurance decreases with increased
weight.

Engine Considerations
For maximum endurance, the engine must be operated with minimum gross
fuel consumption (GFC)-minimum fuel flow. This is achieved at the lowest
permitted RPM in the lean range, with MAP adjusted to maintain the 'minimum
power' speed. As with flying for range, correct leaning of the mixture has a
significant effect on the endurance achieved.

Practical Application
In practice, the best lAS to fly for endurance can be found with a little judicious
experimentation.

• Fly at about the recommended gliding speed and, with small adjustments to
power, determine the lowest power setting which will comfortably hold the
aircraft in level flight.

o Use the lowest permitted rpm for the lean range, which will give smooth
running, and enable the generator/alternator to charge. Adjust MAP to
maintain the selected speed.

o In turbulent conditions, or for manoeuvre, fly at a slightly higher speed (e.g.


I 0% higher) to avoid having to apply large increases in power to overcome
the effects of gusts/increased drag.

o Ensure that the mixture is correctly leaned.

Principles of Flight Range and Endurance 15-9


• Fly at the lowest practical altitude, but if you have the luxury of high altitude,
descend slowly, power on at the endurance speed, until the lower altitude is
reached. (The aircraft remains airborne during the time in descent with less
power than that needed for level flight, thus increasing the endurance.)

Summary-Piston-Engine Range and Endurance Speeds.

As shown in Fig. 15-5:

• For maximum range, the aircraft is flown at the minimum drag speed.
Although more power is used than when flying for endurance, this is the
speed which provides the highest ratio of TAS obtained for power applied
(and therefore, the greatest distance obtained for the amount of fuel used).

• For maximum endurance, the aircraft is flown at the minimum power


speed. Although the drag is higher than when flying for range, the least
amount of power is needed to maintain level flight and, therefore, gross fuel
consumption is lowest (resulting in the greatest amount of time which the
aircraft can remain airborne). ·

For a reminder on the difference between the minimum drag speed and the
minimum power speed, refer back to Chapter 9, page 9-9, and Fig. 9-11.

power
required for
level fiight

0
best J t . -~est
endurance ·range
TAS

Fig. 15-5. Maximum range and maximum endurance speeds compared.

15-10 Range and Endurance The Commercial Pilot Series


Review 15
I. Maximum range is achieved when the greatest .............................. is flown
for the least amount of ...................... used.

2. Specific Air Range (SAR) is defined as the .............. distance flown


per ................................................. of fuel used (e.g. anm/USG).

3. Gross fuel consumption (GFC) is the amount of fuel (e.g. USG or litres) used
per ......................... .

4. For maximum SAR, the aircraft must be flown for the maximum product
of ................................... efficiency and ................................ efficiency.

5. The best airframe efficiency (and therefore best SAR) is obtained at the lAS
for minimum ................ ./maximum ................................ ratio.

6. Increased altitude (affects/does not affect) the lAS for best SAR.

7. A reduction in weight will (improve/reduce) SAR.

8. Specific ground range (SGR) is the distance flown per unit


quantity of fuel used.

9. In a strong headwind, the optimum speed for best SGR (increases/


decreases) and in a strong tailwind it (increases/decreases).

I 0. Piston engine considerations determine that for best range:


(a) (high/low) rpm and (high/low) MAP should be used;
(b) the mixture should be (rich/lean) with, if possible, carb. air in (hot/
cold) and air filter out; and
(c) the aircraft should be flown (below/at) the full throttle height for the
power setting used.

11. Flying for endurance involves .remaining airborne for the longest amount
of .................. for the least amount of ................. used.

12. The gross fuel consumption at the best endurance lAS (increases/
decreases) with altitude. Th'erefore, the best altitude for endurance is as
(high/low) as practicable.

13. Increased weight (reduces/has no effect on) endurance.

14. To obtain maximum endurance/minimum GFC, the:


(a) rpm should be the (highest/lowest) permitted in the lean range;
(b) ....... should be adjusted to the minimum necessmy to maintain the
required speed; and
(c) ................................ should be properly leaned.

15. For range, the aircraft is flown at the minimum ....................... speed. For
endurance, the aircraft is flown at the minimum ............................... speed.

Principles of Flight Range and Endurance 15-11


15-12 Range and Endurance The Commercial Pilot Series
High Speed Flight
Introduction

When an aircraft moves through the air, it creates a series of pressure waves
which we know as sound waves. These pressure waves radiate out in all
directions from the aircraft at the speed of sound-which, under standard sea
level conditions, is a little over 660 knots. With aircraft speeds well below the
speed of sound, the faster-moving pressure (or sound) waves travel well ahead
of the aircraft and forewarn of its approach. We can hear the aircraft coming
but, more importantly, the pressure wave 'signals' give the air time to begin
moving aside to permit the passage of the aircraft with the least disturbance.

Fig. 16-1. In subsonic flight, pressure waves


travel ahead of the aircraft to warn of its
approach.

This behaviour of the air in reacting to the pressure wave signals sent out ahead
of an aircraft can be readily demonstrated using smoke streamers in a wind
tunnel. If, for example, an aerofoil with a flap is placed in a relatively slow speed
airstream, the amount of upwash ahead of the aerofoil can be seen to increase
markedly at the same moment as the flap is lowered. The reason is that the
increase in pressure differential above and below the aerofoil has been reflected
in a stronger pressure wave signal being sent forward, and the air has reacted by
moving so that more of it will pass through the lower pressure area-along the
path of least resistance.

------------------------
-------------------- --------

~~~~~~~~~~~~~~~~~~~~~~~~~~2~
--------------~~~
-------- _____ ,.____._.
-- _____.::-;-=====:~ ,,,......
---------
-, ..::::::.::::::-..._'
....... .::::::.

Fig. 16-2. In slow-speed flight, the pressure wave 'signals' give the air time to react to the
approach of the aircraft.

Given sufficient warning by these pressure waves of the approach of a solid


body, the air is in effect able to begin moving 'out of the way' and, to a large
extent, avoid being compressed by the body as it passes. This characteristic
enables us to consider air as being 'incompressible' when dealing with flight at
speeds which are relatively low by comparison with the speed of sound. The
errors which arise from mal<ing this assumption are negligible and it greatly
simplifies the treatment and handling of the subject of slow-speed flight.

Principles of Flight High Speed Flight 16-1


At speeds above about 250 knots/0·4 times the speed of sound, we can no longer
afford to ignore the effects of compressibility. As the speed of an aircraft is
increased, the distance which the pressure waves are able to penetrate ahead of
it is decreased. The 'free' air ahead of the aircraft receives less and less warning
of its approach; is less able to 'move out of the way'; and hence becomes more
compressed by the aircraft as it passes. This compression (and subsequent
expansion) of the airflow passing the aircraft means that, in addition to the
changes in speed and pressure of the airflow, changes in the density and
temperature also become important.

Initially, the onset of these compressibility effects with the increased speed is
gradual. However, as the aircraft more closely approaches the speed of sound, it
begins to 'catch up' with its own pressure wave signals through the airflow
ahead of it. When it does, the pressure waves pile up on top of one another (or
coalesce) and shock waves are formed. The flow pattern around the aircraft
now becomes dominated by these shockwaves, through which sudden changes
in the speed, pressure, density and temperature of the airflow occur.

Fig. 16-3. Approaching the speed of sound,


the aircraft 'catches up' with is own pressure
wave signals.

Flow Regimes
The airflow and pressure patterns around an aircraft thus vary, depending on
how fast the aircraft is travelling in comparison with the local speed of sound.
These different flow patterns can be characterized under different flow regimes
(or speed ranges) as follows:
• Low-speed subsonic flow. This is the flow below about 250 knots where
the air behaves for all practical purposes as if it were incompressible. It
produces the streamline and pressure patterns with which we have become
familiar in the previous chapters.
• High-speed subsonic flow. This is the range in which the speed is high
enough to give significant compressibility effects, but in which nowhere
around the aircraft does the local airflow exceed the local speed of sound.
While fundamentally remaining the same as for low-speed subsonic flow,
the streamline and pressure patterns undergo subtle changes which we
discuss in more detail later.
• Transonic flow. At faster speeds, as the airflow around the aircraft is
accelerated, some of it will reach or exceed the speed of sound. The
transonic flow regime is characterized by having a mixture of subsonic and
supersonic airflow, and by the formation and movement of shock waves
(explained in more detail shortly). Transonic flow tends to be unpredictable
and changeable and the coefficients of lift and drag can vary considerably.
• Supersonic flow. In fully developed supersonic flow, the local speed of the
flow is everywhere greater than the speed of sound. Although shock waves
are still present, their position becomes fixed and the airflow around the
aircraft again 'settles down' and becomes more predictable.

16-2 High Speed Flight The Commercial Pilot Series


NOTE: In the context of high speed flight, 'sonic' means 'at the speed of sound'.
Hence it follows that 'subsonic' applies to speeds below the speed of sound and
'supersonic' relates to speeds above the speed of sound. The term 'transonic'
literally means 'across the speed of sound' but, in high-speed aerodynamics, it
applies to that speed band where there is a mixture of subsonic and supersonic
flow around the aircraft.

In this study of high speed flight, we will be concentrating on the high speed
subsonic and transonic regimes. These are the regimes that concern modern
commercial jet transport aircraft which have cruising speeds which 'butt into'
the early part of the transonic region. We will not be much concerned with
supersonic flight, or for that matter, with what is termed hypersonic flight at
again higher speeds.

It is clear frorn the foregoing that the relationship of the speed of the airflow
around the aircraft to the local speed of sound is fundamental to determining the
characteristics of the flow itself. In this study of high speed flight we therefore
turn first to look more closely at the nature of sound waves and the speed at
which they travel.

Sound Waves
A sound-wave is a very small pressure wave (or pressure disturbance) which
travels through air at a definite speed. A sound source is anything which causes
air to become slightly compressed-e.g. a firecracker; a tuning fork, a slow-
moving aircraft. Air is resilient (springy), and once compressed, it quickly
'springs back' and expands. This expansion causes the adjacent air to become
compressed and it also springs back and expands .... and so the process goes on.
Successive waves of slightly compressed (higher pressure) air travel out in all
directions in 'domino' fashion from the source at the speed of sound.
compression compression

expansion expansion

successive compressions/expansions travel at the speed of sound

Fig. 16-4. Sound (pressure) waves are transmitted through intermolecular collision.

It should be noted that in the transmission of sound waves, air does not become
physically displaced. Its molecules vibrate about a mean position and the
successive compressions and rarefactions of the wave are transmitted through
intermolecular collision. The size of a sound wave is measured by its amplitude,
i.e., the maximum change in pressure from static pressure. An average sound
wave will have an amplitude of about one millionth of an atmosphere and so by
any measure, sound waves are very small pressure disturbances.

In free air, sound waves attenuate (weaken) with distance from their source and
eventually disappear as the energy in the wave becomes spread over an ever-
increasing surface.

Principles of Flight High Speed Flight 16-3


The Speed of Sound
The speed at which a sound wave travels, depends only on the temperature
of the air in which it is travelling. As just explained, sound waves are
transmitted through intermolecular collision. The temperature of any substance
is a measure of the speed at which molecules in that substance vibrate and
move randomly about. With increased air temperature, the air molecules
become more energetic and sound waves are transmitted by adjacent
molecules at a faster rate. Conversely, in colder air, the molecules become
sluggish and sound waves are transmitted at a slower rate.

Although not normally required for practical flight, if necessary, the speed of
sound can be calculated by the formula:
a= 39,.IT
where a = speed of sound
T =absolute temperature of the air ("C + 273)

Fig. 16-5 shows the variation in the speed of sound in the International Standard
Atmosphere. It is emphasised that the reduction in the speed of sound with
increasing altitude is solely due to the reduction in temperature. Changes in
pressure and density have no effect provided the temperature remains constant.

Feet oc Knots

60 000 -56·5 573·8


50 000 -56·5 573·8
40 000 -56·5 573·8
35 000 -54-3 57~-*-
30 000 -44·4 589·6
25 000 -34·5 602·2
20 000 -24·6 614·6
15 000 -14-7 626·7
10 000 -4-8 638·6
5 000 5·1 650·3
Sea Level 15·0 661·7

Fig. 16-5. Variation of the speed of sound with altitude

Mach Number

The changes to the airflow and pressure patterns which occur as the TAS of an
aircraft approaches the speed of sound, have a significant effect on its
performance, manoeuvrability, stability and control. The speed of sound varies
with temperature-widely, as we have just seen from Fig. 16-5-and therefore
the TAS at which these changes occur will also vmy widely with temperature.

There is , therefore, a need to know what the aircraft-speed is in relation to the


speed of sound in that part of the atmosphere in which it is flying. Such a speed
measurement system was devised over a century ago by the Austrian physicist
Ernst Mach and from his name comes the term Mach number which is the ratio
of:
speed of the aircraft- or of an airflow (TAS)
local speed of sound (a)

Mach number is designated by the letter M and, since it represents a ratio of one
speed to another, it has no units.

16-4 High Speed Flight The Commercial Pilot Series


If an aircraft is flying at the local speed of sound (a), its TAS (V) will be equal to
a. The ratio of V/a is one and the aircraft will thus be flying at Mach I. An aircraft
flying at half the speed of sound has a Mach number of 0·5 and one travelling at
twice the speed of sound would have a Mach number of2.

For Mach numbers less than one, it is usual to drop the zero and express it as
"point seven Mach" for 0· 7 Mach. Throughout this chapter, for clarity, Mach
numbers will be written as follows; 0·7M and I ·2M.

Although we may initially think that an aircraft has only one Mach number at any
given time-i.e. the ratio of its TAS at that time to the local speed of sound-it is
important to remember that the speed of the airflow around the aircraft is not
the same at every point. For all the different speeds of flow around an aircraft
(flows with different local true airspeeds if you like), there will be different Mach
numbers. Consequently, there is a need to define different types of Mach
number.

Free-stream Mach Number (Mfs)

Mfs is the Mach number of the flow past the aircraft, but sufficiently remote to be
unaffected by it.
Mfs = (TAS)
local speed of sound (a)
Mfs is of course the Mach number of the aircraft as a whole and is also
sometimes called the flight Mach number. If the usually small instrument errors
are ignored, Mfs is the true Mach number of an aircraft as indicated on the
Machmeter.

Local Mach Number {ML)

At any given Mfs, the flow around an aircraft will be accelerated in some places,
slowed down in others. The speed of sound may also change in various places
because of the temperature changes which occur through compression or
expansion. Hence:
speed of flow at a point
ML =
speed of sound at the same point
ML may be higher than, the same as, or lower than Mfs.

Critical Mach Number (Merit)

As Mfs increases so do some of the local Mach numbers. That Mfs at which any
ML has reached 1·0 is called the critical Mach number. Since there is an
increase in the acceleration of the flow over the wing, Merit will occur at a lower
Mfs if the angle of attack is increased. Merit marks the bounda1y between the
subsonic and transonic speed ranges.

Detachment Mach Number (Mdet)

Mdet is that Mfs at which the bow shockwave (discussed later) can for practical
purposes be considered attached to the leading edge. Above Mdet there is
virtually no movement of the shockwaves and all local flows are supersonic
(except in the lowest layers of the boundary layer and perhaps in small areas in
front of the leading edges). Mdet marks the boundary between the transonic
and supersonic flow regimes.

Principles of Flight High Speed Flight 16-5


Critical Drag Rise Mach Number (Mcdr)

Mcdr is that Mfs above which, due to the formation of shockwaves, there is a
significant rise in drag. Mcdr will usually occur at a higher flight speed than Merit
and before a Mfs of 1.0 is reached.

Speed Ranges
We have previously described in broad terms the characteristics of the various
flow regimes (subsonic, transonic and supersonic). We are now in a position to
define these speed ranges more specifically in terms of Mfs, as shown in
Fig. 16-6.

all local Mach nos. <1·0 some local Mach nos. > 1·0 all local
some< 1·0 Mach nos.> 1·0

compressible flow

low high

I
0·4 Merit 1-0 Mdet

Mfs (fiight Mach no.)--~

Fig. 16-6. Speed ranges related to free-stream Mach number.

Subsonic

The subsonic speed range starts at zero and has its upper limit at Merit.
Remember that Merit is the Mfs at which the highest local Mach number reaches
unity. The subsonic speed range contains no flow whatsoever that is at or above
the speed of sound. The subsonic range is subdivided at about 0·4M as
explained previously into the low range, where the effects of compressibility are
negligible; and the high range where compressibility effects become
increasingly significant.
Transonic

The transonic range begins at Merit and extends up to Mdet. Between these
Mach numbers, shock waves form and move about on the aircraft and there is a
mixture of subsonic and supersonic local flows around the wings. In very
general terms (and it depends on aircraft design and angle of attack) the
transonic range extends from an Mfs of 0· 75M to about !·2M.
Supersonic

Mdet, which marks the beginning of the supersonic range, is the point where all
of the main local flows become supersonic; the bow shockwave is considered
attached to the leading edge; and the shockwaves become more or less fixed in
position.

16-6 High Speed Flight The Commercial Pilot Series


Formation of Shockwaves
Pressure Waves from a Moving Source
Source Moving at Subsonic Speed

Fig. I 6-7 shows the pattern of pressure waves from a source moving at subsonic
speed. Each circle represents the position, at a given instant in time, of
subsequent waves which originated at a regular time interval. The numbers and
the marks represent the position of the source at each of those time intervals
starting with 0 (at this instant) and ·I, -2, -3, representing its position I, 2, 3 time
intervals ago .... and so on.

-2 ·3 -4
v-+

a = speed of sound
V=TAS

Fig. 16-7. Pressure wave pattern from a source moving at subsonic speed.

The distance travelled at source speed (V) is less than the distance radiated by
the pressure wave at the speed of sound (a), hence the waves maintain their
separation although they are closer together ahead of the source than behind it.
The pattern shown in Fig. 16-7 is representative of a source moving at high
subsonic speed where the closeness of the waves is indicative of the short
warning time given of the approach of the source.

Source Moving at Sonic Speed

The pattern produced when the source is moving at sonic speed is shown in Fig.
I 6-8. In this case, the speed (V) of the source is the same as the speed of
propagation of the waves (a). They are thus unable to penetrate the air ahead of
the source, and they 'bunch up'-or coalesce-to form a Mach wave (or Mach
line). This Mach wave is a formed limit to the influence of the source-the air
ahead of it receives no warning of the approach of the source. As the Mach
wave is at right angles to the direction of movement of the source the wave is
called a normal Mach wave.

Principles of Flight High Speed Flight 16-7


Fig. 16-8. Pressure wave pattern from source moving at sonic speed.

Source Moving at Supersonic Speed

At supersonic speeds yet another pattern is produced. With the speed of the
source (V) greater than the speed of propagation of the pressure waves, the
source in effect 'leaves them behind' within a three-dimensional cone shape.
The bounda1y of the cone is formed of an oblique Mach wave. It can be seen
from Fig. 16-9 that the angle which the oblique Mach wave makes with the flight
path (called the Mach angle (rn)) is given by:
sin rn = a/V, but as Mach no. (M) = VIa , sin rn = 1/M
In other words, the Mach angle is inversely proportional to the Mach number of
the source, the faster the source is travelling, the more angled back will be the
oblique Mach wave.

0 -2 -3 -4

a = speed of sound
V=TAS

Fig. 16-9. Pressure wave pattern from source moving at supersonic speed.

16-8 High Speed Flight The Commercial Pilot Series


Shockwaves
The simple pressure patterns illustrated in Figs. 16-7 to 16-9, are those which
would result from an infinitely small pressure disturbance arising from a moving
point source. The pressure change across the wave would be infinitely small
and the airflow passing through the wave would be infinitesmally affected by it.
A Mach wave (or line) is thus a line along which an infinitely small pressure
disturbance is felt. The significance of Mach lines will be appreciated when
expansions in supersonic flow are covered later.

The formation of shockwaves in the flow around an aircraft is more complex,


but the principle of their formation remains the same as for the simple Mach
waves just discussed. By comparison with a Mach wave, a shockwave is a
compression wave with a much greater amplitude. Shockwaves are very thin
regions (much less than 1mm across) and can be imagined as a kind of front
formed in the supersonic airflow around an aircraft, through which sudden
changes to the speed, pressure, density and temperature occur. Depending on
the speed of the aircraft and its shape, like Mach waves, two kinds of shockwave
occur-normal and oblique. The physical changes to the airflow which occur
across them are summarized in Fig. 16-10. Mach number, TAS, pressure, density
and temperature are represented respectively by the symbols M, V, p, p,and T,
while subscripts 1 and 2 represent the flow before and after the shockwaves.
normal shock oblique shock

M1 >1 M2 <1 M1 >1 M2< M1

8
V1

p1
~
V2

p2
•t 0 8
V1

p1
V2t-

p2 t 0
pi p2 t pi p2 t
Ti T2 t Ti T2 t

Fig. 16-10. Changes in the airflow which occur through a shockwave.

NOTES:

I. Both types of shockwave are similar. Before either can form, the airflow
must have a Mach number of at least 1·0. Through each type, Mach number and
TAS of the flow is reduced, while the pressure, density and temperature all rise-
in other words, a sudden compression of the air has taken place across the
shockwave.

2. The only differences are:


• the normal shockwave does not change the direction of the flow, the
oblique shock does;
• the speed of the flow is always reduced to less than Mach 1·0 across a
normal shock, whereas across the oblique shock, the Mach number may or
may not be reduced to below 1·0 depending on the initial speed of the flow
and the strength and angle of the shockwave.

3. The physical properties of the airflow do not necessarily remain constant


after passing · through the shockwave. In particular, if the conditions are
favourable, the airflow can expand and accelerate rapidly resulting in a
reduction in pressure, density and temperature.

Principles of Flight High Speed Flight 16-9


Shockwaves form whenever a supersonic flow is required to 'turn into itself or
converge at what is called a compressive corner. To explain these terms more
fully, refer to Fig. 16-11, in which a supersonic flow around the upper half of a
wing section is shown.
A------------------------------~A1

Fig. 16-11. Compressive corners in the supersonic flow above a wing section.

A -AI represents the streamline above the wing which is sufficiently remote to
be unaffected by it. B - BI is one of the streamlines in the flow nearer the wing
which is obliged to change direction to let it pass. There are two points on B - BI
at which the flow must turn towards the remote flow A- AI. These are shown
as I and 2 in the diagram where the flow must turn at the leading edge and
again toward the trailing edge near the separation point where the airflow must
resume its freestream path. By comparison with the flow over the top of the
wing (along the dashed line) where expansion can take place, compression
must occur at points I and 2-the compressive corners in the flow-and this is
indeed where shockwaves form in fully supersonic flow. As the flow situation is
mirrored below the wing, shockwaves form in the lower flow also (Fig. 16-12).

bow
shockwave tail

~============::::~~-:shockwave
- wake

Fig. 16-12. Shockwaves in fully supersonic flow.

The Bow Shockwave


Consider a wing section which is travelling at Mach 1·0 (i.e. Mfs = 1·0). As
shown in Fig. 16-13, ahead of the leading edge there will be an area around the
stagnation point where the air (on meeting the wing) becomes highly
compressed and accordingly has an increased temperature. This area can be
imagined as a bubble of compressed, dense air with its highest temperature at
the stagnation point but reducing over a certain distance to ambient
temperature.
bow shockwave forms
ahead of leading edge
pressure waves propagate forward
initially faster than Mfs 1·0

Mfs = 1·0 - - - - - •

'bubble' of compressed air


at higher temperature

Fig. 16-13. Formation of the bow shockwave.

16-10 High Speed Flight The Commercial Pilot Series


As the speed of sound is proportional to temperature, the pressure waves from
the leading edge propagate ahead initially at a speed which is higher than the
speed at which the wing is travelling (Mfs = 1·0). However, as the temperature
drops to ambient, they slow up to the same speed as the oncoming air. They
therefore coalesce at the edge of the temperature 'bubble' and a normal
shockwave is formed which 'stands off ahead of the leading edge.

As the flight speed is increased above Mach 1·0 the higher speed of wave
propagation within the temperature bubble is partially offset by the increased
speed of the aircraft. The stand off distance of the bow shock is accordingly
reduced and it begins to bend back and become more oblique at the top and
bottom. When the Mfs becomes equal to the speed of propagation, the wave
will in theory attach itself to the leading edge. Remember that the speed at
which this occurs is defined as Mdet.

As shown in Fig. 16-14, the flow behind the normal part of the bow shockwave is
subsonic and unchanged in direction by the shockwave itself. However, once it
is reduced to subsonic speed by the shockwave, the air is able to negotiate the
leading edge in the usual way as a subsonic flow. When the deflection angle is
reduced, it is then free to accelerate rapidly once again to supersonic speed.
Unless the wing has a leading edge which is 'razor sharp' this means that in
practice, there will always be a small area of subsonic flow ahead of the leading
edge regardless of the flight speed.

supersonic
M1 > 1·0
Fig. 16-14. The bow shockwave
and streamlines at higher Mach
numbers

The ability of the bow wave to attach is basically determined by the shape of the
aerofoil and the amount of deflection required to split the airflow to enable it to
pass above and below the wing. The blunter the leading edge, the greater the
compression and temperature rise and consequently the higher the increase in
the local speed of sound. The greater the increase in the local speed of sound,
the more difficult it is for the bow wave to attach itself to the leading edge.

Most aerofoil shapes do not allow the bow wave to attach, and in reality it will
stand off slightly in front of a small leading edge area of subsonic flow at all
speeds. To encourage bow wave attachment-desirable to achieve stable flow
and reduce supersonic drag-a sharp leading-edge radius is used for supersonic
aircraft. Once attached (or as near as it is going to get) an increase of Mach
number into the supersonic band results in the bow wave becoming more
oblique.

Wing Shockwaves
Wing shockwaves form first above and then below the wing, well before the
bow wave appears. After forming at about mid section, the wing shockwaves
move at different rates toward the trailing edge, eventually joining there to form
the tail shockwave.

Principles of Flight High Speed Flight 16-11


Wing shockwaves form for the same reason as the bow wave. As Merit is
passed, a bubble of supersonic flow appears above, and then later below, the
wing. The pressure waves emanating from the 'compressive corner' located
near the separation point, travel forward along the wing. On meeting the
supersonic flow, they are unable to proceed any further fmward and coalesce to
form a shockwave.

At the base of the shockwave is the boundary layer, part of which is still
subsonic. This slower moving flow in the boundary layer enables the pressure
rise at the shockwave to be communicated fmward, thickening the boundary
layer and creating a further compressive corner. Small oblique shockwaves
emanating from this compressive corner merge with the main shockwave
resulting in a lambda foot, named for its resemblance to the Greek letter
lambda (A.). This interaction between the shockwave and boundary layer results
in boundary layer separation from the base of the shockwave leading to the
development of a thicker turbulent wake.

Fig. 16-15 shows a diagram of upper and lower wing shockwaves at a


representative subsonic flight Mach number (about Mfs 0·85 for that wing
section). At that point the upper shockwave tends to become anchored for a
time at about 70o/o chord. As there is less acceleration below the lower surface
and less flow breakaway, the lower shockwave forms later, is weaker than the
upper and, once formed, moves more quickly back toward the leading edge
lambda foot
normal
supersonic
shockwave

Fig. 16-15. Typical wing


Mfs=0·85-
shockwaves.

normal
shockwave
Depending on wing section and angle of attack, wing shockwaves will form well
below an Mfs of 1·0 and they are common above an Mfs of 0·85. The formation
of wing shockwaves can be delayed by using a thinner wing section which
reduces the flow acceleration while highly cambered sections encourage their
formation at lower speeds. Any increase in angle of attack, e.g. during
manoeuvring, will result in increased flow acceleration (for a given Mfs) and
result in the formation of shockwaves at even lower Mfs values.

Expansion Waves
Whenever a supersonic flow is able to turn away from itself-or diverge-it will
expand and accelerate. Refer back to Fig. 16-11. Note how, between points I
and 2, the direction of the streamline is gradually and smoothly turning away
from the remote airflow A- AI. In this area above the wing, expansion is taking
place. In direct contrast to compression in a supersonic flow-which takes
place suddenly (almost explosively) across a shockwave-expansion occurs
through an infinite number of Mach lines, where the changes to the flow are
infinitely small. As shown in Fig. 16-16, a convex upper surface of a wing can be
considered as an infinite number of small steps called expansive corners.
Consider the two streamlines in the supersonic flow as they reach each corner.
Immediately the particles in the flow adjacent to the surface reach the corner,
they are 'given more room' and, as a result, the flow expands, accelerates and
the pressure decreases. This pressure disturbance is then felt along a Mach line
with an angle (m) which depends on the Mach number of the flow (the higher
the Mach number, the smaller the angle).

16-12 High Speed Flight The Commercial Pilot Series


M1 > 1·0

Fig. 16-16. Supersonic flow around expansive corners.

The streamline further away from the surface is unaware of the corner until it
reaches the Mach line originating at the corner. This Mach line represents a
'front' through which there is an infinitely small drop in pressure. The particles
in this streamline also then expand and accelerate. The expansion and
acceleration in the successive streamlines of the flow is such that it can conform
with the change in direction of the surface.

The process is repeated at each expansive corner. Note that as the Mach
number is increasing, the angle (m) of each successive Mach line is decreased
and they therefore fan out. This results in a smooth expansion and acceleration,
even around a relatively sharp corner, such as depicted in the lower diagram of
Fig. 16-16. The decrease in pressure around an expansive corner in supersonic
flow provides a favourable pressure gradient and allows an attached boundary
layer to be maintained. This is exactly the opposite to subsonic flow where a
sharp corner would result in an adverse pressure gradient, boundary layer
thickening and flow breakaway. Consequently, there is no objection to such
corners on aircraft designed primarily for supersonic flight.

The Nature of Supersonic Flow


From the foregoing, we can summarize the nature of supersonic flow as follows:

Where the flow is required to converge (or turn towards itself), compression
takes place. This compression occurs across a shockwave where suddenly, the
velocity of the flow is decreased, and its pressure, density and temperature
increased.

Where the flow is able to diverge (or turn away from itself), it expands. This
expansion of the flow takes place smoothly and gradually through expansion
waves (or fans) which are the direct antithesis of shockwaves. During an
expansion, the velocity of the airflow increases, and the pressure, density and
temperature decrease.

It can be seen that the characteristics of supersonic flow are completely different
from subsonic flow. Whereas the velocity of subsonic flow increases where it
converges (and the pressure reduces accordingly), in supersonic flow, the
velocity of the flow is reduced through the shockwaves caused by convergency,
and the pressure is suddenly increased. Conversely, in an area of divergency,

Principles of Flight High Speed Flight 16-13


subsonic flow slows down and its pressure rises, whereas supersonic flow
accelerates and its pressure reduces. The table at Fig. 16-1 7 summarizes and
compares the different characteristics of subsonic and supersonic flow.

~· w
Convergent Divergent

Subsonic vt Pt P't vt Pt P*t


Supersonic vt Pt Pt vt Pt Pt
thru shockwave

*NOTE: at low subsonic speeds, these changes are so small as to be negligible

Fig. 16-17. Comparison between subsonic and supersonic flow characteristics.

Fig. 16-18 illustrates the difference between subsonic 'incompressible' flow and
supersonic 'compressible' flow through a convergent/divergent duct. You will
be familiar with the top diagram showing the conditions for subsonic flow
through a venturi. The lower diagram shows an 'idealized' supersonic flow
through the same duct. 'Idealized' in the sense that such flow is dominated by
the formation and positioning of shockwaves which depends to a large extent on
the speed of the flow and the shape of the duct. [The design of ducts (diffusers
and nozzles) for various types of supersonic flow is complex, and beyond the
scope of this manual].
subsonic 'incompressible' flow
speed increasing speed decreasing
pressure reducing pressure increasing

--------
--------=-------- ---
-...:.:. ::.::
----------------
--------
M < 1.0
--------
------------------------
--------------------------
--------------------------
------
----------=---------------
constant density

supersonic 'compressible' flow


speed decreasing speed increasing
pressure increasing pressure decreasing
density increasing density decreasing

M > 1.0

compression through expansion through


shockwave(s) expansion waves

Fig. 16-18. Subsonic and supersonic flow through a convergent/divergent duct.

16-14 High Speed Flight The Commercial Pilot Series


The important thing to note from Figs. 16-17 and 16-18 is that the characteristics
of subsonic and supersonic flow are reversed insofar as convergency/divergency
are concerned:
• Where the streamlines converge, subsonic flow-which we take to be
incompressible-speeds up and the pressure reduces. In supersonic flow,
the opposite occurs-the flow slows down, the air compresses, and its
pressure rises.
• Where the streamlines diverge, subsonic flow slows down and its pressure
rises. In supersonic flow, the opposite occurs-the air speeds up through
expansion, and its pressure reduces.

Note that subsonic and supersonic flow are similar in one important aspect.
That is, that where the flow is accelerated the pressure reduces. Thus an
aerofoil which is able to produce a greater acceleration of the flow over its upper
surface will produce lift regardless of whether that flow is subsonic or
supersonic.

The Effects of Compressibility on Lift


As mentioned earlier, in the high-subsonic speed range, compressibility effects
bring about subtle changes to the streamline and pressure patterns around an
aerofoil. As a result, the coefficient of lift (Cd is increased for a constant angle of
attack. In the transonic range, the presence of shockwaves, either in forming or
moving, creates significant changes in the flow which varies the CL. In extreme
cases, particularly where the aerofoil is not one designed for high speed, these
changes can lead to stability and control problems. When the flow becomes
fully supersonic above Mdet, the position of the shockwaves becomes fixed, the
CL settles down and once again becomes predictable.

B
A
I

I
I
I
D
I
E
r---,--------+----- --T------~--

1
1
!+-----:: subsonic _ ___., c 1+---- supersonic
I I
I transonic ---+; I
I I

0-4 Merit Mdet

Fig. 16-19. Variation of CL with Mach number at a constant angle of attack.

Fig. 16-19 shows the variation of CL with Mach number at a constant angle of
attack for a subsonic aerofoil with a thickness/chord ratio of 12%. The changes
in lift in the transonic region (discussed in more detail shortly) are so large that
this section would not be suitable for transonic/supersonic flight. It does
however se1ve to demonstrate clearly the effects of compressibility on lift. These
effects are described in the following paragraphs with reference to the various
points marked on the above graph.

Principles of Flight High Speed Flight 16-15


Leading up to Point A (Fig. 16-20). Above 0·4 M, the effect of compressibility on lift becomes
significant. With less warning time of the approach of the aircraft, the streamline pattern is
changed at high-subsonic speeds to that shown in the diagram. The change in upwash
ahead of the wing is more abrupt, leading to an effective increase in angle of attack. In
addition, the streamlines above the wing are closer together. Both of these factors lead to an
increase in lift which (for a given angle of attack) is more than proportional to V2 ; i.e. there is
an increase in CL.

Fig. 16-20. Changes to the streamline pattern at high subsonic speed.

Point A - Point B (Fig. 16-21). Merit is reached at point A. Beyond Merit a bubble of
supersonic flow begins to expand above the upper surface. At this early stage in the
development of supersonic flow, only very weak shockwaves (or compression waves called
'whiskers') form inside this bubble and, as they have little effect in retarding the airflow over
the upper surface, the CL continues to increase for a while after Merit is passed.

supersonic 'whiskers'
'bubble'

subsonic subsonic

Merit + ______.....

Fig. 16-21. Past Merit, a bubble of supersonic flow begins to expand over the upper surface.

Point B (Fig. 16-22). The 'whiskers' are swept together to form the upper shockwave. As it
strengthens, the pressure gradient becomes more adverse: the boundaty layer thickens and
the separation point moves fon.Yard. As a result, the CL begins to decrease. This loss of lift
from point B-which is typical for an aerofoil of this type-is sometimes called the 'shock
stall' (explained in more detail shortly).

supersonic flow
moving back

r sep<>ration point moving forward

Mfs=O·S-+

Fig. 16-22. CL reduces once the top shockwave forms.

16-16 High Speed Flight The Commercial Pilot Series


Point C (Fig. 16-23). The upper shockwave has strengthened and become
anchored at about 70o/o chord. The lower shockwave has formed and moved
rapidly to the trailing edge. Under these conditions, there is high pressure above
and lower pressure below the rear part of the wing, and lift is lost as a result.
Consequently, the CL is substantially reduced by comparison with the basic
subsonic value.

Mfs=0·89-

Fig. 16-23. Flow conditions at point C.

Point D (Fig. 16-24). The top shockwave moves to the rear of the wing which
now means that most of the wing is bathed in supersonic flow. The size of the
wake is restored to approximately that which existed prior to the shockwaves
forming. With these improvements in the flow, the CL is also restored to slightly
above the basic subsonic value.

Mfs=0·98-

Fig. 16-24. Flow conditions at point D.

Point E (Fig. 16-25). The bow shockwave has now formed. As the flow behind
the bow wave (which is the flow over the wing producing lift) has had its energy
reduced by the shockwave, the CL is once again reduced.

supersonic supersonic

subsonic

Fig. 16-25. Flow conditions at poinf E.

Beyond Point E. All shockwaves become more oblique and the bow wave will
almost attach to the leading edge. Although the flow is stable and the CL
relatively steady, by about Mfs 1·4, it is reduced to about 70o/o of its basic
subsonic value for that angle of attack.

Principles of Flight High Speed Flight 16-17


The Shock Stall

Referring back to Fig. 16-19, note the significant peak in the CL Clllve which
occurs at a speed which is a little beyond Merit for this type of (basically
subsonic) aerofoil. As we have seen from Figs. 16-22 and 23, the reason for this
sudden drop in CL is the formation of the shockwave over the upper surface of
the wing. At the same time, the positioning of this 'top' shockwave over the
upper surface, causes the turbulent wake behind the wing to be much thicker
than it is at higher or lower speeds.
These two effects-a relatively sudden loss of lift accompanied by airframe
buffeting resulting from the thick turbulent wake-cause the aircraft to react in a
very similar manner to the low speed/high angle-of-attack stall, hence the term
'shock stall'. In transonic/supersonic aircraft, these effects of the shock stall can
virtually be eliminated by good design.

The Effects of Compressibility on Drag


The main effects of compressibility on drag arise from the formation of
shockwaves in the transonic region. This type of drag is called wave drag or
Mach drag and it is made up of the drag from two separate sources-energy
drag and boundary layer separation.

Energy Drag
As the airflow crosses a shockwave, energy is required to provide the
temperature rise, and this energy demand is placed on the aircraft in the form of
increased drag. This increase in drag is exerted on the aerofoil through the
rearward facing surfaces experiencing a greater reduction in pressure than the
forward facing, once the shockwaves have formed. Normal shockwaves
generate more wave drag because the temperature rise across them is
proportionally higher. The more oblique the shockwaves are, the less energy
they absorb, but because they become more extensive laterally and affect a
greater volume of air, the energy drag rises as Mfs increases.
Boundary Layer Separation
At certain stages of the development and movement of the wing shockwaves,
there is considerable flow separation-refer back again to Figs. 16-22 and 23. At
these points in the development of transonic flow, the thick turbulent wake
increases the drag considerably. As both shockwaves then move to the trailing
edge, the size of the wake reduces again, and so does the drag arising from
boundary layer separation.
Taken together, these two I
sources of drag-energy drag I
I
and boundary layer I
separation-modify the drag boundary layer
1
1parasite

curve as shown in Fig. 16-26. separation drag 1 drag


D I

energy drag
Fig. 16-26.
The effect of wave drag.

Mfs

16-18 High Speed Flight The Commercial Pilot Series


This change in the drag characteristics (called the transonic drag hump) is also
reflected in the curve of Co against Mach number.

Fig. 16-27.
Variation of Co with Mach number
at constant angle of attack
Mach
Merit Mcdr number

NOTES:

I. In the high subsonic range, there is little change in Co until Merit is reached.

2. At Merit, there is only a small increase in C0 . At slightly higher Mach number,


in concert with the formation of the upper shockwave, there is a significant
increase in C0 . This is the critical drag rise Mach number (Mcdr), sometimes
also called the drag divergence Mach number.

3. The peak in Co occurs at about M 1·0 where, although the contribution of


boundary layer separation is reducing, that of energy drag is increasing with the
formation of the bow wave above M I ·0.

4. Above Mdet, the Co settles at a value which is approximately 1·5 times the
low subsonic value at the same angle of attack.

Control at High Speed


Longitudinal Control
As most aircraft enter the transonic speed range they experience a nose-down
trim change (called 'Mach tuck') which is caused by two factors:

• The rearward movement of the centre of pressure. Before Merit is reached,


the CP will probably lie at about 20-25% chord (depending on wing section).
Through the transonic range, the CP moves rea1ward as (with the rearward
spread of supersonic flow) pressures beyond the point of maximum
thickness continue to decrease and the rear part of the wing makes an
increased contribution to lift. When the flow is fully supersonic, the CP will
have typically shifted to be at about 50% chord. This rearward movement of
the CP can cause a significant nose-down pitch.

• A reduction of downwash on the tailplane. As the wing shockwaves form


and strengthen, there will generally be increased flow separation,
particularly behind the upper shockwave. For most aircraft with a
conventional tailplane configuration, the flow separation has the effect of
reducing the downwash over the tailplane which effectively increases its
angle of attack, resulting in a nose-down pitch.

Principles of Flight High Speed Flight 16-19


If uncorrected, the nose-down pitching moment will cause increased
acceleration, the Mach number will increase and the aircraft will pitch even
more nose-down. Alternatively, if the pilot corrects by applying up elevator (to
hold the same nose attitude) this can easily lead to an inadvertent increase in
angle of attack-and an acceleration in the formation of the shockwaves and the
ensuing nose-down pitch.

This destabilizing nature of the nose-down trim change places a very real limit to
the Mach number to which an aircraft can be safely flown and is normally the
basis for establishing an aircraft's maximum operating Mach number (MMo).
This is the 'red line' Mach number on the combined airspeed/Mach indicator. In
some aircraft, where cruise at higher Mach numbers may be desired to take
advantage of favourable drag figures, MMo may be increased by installing a
'Mach trimmer' which is simply a Mach sensitive device which automatically
deflects the tailplane or elevator slightly more than is needed to counter the
nose-down pitch. The reason for the 'over-deflection' is so that positive
longitudinal stability is maintained and the aircraft must still be trimmed nose-
down as speed increases.

MMo for an aircraft also takes account of the Mach number at which shockwave
intensity will cause enough separated flow to reduce elevator effectiveness, or
cause control 'buzz', or both. This control 'buzz' will become control buffet if the
aircraft is accelerated further and is formally termed high-speed buffet.
Eventually, the buffet leads to loss of elevator control. In the early days of high-
subsonic speed flight, many aircraft experienced this high-speed buffet by
inadvertently accelerating beyond MMo in turbulence and then wrongly
mistaking the buffet for low speed pre-stall buffet. Similarly, aircraft in high-
speed descents with speed brakes deployed (which produce buffet) have flown
fast enough to get high-speed buffet, but it has not been identified due to the
masking effect of the speed brakes.

push
50

25
stick
force 0 1--"'T---r----.----,,--'""---,---,---,-----,-----
(!bs) ·90 Mach number
25

50
pull

push Mach trim input~/----,


50
/
/ ''
/ '\
25 ~--«:_· ----~----~~~---~
/
stick

·90 Mach number


25

50
pull

Fig. 16-28. Change in stick force with Mach number


and resultant force with Mach trimmer.

16-20 H(gh Speed Flight The Commercial Pilot Series


If an aircraft is trimmed for a given cruise speed and accelerated, a progressive
push force on the elevator is required to maintain the initial pitch attitude,
assuming the elevator trim is not used. However, as already described, at higher
Mach numbers, the aircraft becomes progressively subject to a nose-down pitch.
Consequently, as Mach number is increased, the elevator push force
progressively changes to a pull force to maintain the set pitch attitude. Fig. !6-28
illustrates the change in stick force required for a typical subsonic wing/tailplane
combination, as the aircraft is accelerated.

To solve this confusing control problem, aircraft designed to operate at high


subsonic cruise speeds but which have wing sections not fully optimized for
transonic flight, are fitted with a 'Mach trimmer' as mentioned previously. The
Mach trimmer provides a control input in the opposite direction to the nose-
down pitch, of sufficient magnitude to still require a progressive increase in
elevator push force as Mach number is increased. The lower diagram in Fig. 16-
28 shows the result of Mach trim input.

Lateral Control
Disturbances in the rolling plane are often experienced with aerofoil sections not
designed for transonic flight by unequal formation of shockwaves on either wing.
Apart from the different amount of lift available from each wing (causing
uninvited roll), the formation of shockwaves can result in loss of aileron
effectiveness through flow separation ahead of the aileron surface. To
overcome this problem at high subsonic speeds, in many aircraft the outboard
ailerons are disengaged (or faired) for high-speed flight and lateral control is
achieved through the use of spoilers or a combination of spoilers and inboard
ailerons. As shockwave formation normally occurs first on the outboard sections
of the wings, inboard ailerons are less subject to the effects of compressibility.
Where combined inboard ailerons/spoilers are employed, the ailerons will
normally be used alone to provide low rates of roll, with the spoilers deploying
automatically when a higher rate of roll is demanded.

The use of spoilers and/or inboard ailerons at high speed also provides a solution
to the problem where, for example, the upward force generated by the
downgoing (outboard) aileron tends to cause the outer wing to twist nose-down
about its torsional axis-resulting in a decreased angle of attack and a roll in the
opposite direction to that demanded. As described in Chapter 7, this effect is
known as aileron reversal. As they are attached to a thicker and stiffer part of
the wing, any twisting moment caused by inboard ailerons is more easily
resisted and, because of their action and positioning, spoilers do not produce a
twisting moment.

Directional Control
As with other control surfaces, the rudder will normally have reduced
effectiveness in the transonic range, when shockwaves form ahead of the main
hinge line. Some modem high-speed fighter aircraft, required to manoeuvre in
the transonic region, are fitted with an all-moving slab fin. However, for
transport aircraft designed for high-subsonic speed cruise, a conventional fin/
rudder combination is retained so that slow speed directional stability
requirements are satisfied.

The use of rudder at high speeds near Mcdr can result in a yaw in the opposite
direction. Application of the rudder will cause one wing to travel faster than the
other which drives it further into the high-speed flow regime with a resulting
increase in drag. This increase in drag will result in a yaw in the opposite

Principles of Flight High Speed Flight 16-21


direction to that demanded. As a result of this sensitivity to yaw control at high
Mach number, all aircraft designed for high speed/high altitude flight employ a
yaw damper which rigidly monitors directional control requirements and
provides very small inputs of rudder at the earliest possible stage when required.

Design for High Speed Flight


There is a need with modern transport aircraft to cruise at the highest possible
subsonic speed with the best possible economy-i.e. having the least drag
possible at the desired cruising speed so that the fuel burn is minimized. It is
therefore essential to be able to cruise at the highest possible Mach number
before the effects of the drag rise at Mcdr are felt. The following are the main
design features employed to achieve this.

Wing Thickness/Chord Ratio


One of the biggest obstacles to cruising efficiently at high Mach number, is the
drag associated with flow separation behind a strong shockwave. If the onset of
shockwave formation can be delayed, and/or the shockwave strength can be
reduced, the drag penalties will also be reduced.

In the earliest days of design research for high speed flight, it was found that a
wing with low thickness/chord (tic) ratio had much better drag characteristics
than one of high t/c ratio. With a low t/c ratio, the airflow over the top surface at
the cruise angle of attack, has much lower local Mach numbers, thus delaying
Merit and with it, Mcdr. In addition, the intensity of the shockwaves is reduced
which in turn reduces the amount of boundary layer separation and the amount
of total drag generated when Mcdr is reached. The Co curves in Fig. 16-29
demonstrate the benefits of a low t/c ratio in terms of the delay to Mcdr and
reduction in total drag.

12% tic

'
/'"-
/
1
'-?%tic
------~~ : -------
1 I '
I I
I I

~ 1·0 Mach
Fig. 16-29. Effect of thickness/chord ratio on Co. Mcdr Mcdr number
at 12% at 7%

I
I
I 12% tic
',- ---:::-::.----
1
---~---.
---- I I r----
I I : [ : 7%Vc-
I I
I I l I
Merit Merit 1·0 Mdet Mdet
12% 7% 7% 12%
Mfs
Fig. 16-30. Effect of thickness/chord ratio on CL.

16-22 High Speed Flight The Commercial Pilot Series


Low t/c ratio wings also give significant improvements in stability and control
through the transonic range. The CL curve at Fig. 16-30 shows how rapid
transonic changes in lift can be avoided with obvious advantages in longitudinal
control and stability. Note however that, as for slow-speed flight, the lifting
capability of a low t/c ratio wing is also reduced throughout the transonic range
by comparison with the thicker wing.
Although low t/c wings offer advantages for transonic/supersonic flight, they
have significant disadvantages for slow speed flight. With a much lower CLmax
these 'thin' wings have stalling speeds which are much higher than for a high t/c
wing and the stall tends to be more sudden. High-lift devices (leading and
trailing edge flaps etc) must be used to achieve reasonable approach speeds
and landing distances.

Supercritical Wing Section


The supercritical aerofoil was first developed in 1965 with the purpose of
increasing the value of Mcdr but without incurring the penalty of a reduced
CLmax which is brought by a low t/c (i.e. 'thin') wing section. A comparison
between what we could consider to be a more conventional aerofoil shape and
the the supercritical aerofoil is depicted in Fig. 26-32.

12% 15%
conventional supercritical
section section

Fig. 26-32. Typical supercritical aerofoil shape.

The flatter upper surface of the supercritical aero foil reduces the acceleration of
the top surface airflow and thus delays Merit to a higher figure. In addition,
when the shockwaves form, they are weaker than for the conventional aerofoil.
This extends the gap to Mcdr and improves the transonic drag rise
characteristics.
The flattened upper surface reduces the amount of lift which would be available
from a wing built with this section. To compensate, the undersurface of the
supercritical section has a pronounced 'reflex camber' over about the rear 30%.
This provides increased lift from the lower surface, and also helps to stabilize the
trailing-edge flow, reducing the wake and the drag.
A supercritical wing section provides an increased efficient cruising speed
before Mcdr is reached while retaining a relatively high CLmax for good low
speed performance. Additional advantage is gained by its relatively high t/c ratio
providing sufficient depth for lighter-weight construction and more room for
carriage of fuel etc, in the wings. A number of modern aircraft have this type of
wing section, including the Boeing 767/777.

Sweepback
A brief consideration of all modern transport aircraft designed for high speed
cruise would indicate that wing sweep is an essential design feature of such
aircraft. Following the first exploratory flights to high Mach numbers in the late
!940' s, it was soon discovered that wing sweep had a significant effect in
delaying the formation of shockwaves and allowing higher speeds to be
achieved.

Principles of Flight High Speed Flight 16-23


Fig. 16-33 shows a wing which is swept back at an angle relative to the lateral
axis of the aircraft. The TAS vector (V) can be resolved into one component at a
right angle to the leading edge and the other parallel to the leading edge CV2l.
Since the flow component V2 has no effect on the flow across the wing, it is the
component V1 which is responsible for determining the pressure pattern
developed by the wing. Thus it it this component of flow across the wing which
dete1mines the value of Merit and Mcdr. Consequently, a much higher value of
Mfs (V in the diagram) can be flown before V1 reaches an ML of 1·0, so delaying
the formation of shockwaves.

0 = angle of sweep

if V = Merit (straight)
V1=Vcose
(a lower speed)
/

/
/ ''
I \ v V2 = component of V
parallel to leading edge
I \
I \
I
I /
I //
I//
/

Fig. 16-33. The effect of sweep back on Merit.

Because it is the component of TAS at right angles to the leading edge (V1)
which determines the amount of lift at that section, and that is less than theTAS
(V), then the amount of lift developed at any given swept-wing section is lower
than for the same TAS with no sweep. Another simple way of looking at it, is to
consider the wing section which lies across the wing at V1 in the above diagram.
That wing section has a certain t/c ratio, but when viewed at a right angle to V-
the actual direction of the freestream airflow-the effective t/c ratio is much
lower. Consequently, the CLmax for a given aerofoil is less for a swept wing than
for a straight wing, and the slope of the CL cmve is lower, as shown in Fig. 16-34.

Disadvantages of Sweepback
straight

swept
Fig. 16-34. Comparison of straight
and swept wing CL curves.

The great advantage of wing sweep for high-speed transport aircraft, is that Mcdr
is delayed and a higher economical cruising speed can be obtained. The use of
sweepback to obtain this advantage, however, has a tendency to create several
undesirable handling characteristics, which the designer must be careful to
overcome. In summary, these are:

16-24 High Speed Flight The Commercial Pilot Series


Sample Examination
Part I Principles of Flight
I. An aircraft in a level turn at constant speed:
A is not accelerating, since its speed is steady;
B is in equilibrium and thus not accelerating;
C is accelerating because it is changing direction;
D is at constant velocity and is accelerating.

2. The physical property of weight:


A is the same as mass;
B is the force produced when a mass is acted upon by gravitational
attraction;
C remains constant regardless of distance from the centre of the earth;
D is measured by mass times velocity.

3. If an aircraft lands at normal weight but at a faster speed than normal, it


will be more difficult to bring to a stop because:
A of its greater momentum;
B it has higher inertia;
C the wheels have less traction with the runway;
D its kinetic energy has decreased.

4. Dynamic pressure (Y, pV') is obtained by;


A subtracting static pressure from total pressure;
B adding static pressure and total pressure;
C subtracting total pressure from pilot pressure;
D measuring the pressure in the pilot tube.

5. In accordance with Bernoulli's Theorem, in a streamline flow of air around


an aerofoil at low subsonic speed:
A where the speed is increased, the static pressure is increased;
B where the speed is decreased, the static pressure is decreased;
C where the speed is increased, the static pressure is decreased;
D static pressure remains constant at all points in the flow.

1-2 Appendix The Commercial Pilot Series


6. The centre of pressure of a cambered aerofoil:
A remains at about the mid-chord position over the normal operating
angle of attack range;
B moves fmward as angle of attack is increased until, passing the
stalling angle, it moves to the rear;
C increases until the stalling angle is reached where it suddenly tilts to
the rear;
D moves forward as airspeed is increased and rearward as it is
decreased.

7. When an aerofoil is at the angle of attack for best lift/drag ratio, the total
reaction is:
A as near to a right angle to the relative airflow as it can be;
B at a right angle to the relative airflow;
C parallel with the relative airflow;
D at a right angle to the effective airflow.

8. The coefficient of lift incorporates the following factors:


A angle of attack, TAS, wing area;
B angle of attack, shape and condition of the aerofoil;
C angle of attack, density and velocity;
D size and shape of the aerofoil, angle of attack.

9. By comparison with a wing of low aspect ratio, a wing of high aspect ratio:
A produces lift more efficiently;
B stalls at a higher angle of attack;
C is not as efficient in producing lift;
D has a lower thickness/chord ratio.

I 0. Separation of a streamlined airflow around an aero foil occurs when:


A the boundary layer changes from laminar to turbulent flow;
B the lower flow in the boundary layer is brought to a stop and begins to
reverse;

C the boundary layer flow meets an adverse pressure gradient;


D the upper flow in the boundary layer mixes with the lower flow and
re-energises it.

II. Some wings are designed with washout in order to:


A prevent stalling from the wingroots first;
B reduce the effects of adverse yaw;
C avoid wingtip stalling;
D improve lateral stability.

Principles of Flight Appendix 1-3


12. In straight and level flight, induced drag:
A remains constant;
B increases as lAS is increased;
C decreases as lAS is increased;
D reduces until the speed for minimum drag is reached, then increases
again.

13. Choose the selection of words that correctly completes the following
statement.
When trailing edge flaps are lowered, the coefficient of lift
is ......................... and the stalling angle of attack is ............................... :
A reduced reduced;
B reduced increased;

c increased reduced;
D increased increased.
14. Choose the selection of words that correctly completes the following
statement.
Movement of an aircraft about its ...................... axis is
called ......................... . which is controlled through the use
of ............................ :
A longitudinal yaw rudder;
B normal yaw rudder;
C normal roll aileron;

D lateral pitch flaps.

15. If the elevator trim control is moved fmward (i.e. to relieve forward stick
pressure) the trailing edge of the trim tab moves:
A up;
B back;
C down;
D across.

16. the primary purpose of aerodynamic balancing is to:


A trim out the stick forces;
B prevent a control from running to full deflection when moved;
C reduce adverse yaw;
D adjust the ease with which a control can be moved in flight.

1-4 Appendix The Commercial Pilot Series


17. Choose the selection of words that correctly completes the following
statement.
If the load factor on an aircraft is increased, the stalling speed will
be ........................ and the stalling angle of attack will ............................. :
A increased remain the same;
B decreased remain the same;
c increased. be increased;
0 increased be reduced.

18. In a piston-engined aircraft, the least amount of power is required for


steady straight and level flight:
A at the minimum drag speed;
B at the speed for best lift/drag ratio;
C at the minimum thrust speed;
0 at a lower speed than the minimum drag speed.

19. In a steady climb at a constant lAS:


A the forces are not in equilibrium, since lift must be greater than the
weight;
B the forces are in equilibrium, with the resultant of lift and thrust
balancing the resultant of weight and drag;
C lift must be greater than weight; thrust must be greater than drag;
D there must be an excess of power required over power available.

20. Maximum rate of climb is achieved at theTAS where:


A there is the greatest excess of lift over weight;
B there is the greatest excess of thrust available;
C the angle of climb is the steepest;
0 there is the greatest excess of power available.

21. In a glide at constant lAS:


A weight is balanced by the resultant of lift and drag;
B lift must be equal and opposite to weight;
C weight is greater than the resultant of lift and drag;
D lift equals weight, the forward component of weight balances the
drag.

Principles of Flight Appendix 1-5


22. An aircraft is in a balanced 30' bank level turn at I ,000 ft and an lAS of 120
knots. Assuming constant weight, what happens to the radius and rate of
turn if the aircraft is flown in a turn at exactly the same bank angle and lAS,
but at 8,000 ft?
A Radius decreases rate remains constant;
B Radius and rate both remain exactly the same;
c Radius increases rate decreases;
D Radius and rate both increase.

23. Design manoeuvring speed 0/A), is the speed:


A which should never be exceeded in any circumstances;
B above which, it is possible to exceed 'g' limitations when
manoeuvring;
C should be exceeded only in smooth air conditions;
D below which the aircraft cannot be stalled.

24. In an aircraft with a fixed-pitch propeller, if speed is decreased at a


constant throttle setting, rpm will decrease because:
A propeller torque has decreased;
B propeller torque has increased;
C the angle of attack of the propeller has been decreased;
D there is decreased slippage.

25. Choose the selection of words that correctly completes the following
statement.
If its centre of gravity is moved to the aft limit, an aircraft will
be ......................... stable in pitch and the elevator 'stick force' will
be ................................. .
A more heavier;
B less lighter;
c more lighter;
D less heavier.

26. If an aircraft is displaced in yaw and a skid develops, dihedral will:


A cause roll in the same direction as the yaw;
B cause roll in the opposite direction as the yaw;
c have no effect on roll;
D prevent further yaw.

1-6 Appendix The Commercial Pilot Series


Part II Performance

31. The take-off distance required (TODR) is defined as:


A the speed which gives an adequate margin above the stalling speed in
the take-off configuration;
B the length of runway declared by the aerodrome operator as available
and suitable for the ground run of an aeroplane taking off;
C the length of ground run required by an aeroplane when taking off
from a standing start at maximum take-off power;
D the distance required to take-off from a standing start at maximum
take-off power and reach a given screen height above the nmway at
the take-off safety speed.

32. An aeroplane is climbing at 600 fVmin into a headwind of 10 knots and at


an lAS of 120 knots. A close approximation of its climb gradient with
respect to the ground is:
A 4'·,
B 4%;
c 5·5%;
D 5·5'.

33. An operator under CAR Part 135 is required to ensure that for landing, the
aeroplane will be able to:
A land inside the runway threshold and come to a stop within the
landing distance available, using maximum braking;
B make a full-stop landing from 50 ft above the runway threshold within
85% of the landing distance available;
C make a full-stop landing from 50 ft above the runway threshold within
115% of the landing distance available;
D from 50 ft above the runway threshold, land within the landing
distance declared available for that runway, using maximum braking.

34. If the aerodrome elevation is 1750 ft, QNH 1029 hPa, what is the
aerodrome pressure altitude?
A 1270 ft;
B minus 170 ft;
c 2230 ft;
D 3670 ft.

1-8 Appendix The Commercial Pilot Series


27. Choose the selection of words that correctly completes the following
statement.
In a certain twin-engine aircraft, both wing-mounted engines rotate
clockwise when viewed from the rear. Asymmetric blade effect will offset
the thrust lines of the engines to the ........................... at high angles of
attack, making the ........................ engine the critical engine to have fail.
A right left;
B right right;
c left left;
D left right.

28. The best range speed (still air) is the speed for minimum drag. If altitude
is increased, the best range:
A lAS decreases, TAS remains constant;
B lAS and TAS both increase;
C lAS remains constant, TAS increases;
D lAS increases, TAS remains constant.

29. At cruising speed in a certain aircraft, the centre of pressure is behind the
centre of gravity, but the lines of thrust and drag coincide (i.e. there is no
thrust/drag couple). In order to prevent the nose pitching up or down, the
tailplane must provide:
A a force which will depend on where the elevator trim is set;
B an upward force;
C no force, as it will taken care of by longitudinal dihedral;
D a downward force.

30. If a certain aircraft is fitted with a more powerful engine, but without
increasing its drag or basic weight:
A its climb performance and service ceiling will be improved, but there
will be no change in its ability to turn;
B its climb performance, service ceiling and turning ability will all be
improved;
C turning performance will be improved, but not the climb or service
ceiling;
D there will be no change in performance as the weight has been
unchanged.

(Part II - Performance begins on the following page.)

Principles of Flight Appendix 1-7


Appendix 1
Sample Examination

This appendix contains a sample examination paper. It is in two parts: Part I -


Principles of Flight; and Part II - Aeroplane Performance. The subject matter for
this paper has been drawn from the New Zealand CAA syllabus for the
Commercial Pilot Licence Examination in Principles of Flight and Aeroplane
Performance.

The format and style of the questions are similar to those employed for the
examinations by Aviation Services Limited (ASL).

If you have studied the preceding chapters and completed the chapter reviews,
you should have no difficulty in answering the questions in the sample
examination. It is suggested that you allow yourself two hours to complete the
paper, without referring back to the main text (or to the answers) in the manual.
This will enable you to test yourself in as near as possible to real examination
conditions.

The correct answers are given, together with those for the chapter reviews, at
Appendix2.

NOTES ON MULTI-CHOICE QUESTIONS

Multi-choice questions are made up of a 'stem' and a number of choices. The


stem either poses a question or makes a statement. The choices provide a
selection of answers to the question, or different ways of completing the
statement. Unless stated othe1wise, only one of the choices (A, B, C or D) will
be correct.

You should read the stem carefully, and formulate an answer in your own
mind. Then examine the choices and select the one which coincides with your
considered answer. Look out for the 'poisoned lolly' - a choice which may
seem at first sight to be correct, but which in fact is close to the correct answer
or just 'sounds right'.

If you are unable to identify the correct answer after the first reading, it
sometimes helps to eliminate those choices which are clearly wrong. Then,
with a little concentration and logic applied, the correct response should
become apparent.

For those questions which require some arithmetic, or employ the use of graphs
or charts, take some care to be as accurate as you can. The ASL examinations
permit the use of a small portable electronic calculator. While it is not
absolutely necessa1y for the questions in Part II, the use of such a calculator will
be of some assistance to finding the answers to some of them. Where possible,
complete a double check of your figuring and the use of graphs, charts and
tables. For questions requiring a numerical answer, select the choice which is
the closest to your own calculation.

Most examinations have a time limit. If an answer is not clear to you, or you
cannot see how to proceed after a minute or so, leave the question and return
to it if there is time at the end.

Principles of Flight Appendix 1-1


25. Use the head/tail wind component graph on page I 7-21. The reported wind
is 050/25 and the runway in use is OS. What are the headwind and
crosswind components on this runway?
26. Use the head/tail wind component graph on page 17-21. The reported wind
is 230/IS and the runway is 16/34. What are the head/tail wind and
crosswind components on both runway vectors?
27. Use the take-off graph on page 17-25. Aerodrome elevation is 12SO ft, QNH
is 1019 hPa and the reported ambient temperature is +2JOC. The aircraft all-
up weight is 2420 kg, runway OS/26 has a grass surface and slopes down 2%.
ATIS reports the wind as OS0/40. What is the required take-off distance?
2S. Use the take-off graph on page 17-25. Aerodrome elevation is sea level, the
QNH is I 004 hPa and the temperature is ISA-3. The sealed runway has a
take-off distance available of 620 metres, an up-slope of I% and a headwind
component of 24 knots. What is the maximum all-up weight of the aircraft if
it is to take off within the distance available?
29. Use the landing graph on page 17-33. Aerodrome elevation is 1500 feet,
QNH is 1032 hPa and the ambient temperature is + I2°C. The aircraft all-up
weight is 2220 kg, runway 03/21 is sealed, it has a slope of I% up and the
headwind component is 30 knots. What is the required landing distance?
30. Use the landing graph on page I 7-33. Aerodrome elevation is 5SO feet, QNH
is 99S hPa and the temperature is ISA+6. The grass runway has a landing
distance available of 450 metres, a down-slope of 2% and a headwind
component of 32 knots. What is the maximum all-up weight of the aircraft if
it is to land within the distance available?
31. You are an airfield of 1000 feet elevation which has a single grass runway
06/24. The QNH is lOIS hPa and the ambient temperature is+ !Soc. Aircraft
all-up weight is 2400 kg. The wind is almost totally crosswind giving a
headwind component of 6 kt on runway 06 and a tailwind component of 6
kt on runway 24. The slope on runway 06 is 2% up and on runway 24 2%
down. Using the take-off graph on page 17-25, which runway vector
provides for the shortest take-off distance?
32. Use the take-off chart on page 17-37. For a Part 135 operation, if the mnway
elevation is 250 ft, the QNH I 030 hPa, ambient temperature 24°C, take-off
weight 1522 kg, the runway is paved with a I% downslope, and the
headwind component is 24 kts, what is the take-off distance required?
33. Use the landing chart on page 17-3S. For a Part 135 landing on a grass strip
at 1200 ft elevation and a downslope of I%, QNH I 003 hPa, landing weight
1470 kg, with a headwind component of 10 knots, what is the landing
distance required?
34. Use the single-engine service ceiling graph on page 17-40. You plan a flight
in a twin engined aeroplane over a 12,700 feet mountain barrier. The area
QNH is 1022 hPa and the forecast lists a temperature of -TC at 10,000 ft and-
!Soc at 15,000 ft. What is the maximum all-up weight if you are to maintain
at least 2000 ft terrain clearance in the case of an engine failure?
35. Use the single-engine service ceiling graph on page 17-40. Your twin
engined aircraft all-up weight is 7200 lb, the sea level air temperature is + 19°
C and the area QNH is I 00 I hPa. What is the highest altitude you can
maintain on one engine?

17-48 Performance The Commercial Pilot Series


15. You are about to depart from an airstrip of unknown elevation. You manage
to obtain the area QNH of 1020 hPa over the aircraft's radio. When you set
this in the subscale window the altimeter reads 1280 feet. What is the
pressure altitude of the place?
16. You are at a place of unknown elevation and you cannot obtain the area
QNH. How can you determine the pressure altitude of the place?
17. Fill in the blank spaces in the following table.

Pressure altitude Outside air temperature ISA temperature


oc !SA+ or ISA-
1500 +22
6800 -5
FLISO -8
4000 1SA+4
2500 lSA-6

18. Fill in the blank spaces in the following table.

Elevation QNH Pressure Outside air Density


altitude temperature altitude
2500 1028 +20°C
3600 3180 +Soc
1001 7300 lSA-6
Sea level 998 +22°C
7000 7540 7180
1029 4300 +I4°C
1200 900 420

19. When the !SA temperature deviation is + 10°, you (add/subtract) ....... feet
(from/to) the (elevation/pressure altitude) to obtain the density altitude.
20. Using the typical flight manual graph on page I 7- I8, calculate the take-off
distance to 50 ft under the following conditions; pressure altitude 1000 ft,
ambient temperature soc, take-off weight 6700 lb, headwind component I 0
knots.
21. Assume the flight you were planning for in Q 20 above was for an air
transport operation under Part 135 where the take-off was to be from a grass
runway, with an upslope of I%. Using the extract from CAR Part 135 given at
page 17-23, what is the amended take-off distance once the appropriate
factors are applied?
22. If the take-off run available (TORA) for the take-off in Q 21 was 3000 ft,
would the pilot be able to comply with Part 135 in respect of take-off
distance required (TODR)?
23. Under certain conditions during a Part 135 operation, you have calculated a
landing distance using flight manual data of I 700 ft. If the landing is to be
made on a metal strip with a 2% upslope, what is the corrected landing
distance required?
24. If the published landing distance available for the landing in Q 23 is 670
metres, will the pilot be able to comply with Part 135 in respect of the
landing distance available (LDR)?

Principles of Flight Performance 17-47


Review 17
I. The take-off distance required-TODR-(does/does not) include the
distance required to reach a given screen height above the runway.
2. The take-off distance available-TODA-is defined as the length of take-
off ........ available, plus the length of any ...................... .
3. For aeroplanes operating under CAR Part 135, the TODA must not
exceed ....... %of the take off run available (TORA).
4. A clearway (is/is not) the same as a stopway and (may/may not) be ground
or water.
5. The accelerate-stop distance available specified by the appropriate authority
(may/may not) include a stopway.
6. The shorter the accelerate-stop distance available, the ................ the speed
beyond which the aircraft can be brought to a halt in case of an abandoned
take-off, and the ..................... the allowable take-off weight.
7. You climb your aircraft at an lAS of II 0 kt at sea level under !SA conditions.
The headwind component is 22 kt and the rate of climb is 800 ft/min. What
is the climb gradient (a) with respect to the airmass and (b) with respect to
the ground?
8. The landing distance required is (a) the length of the runway, or (b) the
distance measured from a point 50 ft above the runway threshold to where
the aeroplane can be brought to a complete stop?
9. A wet runway is defined as one with sufficient moisture on its surface to
cause it to appear ......................... but without significant areas
of ........................... water.
10. A contaminated runway is defined as one which has more than ...... %of its
surface area within the required length covered by surface water, slush or
loose snow, or has .......... on any part of the runway surface area.
II. The International Standard Atmosphere (!SA) assumes that the atmospheric
pressure at sea level is ............. hPa, the air temperature at sea level is .........
0

C and the temperature lapse rate is ........ °C/1 000 ft up to ........... feet. (Use
'rounded off figures.)
12. Apart from pressure and temperature, the density of air is also affected by
its .................... content in that the (higher/lower) the humidity of air the
lower its density.
13. In the lower atmosphere, one hPa equals a height of approximately ...... .
feet. When calculating pressure altitude, this value is added to the altitude
(or elevation) for every hPa (above/below) 1013 hPa.
14. Fill in the blank spaces in the following table.

Altitude QNH Pressure altitude

1500 1031
2400 2700
998 8150
Sea level 1004
3720 3360
!027 5390

17-46 Performance The Commercial Pilot Series


We now interpolate for temperature. 16°C is 10°C + 0·6 of the difference
between 10° and 20°C. The take-off distance at 16°C will therefore be 2590 +
(0·6 of 2847- 2590) = 2744 ft.

This distance of 2744 ft is for nil wind Compare this with the figure of 3210 ft we
obtained by the approximation method. To establish what the effect of the
headwind component will be, we must follow a note on the take-off table which
states that for each 10 knots headwind, the distance should be decreased by 7%.
Hence, for 19 knots, the decrease is 2744 x (19/10 x 0·07) = 365. The take-off
distance with 19 knots headwind will thus be 2744-365 = 2379 ft.

Footnote

When we are using a graph or chart, we interpolate when we estimate the


position of a point or a line which is part way between two other guidelines (as
we have done several times when using take-off and landing charts). Although
the result is an estimate, it will be a close approximation to the actual figure
provided we take reasonable care in plotting the positions of the various points
on the graph or chart.

Interpolation with the use of tables is more accurate. Although it is time


consuming, it is a technique which you should be aware of and capable of using.
It is likely that some of the performance planning data for the aeroplanes which
you will fly (including medium and large aeroplanes) will be in tabular form and
will require interpolation in a similar manner described for the above example.

Principles of Flight Performance 17-45


Interpolation Method.

In this method, we must interpolate between the increments given, to fit the
table to the actual conditions. Usually, interpolation between two horizontal
lines of data will be first required, followed by interpolation between two vertical
columns of data. In the extract from the table shown for Example 36, we must
interpolate three times horizontally, because of the way the data has been
presented. In the sense in which we are using the word, interpolation means
calculating or estimating values from known ones in the same range.

First, we must reduce the two lines of data for each of the pressure altitudes of
2000 and 3000 ft to one line representing the data for the actual pressure altitude
of 2400 ft. 2400 ft is 400/100 (or 0·4) of the difference between 2000 and 3000 ft.
Therefore, if we take the differences between between the 2000 and 3000 ft
values, multiply them by 0·4 and add these figures to the 2000 ft value, we will
have a line of data representing a pressure altitude of 2400 ft. For example, for a
weight of 6200 lb the total distance to clear 50 ft at I ooc is 2220 + (2400 - 2220 =
180 X 0·4) 72 ft = 2292 ft. By the same process, at a weight of 6200 lbs the total
distance to clear 50ft at 20°C is 2480 + (2620 -2480 = 140 x 0·4) 56ft= 2536 ft.

Complete the same interpolation (add 0·4 of the differences to the lower figure)
for a pressure altitude of 2400 ft at the higher weight of 6750 lb, and we now
have a table which has been reduced to:

10"C 20"C
Total Total
Take-off Distance Distance
and Climb Pressure Ground to Clear Ground to Clear
Weight Speed Altitude Roll- 50ft- Roll- 50ft-
Pounds KIAS Feet Feet Feet Feet Feet
6750 98 2400 2834 3102

6200 94 2400 2292 2536

Note we have dropped the ground roll figures since these are not required.

Next, we must reduce the two lines shown above to one, representing the actual
weight of 6500 lbs. This weight (6500 lbs) is 6200 lbs +300/550 (0·55) of the
difference between 6200 lbs and the next increment up of 6750 lbs. Interpolate
in the same way as we did for pressure altitude, i.e. multiply the differences by
0·55 and add to the lower figure. We thus produce one line representing the
data for the actual weight of 6500 lbs, thus:

10"C 20"C
Total Total
Take-off Distance Distance
and Climb Pressure Ground to Clear Ground to Clear
Weight Speed Altitude Roll- 50ft- Roll- 50ft-
Pounds KIAS Feet Feet Feet Feet Feet
6500 96 2400 2590 2847

Note we have also interpolated for take-off and climb speed as well as the take-
off distance.

17-44 Performance The Commercial Pilot Series


Example 35 (page 42). It is possible to use the outside air temperature grid
along the x-axis instead of ISA values but when doing so you must follow the
vertical grid lines.

Obstmctions demand that you can maintain at least FLJ50 on one engine. The
forecast stated the freezing level at 9,000 feet. What will be the aircraft's
maximum weight to achieve this?

Since a flight level is pressure altitude we do not have to calculate it. Using an
average lapse rate of 2°C/1 000 ft we can calculate the temperature at FL150 by
multiplying (15 - 9) = 6 x 2 = -l2°C. Draw a line vertically from -12° and draw a
line horizontally from 15,000 feet. Where these two lines intersect, make a mark
which is about l/3rd away from 7450 towards 6800 lb. Thus the answer is
approximately 7200 lb.

Use of Tabulated Performance Data


For many aircraft, some of the performance data is presented in the flight
manual in the form of a table, rather than as a graph or chart. The use of such
tables is straightfOlward provided the increments of temperature, altitude and
other variables given in the table can be matched directly to the conditions being
planned for. However, this does not often happen, and it becomes necessary to
use either approximate values or (if an accurate result is required) to interpolate
between the incremental values given.

The following gives an example of the approximation and interpolation methods


of handling tabulated performance data, using an extract from a typical take-off
distance table for a light twin-engined aeroplane.

Example 36. Given take-off weight 6500 lbs, pressure altitude 2400 ft,
temperature I6°C, headwind component 19 knots, establish the take-off and
climb speed and the total distance to clear 50 ft.
Extract from Normal Take-off Distance Table
(not to be used operationally)

10'C 20'C
Total Total
Take-off Distance Distance
and Climb Pressure Ground to Clear Ground to Clear
Weight Speed Altitude Roll- 50ft- Roll- 50ft-
Pounds KIAS Feet Feet Feet Feet Feet
6750 98 2000 2350 2770 2570 3030
3000 2500 2930 2730 3210
6200 94 2000 1880 2220 2100 2480
3000 2040 2400 2230 2620

Approximation Method.

Extract from the table the figures for the next increments of weight, pressure
altitude and temperature which are more conservative than the actual
conditions. In this case they are 6750 lb, 3000 ft, and 20°C. The approximate
determination of take-off speed and distance to 50 ft, is therefore 98 knots and
3210 ft.

Note that this take-off distance required of 3210 ft is for nil wind conditions. If it
is well within the take-off distance available, then no further take-off calculations
need be made. If however, it is close to or less than the TODA, the take-off must
be determined accurately by using interpolation.

Principles of Flight Performance 17-43


First establish the average !SA value. At 10,000 ft !SA is 1S- (2 x 10) = 1S- 20 =-
soc whereas the forecast temperature is -8°C. This is 3° colder than !SA,
expressed as ISA-3. At FL180 !SA is 1S - (2 x 18) = 1S - 36 = -21 oc and the
forecast temperature is -26°C, so colder than !SA and expressed as ISA-S. Thus
the average temperature is ISA-4, draw this line parallel to the nearest !SA guide
line.
Next draw a line representing the all-up weight of 72SO lb, about one quarter
distance from 74SO Ib towards 6600 lb. Where this weight line intersects the ISA-
4 line, draw a line horizontally to the Y-axis to obtain a pressure altitude of
17,600ft.
The question required you to calculate the single engine ceiling in terms of
altitude, thus you have to convert pressure altitude to altitude (elevation) given a
QNH of 1026 hPa.
E =?
Q 1026
P-17,600

QNH is greater than 1013, pressure altitude is lower than elevation or, turning
this around, elevation must be higher than pressure altitude. Therefore, I 026 -
1013 = 13x30 = 390 + 17,600 = 17,990 feet altitude.
SINGLE-ENGINE SERVICE CEILING
26

24

22

20

E< 18

.''""
0
0
0 18
~

'"
~ 14

~
12

~
"'
~ 10

Fig. 17-15. -40 -30 -20 -10 0 10 20 30 40 50


Example 35.
OUTSIDE AIR TEMPERATURE °C

17-42 Performance The Commercial Pilot Series


Forecasts give temperatures in oc for selected altitudes. For example, a
domestic high level forecast may indicate an expected temperature of MOl (-I}·
at 10,000 feet and MIS (-IS) at FL ISO. The !SA temperature at 10,000 feet is IS-
(2 x 10) = IS - 20 =-S°C while the forecast temperature is -rc which is 4o
warmer than !SA, expressed as ISA+4. At FLISO !SA temperature is IS- (2 x IS)
= IS - 36 = -2 I oc whereas the forecast is -I soc, 3° warmer than !SA, expressed
as ISA+3. The average !SA temperature is therefore around ISA+3, draw this
line on the graph parallel to the !SA guide lines.

Where the pressure altitude line (from 14,390 ft) crosses the ISA+3 line, make a
mark and estimate the associated all-up weight from the weight guide lines
either side of the mark. The mark is about l/3rd the distance from 74SO lb to
6SOO lb. Since the total weight difference is 6SO lb, I/3rd = 220 lb in round
figures. Deduct this from 74SO and you have established the maximum weight
(7230 lb) your aircraft can carry on this particular day to comply with the single-
engine ceiling requirement. In other words, should an engine fail, the aircraft at
7230 lb will be able to maintain a pressure altitude of I 4,390 ft.

Example 34. An alternative to these calculations is to ascertain the maximum


altitude an aircraft can maintain while on one engine when the all-up weight is a
given value. For example, the QNH on the day is 1026 hPa, the all-up weight is
72SO lb and the forecast shows that the temperature at 10,000 ft is expected to be
-S°C and at FL I SO -26°C.
SINGLE-ENGINE SERVICE CEILING
26

24
~ ~1sl~~~ ~~1~t~t~ =~~l~l~ ~~~~~l~ ~~~~~~J~ ~t~t~~~ ~l~l~t~~ ~l~l~~;~
22

20
~~~IWil~[i i i Ill:
~~r: ~~~t~{~r~ l]~it ,~~,~~~~f~~j~ ~!ri~J~ ~mR ~mu
E-< 18

"'~ -Cj: __ ft.li~ ,:':::'ij:t~f =~;::: ,:::t~ :;:::::: :::::::: =N:d:


0
0
0 16
====~===r-k~~,t~~~:~,~=~~ ~~:\:= ~\=~:::: ~::::: :::::::: =f=~=
~
----·--- '"'., '\ --·t?i:::i\'., _,_ ':K;;;ttt\'., -- ---"\'- -------- -----"'-
"' :::::::: ::=~=:~;:~ ~:::E ·~t:~: ~~:~f: :~:t::t ~;0:::: ::::::~:
~ 14

~
12
"'~
fll
fll

"'~ 10 ~ ... _,__._ __

8
--,----- -------- ------\- --'~--',- --Y---',-~"o;- '\'\---0, ~:-
6
~~~~~~~~ ~~~=~~~~ ~~~~~~~~ ~i~~]~: ~i~:~~~ ~~~~:~:~4~
\ \: \ ·-\~ ~~~==~~ ~~=~~~
. I \ H-~
I I I"' '

4
Fig. 17-14. -40 -30 -20 -10 0 10 20 30 40 50
Example 34.
OUTSIDE AIR TEMPERATURE °C
Principles of Flight Performance 17-41
Fig. 17-13. SINGLE-ENGINE SERVICE CEILING
Example 33. 26

24

22

20

8 18
1"<1
1"<1
"'
0
0
0
~
16

1"<1

~ 14

~
1"<1 12
~
rn
rn
~ 10

4
-40 -30 -20 -10 0 10 20 30 40 50

OUTSIDE AIR TEMPERATURE o C

Example 33. It will will probably be easier to understand the principles of


operating this graph if we use an example. You plan to fly your twin-engined
aircraft over a 12,000 ft mountain range which demands that it can maintain at
least 14,000 ft (under lFR conditions) in case one engine fails (you require a
2,000 feet clearance). It is your responsibility to ensure that the all-up weight of
the aircraft is such that its single-engine ceiling is at least 14,000 feet altitude
(elevation).

You must first calculate the pressure altitude which corresponds to the altitude
of 14,000 feet, this depends on the QNH of the day. Let us assume the QNH is
1000 hPa. Since this is lower than 1013 hPa, pressure altitude is higher than
actual altitude by (13 x 30) = 390 ft. Thus the minimum pressure altitude your
aircraft must be able to maintain is 14,000 + 390 = 14,390 feet. From this value
on theY-axis, draw a horizontal line across the graph.

Secondly, you must establish the temperatures you are likely to encounter. The
simplest way is to establish !SA values and draw the appropriate line parallel to
the lSA temperature guides. Temperature information is normally available to
you from forecasts or from actual reports.

17-40 Performance The Commercial Pilot Series


En-route - Engine Inoperative Performance
There is a requirement for multi-engined aeroplanes operating in accordance
with Part 135 that the aeroplane is to be capable of continuing flight at a positive
slope at or above the relevant minimum safe altitudesOJ, to a point 1000 feet
above an aerodrome which is suitable for landing, taking into account the
forecast meteorological conditions, and assuming the critical engine is
inoperative and the remaining engines are operated at the recommended
maximum continuous power setting. When planning a route to take account of
the foregoing, there is also a requirement that the assumed en-route gradient
with one engine inoperative is the gross-gradient-minus-0.5% gradient<2J.

In practice what this all means is that, in a twin-engined aeroplane, you should
not plan to fly a route over which you would be unable to maintain an obstacle
clearance of at least 2000ft under IFR with one engine inoperative (and at least
I OOOft during a VFR operation) and arrive at a minimum of I OOOft overhead a
suitable aerodrome for landing.

NOTES:
(I) Acceptable means of obtaining the minimum safe altitudes for the flight are
given in AC 119-3. Briefly, these can be summarised as being either (a) to
conduct a detailed high terrain or obstacle analysis of the route, or (b), to use the
published Minimum Enroute Altitude (MEA)*, or Minimum Off Route Altitude
(MORA)* as appropriate. Because (a) is time consuming, it is likely that
preference will be given to using MEA or MORA. (* Given in the En-route
Charts.)
(2) What this means is that, at the engine-inoperative ceiling which has been
assumed for planning purposes, the aeroplane must be capable of maintaining a
climb gradient of O·So/o under ideal conditions. (For a typical light twin-engined
aeroplane, this equates to having a capability of climbing at roughly 50 ft per
minute under ideal conditions.)

Single-engine Service Ceiling Graph


To assist route planning, the flight manual for a twin-engined aeroplane should
provide you with a single-engine service ceiling graph. Fig. I 7-13 provides a
typical example of this type of graph. (Fig. 17-13 is not to be used operationally.)
In US-sourced aeroplanes, the ceilings which are calculated using this type of
graph, take into account the required en-route gradient capability as described
above.

When an engine fails during high-altitude cruise flight in a twin-engined aircraft


there is often a loss of height, known as drift-down, to the single-engine ceiling.
This ceiling depends mainly on the all-up weight of the aircraft and the density
altitude at which it can operate on one engine. In general, the heavier the
aircraft and the warmer the conditions the lower the single-engine ceiling.

When planning a flight the all-up weight is adjusted, if necessary, to enable the
aircraft to maintain a minimum safe altitude above terrain should one engine
fail. The graph in Fig. I 7-13 shows the various elements that determine this
capability.

There are two temperature guides, degrees Celsius along the horizontal x-axis
and matching slanted !SA lines. The vertical y-axis shows pressure altitude
values. Combined, these two parameters provide density altitude information.
The solid slanted lines represent all-up weight variations.

Principles of Flight Performance 17-39


D\f-tiCw-!Bb'l I I IIA
~I I I'J I I tEl ~I I I I I'~ I I I 11 ooo
"'
-~
, I , I . , I 'f !"! l , I . ~"II I ··u I 1\
, tI II"J .1\I II. ' t', E~ I.I l'j
'I'
I,I I ill
, , I . Ill lllll,l,'llii I'lis
I~12
~
LANDING J 1
900

2(])
E

11111-fHIIIII~
This chart •was prepared In accordooce ~Mih AdvioOIY D
CifCuklr PC 119-31£1ng mariUfoc!urefs data to
pRXJuce a chart that can be used forcompl;once 800 1 E
0
with NlCAR R:Jrt 135 Subpart D Pelfolmonce. s (.)
The chorl rxovides !he IOiollonding distance T (])
required from a helglll ol50 reet.
The c hort iflCQrpomtes the solely factors of.
A
N
r::
Port 135.211 Slope ol'\d Surface Cooect!ons
Porll35.223(b) Landing £k>tonce
c
700 E

R
E
Q

11111111 uld~U'J lkf 11111111111111 UJ I fHU I fll IRIIJ'HU 1 ~


Model: Sample
600
Not to be used operationally
D
M
,; 500 E
T
§ R
.8 E
s

c, liNE A - PriVate Operations· Paved Runway
UNE B -Air Transport Opemllons- Paved Runway II II Ufl Jz1111111111111111111111 UItHl I tfffi llii"U tR11YfLoo
"'.!!! UNE C -I>Jr Transport Operations - Metal Runway
UNE D - AJr Transport Opemfions- Gross Runway
.l!l
&

~
3 210123 5 0 10 20 300
DOWN UP TAIL HEAD
SLOPE[%) REPORTED WIND

i"'
{KNOTS)
4000 PROCEDURE FOR USE
1, Loca:e the point corresponding to
1l<::
-.J
3000
the pressure akitude for the illlftekl
2. Mo\li! horizontally across to the !roo "'E
N
~
AIRFIELD
PRESSURE
corresponding to the !anding wetght
3 Move vertically 4l to the )jne .g
corresponding to the type of operation
"-
~
STARTHERE • ALTITUDE
{feet) 2000
and~ SU!faca
4. move across to the zero slope l:oo,
<f
co
~
then either back up to the appl~eal>le

1000
da{lrae of down slope, or down the
correct kin line to the degtee of up :2
~

0
"'"'
5, Move across to the zero Wll1d line,
thetl either back up op lha app\lcabla
S.L taimind speed, or doY.n the rorroctiOn
J,ne to !he applicable hoodwind speed
n'~
~<> !ANDING WEIGHT 6. Read the f8qU!red diStance on the
&~ nght hand sJde or the wind pooet
;$]_.?
~(C B
~

J:li! ~

= ~
1J
g-
-o·
--
";"

m- TAKE-OFF 900
: -'
Q;i
Q,
::0 lh!:i Chor! WOS Pf€p::J18d in OCC01dOflCO V.'lh MV.COJY
"''P"
<Q· Crrcuior AC 119·3 USing monulacturMs data to
~
::t pro::!vce a Cllort tho! con be u:>cd lor compcunco
wrlh NZCAA Port 135 Subpart D Pe<lorrnonco
The chart prow.:.Jes me totot tol'-e·oll d::tan.:::e
reqwred to och.~·.<e o hO;ght or 50 reet 800
D
I
~
;<.
The chart Incorporate, u-.:;. safety factors at s

~
Port l35.209(a) Tar.e-olt Crstanco T
Part 135 211 SlOpe and Surface Correct;ons A
N
Model: Sample Chart
not to be used operationally 700
c
E "'"'q
R
E
c5·
Q ::t
u 0
~
600 I
R
E
D

M
500 E
T
R
UNE A- PriVOie Operolions - Paved RUIWIOY E
UNE 8 • /'>Jr Transport Operations • Paved Runwoy s
LINE C -Air Transport Operations - Metal Run'h'Oy
LINE D - Ail Transport Operations - Gloss Runway
400


START HERE

AMBIENT TEMPERATURE (''C) 3 2 0 1 2 3 5 0 10 20 300


0 10 20 30 UP DOWN TAll HEAD
SLOPE(%} REPORTED WIND
(KNOTS)
J'
a- PROCEDURE FOR USE
1. Locate lha poifll CO!Tespood:ng to the
ambent teml)'lrotura and PAll for the day
'3 2 Move horizoola!ly across to the hne
§ c.orrespondmg to lakg..off wrnght
3 Move vertreally up to the ~ne
!SA+20
2
-
R l correspond>ng to the type of operatoo and
E T runway surface
$1 ,eft' 4 move across to the zero slope line, tllcn

~ ISA+lO e;!her tmc:k up to the appl!cab!o degree of


S T up s!opa. or down the rorrect100 line to the
uu
" dog!oo of down slope

-
R 0 .<> 5 Move across to the zero w.tnd l!rn!, tllen
e1!her back up op the apphc.."'b!e tail-hmd
E E '\.& ISA speed, or OO.m lhe o:xmcl100 hne to the
iFD app!icabk! headvliOd speed
eft'
~
-90'
!i"'
<9_t;;
&,i t"'. ~.
§i .zy
TAKE-OFF WEIGHT
6 Read lhe mqu:red d•stance on the nght
hand s:de of the w.nd panel

~"' "'"
,_y· ISA-10
Use of Later P-charts
As an acceptable means of compliance with Part 135, operators may develop new-
style P-charts (from flight manual data) in accordance with the guidance material
given in Appendix C to AC 119-3. Examples of take-off and landing charts developed
in accordance with this AC, are shown at Figs. 17-II and I 7-12 respectively.
Although their layout is slightly different, the use of these charts is very similar to
those which we have just discussed in detail. A small diagram providing a key to
their use, together with written step-by-step notes, is located at the bottom right of
each chart. Some additional notes on their use are as follows.
I. The charts include the safety factors of Part 135 slope and surface corrections,
and take-off and landing distance respectively and use \1, H/W or 1.5 T/W. The
graphs can thus be readily used to check compliance with Part 135.
2. The take-off chart is entered by establishing ambient temperature against
pressure altitude in the normal way. The landing chart requires only pressure
altitude to be entered. (For landing, the effect of correcting for temperature is
insignificant.)
3. Three take-off weights are shown - interpolate for intermediate weights.
Often, the data for only one landing weight (MCLW) is available in the flight manual,
and thus interpolation for lower landing weights is not possible in landing charts
based on this data.
4. The required safety factors are taken account of in the panel above the weight
panel. For Part 135 operations use lines B, C or D as appropriate (Line A does not
include the "85%'' safety factor). Note that because these lines are moderately
steep on the graph, accuracy in tracing upward from the weight line is essential
to obtaining an accurate take-off or landing distance required-
Example 31. Follow through with this example using Fig. 17-11. For a Part 135 air
transport operation from a paved runway, what is the TODR under the following
conditions. Elevation 570 ft, QNH 1005, temperature 19'C, take-off weight 1590kg,
\1,% downslope, I 0 knots headwind component?

Solution_ QNH is down, so pressure altitude is (8 x 30 ft) or 240 ft up from 5 70 = 81 0


ft. Place a mark where the 19'C line meets an imagina1y 810ft line.
For ease of interpolation, convert 1590 kg to lbs [x 2.2= 3498 (say 3500 lb)]. This is
two-thirds of the distance between the 3300 and 3600 lb lines, so draw a line there for
the take-off weight in this case. Where the horizontal line from the temperature
mark meets this line, make another mark.
From there, proceed as indicated by the instructions. Carefully up to the next line
(line B in this case), then across to the slope reference line, down to '/2%, across to
the wind reference line, down to 5 knots (half of 10 kt), and then across to read off
the TODR. This should be 640 metres (say ±I 5 metres). If it isn't, check your
procedure and accuracy of marking again.
Example 32. For a Part 135 air transport operation, a landing is required on a grass
runway under the following conditions: Elevation 1130 ft, QNH 1022, landing weight
1410 kg, 1% upslope, headwind component 14 knots. What is the LDR?
Solution. Complete the example using Fig. 17-12. You will need to calculate the
pressure altitude, and use only the single landing-weight line available. Then,
following the instructions given on the chart establish the LOR. The answer you
should obtain is 540 metres (±I 5). If it isn't, check your procedure and accuracy
again. Note in this case you must use line D and, for an upslope on landing, move
down the guidelines.

17-36 Performance The Commercial Pilot Series


,,, 1!1111111 1111 111111!11 1111 ""
1...1...1... I...I.J..J..j..J..J..J..J ..J.I..I..l .I,J.i.L f-1...1...1...1... 1...1...1...1 ..J..J..J.J_

H~:.w :_~~~-c,_~!~ ~~~~=-


From the grass X mark, draw a horizontal line
un t1.l you s tn·1{e th e sIope re ference 1·1ne and
._,_,_
::::t:r!j::J:::z
,_,_,_,w .J.J.j
:!:l:l :t:t!!
,_,_,_c;::,_,...~::J .... -~ H++
...u ...
~rr 2 ~~~
........ :::i~~- ......
~~, -!-I-l-l-
H-H
''~'
then have a good look where the down-slope :::::: :::::::::tti ~~~ -t-t++ ....... u.:-o-1-1- :::::::: ::::::::::
is, to the left. Draw a line parallel to the slope :::::::: :::::::a::.~~~ ~~H EH!~;~SS :::::::: : : : : : : :
guide lines until you intersect 2% down, make
rrr ,~., .,.,.,,
iN'lc~r-.-
"T"TTT
~~::: ::;:, ~~~~ iiH r'F r r:: :::::::::: ~:::ii
I T
rrt> ,,,..,
,-,-,- 1*"1'1"1 11!"1 TTTT T r rr ITl"l -,"11"1
660
an X mark there. From this mark, draw a :=:=:=~')sQ~:I:l~~~ HH :ft~~ ~~fL:iL:Cf-tc=:=_::::;~_· ?J.~J~J' 'l
horizontal line to the wind reference line and :=~:=~:=:~~~~~ ...... :tU ...... , ~~i'¥
LLL~O~J.lliili.JJ JJ.lJ. .1.1. L ~L LI-I.J ili1J
A

follow the wind guide lines down to the 14 ~~~


~~~
,-_J,...J:l .JJJJ::l :L±±± :tbt
C-\i:j:
tL '- :::::::: ::.J
JJ.lJ. LLL LLI..J ..l..l.J
.J (/)
W
knot value (half of 28 kt) and then draw a ·~~J 1-S;:oo """, :t~
CCL JJJJ 600 L•cL:e:st; ~
horizontal line to the vertical graph axis which from grass tt~~::~ !;;~ HH hh tttt:tjj~ ::jjj :2

indicates that we have a landing distance of x.:;_Ej:E ~~~ ~~ ~~H ~~~ S;:~ ::i~ :~~ w
528 metres (remembering each little square is ;;; :~; ~~~; ;;;;:it;~ ~:~~3: -(s;!~ ;;;; ~
6 metres). ~:::::: q:::: ~ii nn :ttt~~~:: ::~t'iii (/)
FFF :=1::0· ~ri~ HH tftF FF'kf- f1ii ~~t 0
CCL -,333ti:!JJ II!I r~j\t eta -,3JJ ~~~~ ~
To work the graph backwards if the landing 480
::::::1::~: :l::l:f}; f±±± :~tb: !:CCC L'X:l: :l:l:H 0
distance available is limited, the same care tttF83i~:l::l ::11±± :~Ht~~tt t:B"~.J:l:l::l ~
needs to be taken as with the take-off graph. LLC~..l...J
t... LL.t... -l
.JiffiJ
...J.J.i'>l • .l.t.JJ
.J.J.J-1-., .l.-l.l..l
J.*L
J.J. L
LL't-L
'-!-~
'--'-'
'--1...1 ..I.J ...1

The next two diagrams (which assume a :::::: •::::: ... K:i:i:i :!:!!! tt :: ::::~~-- ':!:::::: :::: .. 420
headwind component of 16 kt (use half= 8 kt) """
rt-1-
E:"' " ' _._._;_.-. ,
,_,_,_

1-H-
++++
++++
.,.,1"+
...........

1-1..1
..1
..1~~· ...........

"'"~.,.,
-1-1-1"'
and a I% up-slope) show the right and the ::::::: :::::::: ~~~ .,..,. .... ,.;...).j- ... -..... 3lli"'
~~~~ ,..,..,..,. ........... ~~~ I -,.., .,-, 1
orr ,-,..,-, .,.,.,., ..,..,.tT rrrr .-. .,,.,
wrong way of doing this. rrr ,..,.,-, .,.,.,., rrrr rrrr -,-, -,-,.,.,
TTTT
360

.. "" '""'"' ......... . :::: ·:::

he<e

Notes on Runway Surface Condition

P-charts (whether older-style or new) make a distinction between paved and


grass runways so that the different acceleration and braking characte1istics can
be taken account of when calculating take-off or landing distance required.
When using P-charts you should be aware that they do not differentiate between
wet or dry sealed surfaces or short, long or wet grass. You should be mindful of
this and make an allowance (which must be approximate) for runway
conditions that are likely to result in poor acceleration or poor braking action,
e.g. long wet grass.

Remember also, that for aeroplanes operating in accordance with Part 135, there
is a requirement for operators to adjust the landing distance required to make an
allowance for wet and contaminated nmways. For example, under Part 135.225,
operators of normal category aeroplanes must ensure that if a landing is being
made on a wet runway, the landing distance available must be at least 115% of
the landing distance required by Part 135.223.

Principles of Flight Performance 17-35


Landing Graph
The landing graph (Fig. 17-10 on previous page) is very similar to the take-off graph
and requires the same method of working. There are one or two differences,
however, which need to be noted.

The pressure altitude/temperature block has more !SA lines and is spread out
more. In the all-up weight block there are only three weight guide lines instead of
five in the take-off graph and the runway surface guide lines are much further
apart.

A very important difference to note is the slope block. Whereas a down-slope is


advantageous for take-off, it has the opposite effect on landing, it lengthens the
landing distance. Thus the up-slope and down-slope are reversed in the landing
graph compared to the take-off graph.

The wind component block is the same but note that distances are plotted in 60
metre lots which means that each little square is 6 metres, not 5 as in the take-off
graph.

Example 30. Airfield elevation is


1400 feet, QNH is I 029 hPa, the
temperature is !SA+ 7. The aircraft
all-up weight is 2350 kg. The grass
runway has a down-slope of 2% and
the headwind component is 28 kt.
What is the landing distance?

First determine pressure altitude.


E = 1400
Q = 1029 (higher than 1013 :. P.Alt is
down). 1029- 1013 = 16 x 30 = 480
ft.
p = 1400-480=920.

Draw the 920 ft line parallel to the S/L


and 1000 ft pressure altitude guide
lines and draw the !SA+ 7 line
parallel to the slanted !SA lines, more
or less equal distance between
ISA+5 and !SA+ 10.

From the intersection of the two


lines, draw a line vertically down
parallel to the grid until you strike the
reference line and make an X mark
there. From this mark, draw a line
parallel to the weight guide lines and
note again that these guide lines are
not straight. Draw a horizontal line to
the right from 2350 kg and where this
line intersects the line you have just
drawn, make another X mark. From
this mark, draw a line vertically down
to the grass surface line and place an
X mark there.

17-34 Performance The Commercial Pilot Series


Not to be used operationally

...
·~·~·~·~·.

660

600

540 '
I
l'l

s~
480
"~

420

360

This graph is for an imaginary aircraft on Air Transport operations, Day.

Fig. 17-10. Landing graph.

Principles of Flight Performance 17-33


The WRONG and the RIGHT diagrams show clearly how a mistake can be made.

Assuming that the available take-off distance is 540 metres, there is a 20 kt


headwind component and a 2% up-slope. The WRONG diagram shows that the
wind guidelines instead of the horizontal grid were followed initially to the I 0 kt
wind (half of 20 kt). It then repeats the mistake by going to the slope reference
line first instead of to the 2% up slope. The RIGHT diagram shows how the lines
should have been drawn.

17-32 Performance The Commercial Pilot Series


Where the two lines intersect, draw a horizontal line parallel with the grid to the
left and you establish the all-up weight appropriate to the take-off distance
available.

Until this technique is thoroughly understood it is wise to draw all lines rather
than just make X-marks at the various positions. Having drawn all lines and X
marks you can then run your pencil along them forward to ensure that you have
followed the correct procedure.

Example 29 Take-off distance available is 5SO metres, headwind component is


24 kt, slope is 2% down and the runway surface is seal. Aerodrome elevation is
1200 ft, the QNH is I 004 hPa and the ambient temperature is +I soc.

What is the aircraft's maximum-possible all-up '.


weight for the limited take-off distance available? .. ,

Solution. Establish the density altitude first. ::


;:
E = 1200 ;,.

Q = 1004 (lower than 1013 :. P. All is up). 1013- ·, ~

1004 = 9x30 =270ft. ::·. ·:!

p = 1200 + 270 = 1470 ft.


.. ..
Draw a 14 70 line parallel to the pressure altitude ""
lines, almost exactly halfway between I 000 and :i :·;

2000 ft. Then draw a horizontal line to the right '""' ..


parallel with the grid starting from + IS°C ambient
..
and from where the pressure altitude and ,.., '. ,.
,.
,;.

down to the all-up weight reference line, make an


X mark there. """
:I

From the X mark, draw a freehand line parallel to


the weight guide lines. This is as far as you can """ :.·

work forward.

From 5SO metres on the far 1ight of the graph, draw a horizontal line to the left
until you come to a headwind component of 12 kt. Now follow the wind guide
lines up until you intersect the reference line, make an X mark there. From this
point draw a horizontal line to 2% down slope, then follow the slope guide lines
to the reference line and make another X mark there. From that mark draw a
horizontal line to the left to intersect the seal surface line and make another X
mark there. From this mark draw a line vertically up through the weight guide
lines.

Where the vertical line you drew intersects the earlier one you drew parallel
with the weight guide lines, make an X mark and from that mark draw a
horizontal line to the left. Where this line intersects the vertical graph axis, read
the maximum possible all-weight that can be carried bearing in mind the limited
take-off distance available.

Having drawn all these lines, now run your pencil forward from the X mark in
the weight guide line section and ensure you have followed the slope and wind
sections correctly.

Principles of Flight Performance 17-31


Continue on from mark X at 2% up-slope (example 22), draw a horizontal line
parallel to the grid until you strike the wind reference line. From that point,
follow the slanted wind guide lines until you intersect the II knot line you drew
vertically up.
From the II knot wind speed intersection, draw a horizontal line parallel to the
grid and intersect the take-off distance line. Read off the take-off distance
required, 600 metres in this example.

Example 28. This example establishes the take-off distance required when
there is a tailwind component of 6 knots (use 1.5 x = 9 kt). We will assume you
have worked through the various parts of the graph to mark X at 2% down slope.

Draw a line straight up from 9 knots tailwind parallel to the grid.

Continue on from mark X at 2% down-slope, draw a horizontal line parallel to


the grid until you strike the wind reference line. From that point, follow the
steeply slanted tailwind guide lines until you intersect the 9 knot tailwind line
you drew vertically up.

From the 9 knot tailwind speed intersection, draw a horizontal line parallel to the
grid and intersect the take-off distance line. Read off the take-off distance
required, 1230 metres in this example. Note the substantial influence of a
tailwind on take-off distance required.

Calculating Maximum All-up Weight


There are occasions where the take-off distance available does not allow the
aircraft to be loaded to its maximum permissible all-up weight and it is then
necessary to calculate an all-up weight that does permit a safe take-off within
the available distance.

The technique requires you to commence with the density altitude section and
goes as far as drawing a line parallel to the all-up weight guide lines. Then you
must start on the right of the graph and work backwards. You will finish with an
intersection of lines in the weight section and by drawing a horizontal line to the
left from that point you will establish the maximum weight of your aircraft for the
take-off distance available.

There is a very common tendency to misuse-use the graph guide lines when
working backwards. To avoid mistakes, have a good look at the wind and slope
sections when working forward. Note that allowance for head or tailwind
component is made by drawing a line parallel to the wind guide lines from the
reference line and when the appropriate wind strength has been reached you
draw a horizontal line to ascertain the take-off distance.

Similarly, slope is allowed for by drawing a line from the reference line parallel to
the slope guide lines until the appropriate degree of slope is intersected, then
draw a horizontal line to the wind reference line.

Thus if you work backwards, draw a horizontal line (from the take-off distance
available) parallel with the grid until you meet the wind strength, then follow the
wind guide lines to the reference line. From there, draw a horizontal line
parallel with the grid to the degree of slope and then follow the slope guide lines
to the reference line. From there draw a line parallel to the grid to the type of
surface and then a vertical line through the weight guide lines.

17-30 Performance The Commercial Pilot Series


Allowing for Slope
If a runway has a slope it will affect the aircraft differently on take-off compared to
landing. Whereas a down-slope assists the take-off and produces a shorter take-off
distance required, it has an adverse effect on landing. Thus it is most important that
you pay particular attention to the values of
the slope guidelines, they are opposites for
take-off and landing.

Example 24. Assume the grass runway has .'


a I% down-slope. Continuing from the X
mark on the grass surface of example 22,
draw a horizontal line parallel to the grid
until you strike the slope reference line.
From that point, draw a line of best fit to the
slope guide lines, down to I% (i.e. only half-
way down the lines) and place a mark X
there.

You will note that the slope guide lines are


not parallel to each other. You must
therefore estimate the mean distance
between the guide lines when drawing your
line of best fit.

Example 25. Assume the sealed runway has a 2% up-slope . Continuing from the X
mark on the hard surface of example 23, draw a horizontal line parallel to the grid
until you strike the reference line. From that point, draw a line parallel to the slope
guide lines, up to 2% and place a mark X there.

Allowing for Wind Velocity


The graph makes allowance for headwind
components up to 30 knots and tailwind
components up to I 0 knots.
Example 26. Following on from example 24,
assume there is a headwind component of 30 kt
(use half = I 5 kt). Draw a line straight up from
15 knots (headwind) parallel to the grid.
from
Continue on from mark X at I% down-slope "omp~o
(example 22), draw a horizontal line parallel to
the grid until you strike the wind reference line.
From that point, follow the slanted wind guide
lines until you intersect the 15 knot line you
drew vertically up. Note again that the wind
guide lines are not parallel to each other which
means that you must estimate a line of best fit
between them. ..
,,,,

From the 15 knot wind speed intersection, draw a horizontal line parallel to the grid
and intersect the take-off distance line. Read off the take-off distance required, 770
metres in this example.
Example 27. Following on from example 25, assume there is a headwind
component of 22 knots (use half = II knots). Draw a line straight up from II knots
(headwind) parallel to the grid.

Principles of Flight Performance 17-29


Allowing for All-up Weight
From where the density altitude line has stmck the reference line, draw a line
parallel to the weight guide lines. Note that these lines are slightly curved and
therefore you cannot use a mler.

Draw a horizontal line parallel to the grid from the aircraft all-up weight value
and where this line
intersects the previous line
you drew, make a mark.

Example 22. Using point X


from example 20,
determine the intersection
point if the aircraft all-up
weight is 2450 kg.

Draw a parallel line


between the appropriate
weight guide lines starting
at point X of example 20.
Then draw a horizontal line
to the right and parallel to
the grid from 2450 kg. Where the two lines intersect, place another mark X.

Example 23. Using point X from example 2 I, determine the intersection point if
the aircraft all-up weight is 2220 kg.

Draw a parallel line between the appropriate weight guide lines starting at point
X of example 21 Then draw a horizontal line to the right and parallel to the grid
from 2220 kg. Where the two lines intersect, place another mark X.

Allowing for Runway Surface


Most graphs make an allowance for sealed or grass runways, but they do not
differentiate between wet or dry sealed surfaces or short, long or wet grass. You
should be mindful of this and make a further allowance, which has to be
approximate to be sure, for
runway surfaces that cause slow ""
acceleration or that produce poor '""
braking action, e.g. wet grass.

Following on from point X on the '"' '


weight guide lines of example 22,
if the runway surface is grass, .. :
draw a line straight down until it '
intersects the slanted grass guide
line and place a mark X there. : ..

Similarly, following on from


'.
example 23 draw a line from the
X mark at 2220 kg straight down :: . ·~
to hard assuming the surface is
·.. ;
seal and place a mark X there.
.
'·.! .·
' '•.! f> .•

17-28 Performance The Commercial Pilot Series


Example 20. The aerodrome elevation is
1250 feet, QNH is I 02S hPa and the i-i· '''H"···
f+:;:;:i:i::i.:::::::::
! : i: k dti;i+i::Hi:-+-+:
temperature is reported as ISA+5.
Establish the density altitude intersection.

E = 1250
Q = I 02S (higher than I 013:.
P. All is down). 102S- 1013 = 15 x 30 = ·'··'··:-·:-· ·)·f·t·t·
450ft. . .• ,.,.,. ·i·i·i·i·
+H+
•f•l•i•i• ·l·l··:··
·i·l·i•(• .,•.••.

p = 1250-450 = soo ft. ·:·:·:·:·

Draw the SOO ft line between the slanted


S/L and I 000 ft guidelines.

Temperature was given as ISA+5, therefore draw a line parallel to the slanted
!SA lines at the appropriate distance (halfway in this case) between !SA and
!SA+ I 0. From where the two lines intersect, draw a line down to the reference
line and place a mark there.

Go back briefly to the intersection of the pressure altitude line and the ISA+5
line. If you draw a thin line horizontally to the left (parallel to the main grid) you
will see that it intersects the ambient temperature line at I soc which is the
equivalent temperature in oc for ISA+5 at this pressure altitude. (!SA should be
!3°C at SOO ft [nearest 1000 ft] and thus ISA+5 is !Soc at that altitude.)

Example 21. The aerodrome elevation is 24SO feet, the QNH is IOOS hPa and the
ambient temperature is + !6°C. Establish the density altitude intersection.

E = 24SO
Q = IOOS (lower than 1013 :. P. All is up).
1013-IOOS = 5x30 =150ft.
p = 24SO + !50= 2630 ft.

Draw the 2630 line parallel to the pressure


altitude lines at the appropriate distance
between the slanted 2000 and 3000 ft guide
lines.

Temperature was given as + l6°C. Draw a


horizontal line, parallel to the grid, from 16°
to the right. From where this line intersects
the 2630 pressure altitude line, draw a line
down to the reference line and place a mark
(X) there.

Go back again briefly to the intersection of the temperature and pressure altitude
lines. You can see that l6°C at 2630 ft is equivalent to ISA+6. At 2630 ft !SA is 15
- (2 x 2Yz) = 15- 5 = 10°C and since the reported ambient temperature was [ 6°C
you can see that this is 6° warmer than !SA.

Thus if the temperature is given in !SA, draw an appropriate line parallel to the
slanted !SA lines but if the temperature is given in oc ambient, use the oc scale
and draw an appropriate line horizontally to the right, in line with the grid.

Principles of Flight Performance 17-27


The Use of Older-Style P Charts
The take-off and landing graphs (P-charts) which we use in the following
examples are similar to those which were formerly provided by the CAA for
inclusion in the flight manuals for light transport aeroplanes. Although the
provision of these P-charts has been discontinued by the CAA, those charts
which are still contained in the flight manual may continue to be used for
perfmmance planning under Part 135. As we have indicated, they incorporate
the required safety, runway surface and slope correction factors. Hence, when
using this type of chart, there is no need to apply the correction factors shown at
page 23 of this chapter. The take-off and landing distances required (obtained
from these charts) may also be directly compared with the take-off and landing
distances available. That is, the 85% safety factor required for transport
operations has also been incorporated in the charts. Note that the 50%
headwind and 1.5 tailwind factors remain in place and must be applied.

Take-off Graph
Take-off distance required can be calculated using the graph at Fig. I 7-9 on the
previous page. This graph allows for density altitude, aircraft all-up weight, type
of runway surface, runway slope, and head or tailwind component. As the graph
is labelled for "Air Transport Operations" it can be taken that it incorporates the
85% safety factor mentioned above.

You can start at the top of the graph and work your way through to the right
where the take-off distance required can be read off. Alternatively, if the take-off
distance available is limited, you can calculate the maximum gross aircraft
weight which will be acceptable by working from right to left. Both systems will
be explained in the following section.
The graphs we use in this chapter are for an imaginary aircraft. The actual
"CASO 4" graphs provided for the aircraft by CAA will probably have arrows and
dotted lines to indicate the method of usage (like the aircraft manual graphs). If
you are going to use the old-style graphs for Part 135 compliance, be sure to use
the graphs which are marked "Air Transport Operations", and the lines marked
for "day" operations (there is no longer a distinction between the day and night
safety factors).

Establishing Density Altitude


As explained previously, to establish density altitude we must first work out the
pressure altitude and then correct this for any temperature deviation from !SA.
Thus it is necessary first to convert the elevation of the aerodrome to a pressure
altitude and then allow for temperature, either in degrees Celsius or as degrees
deviation from !SA (indicated on the graph by 'ambient' temperature).

When you have calculated the pressure altitude of the aerodrome, draw a line
parallel to the pressure altitude lines. Then, having established the temperature,
draw a line representing this temperature parallel to the !SA lines or parallel to
the grid lines, depending on how the temperature was expressed. From where
the two lines intersect, draw a line straight down parallel to the grid until you
intersect the reference line below.

17-26 Performance The Commercial Pilot Series


Not to be used operationally
TAKE-OFF GRAPH

.,., .,. '····~-~-~-~- .;.;.;., .. ;.~.;-~ .. , ....... ,., .


. , ...,., .. ,., ... , .. ,.,.,., ..
....

.........
'' ········· ....
2720
········· ····· .....
:REF LINE

2600

2400

2200

2000

This graph is for an imaginary aircraft on Air Transport Operations, Day.

Fig. 17-9. Take-offgraph.

Principles of Flight Performance 17-25


Example 19. You are operating in accordance with Part 135. For a landing on a
paved runway under the prevailing conditions you have calculated from the
aeroplane flight manual that the landing distance required is 820 metres. The
runway you intend to use however has a published downslope of 2% in the
landing direction. What is the corrected LOR?

The LOR must be increased by 5% for each I% downslope. Thus corrected LOR
for the runway in question is 820 x II 0% = 902 metres.

Summary of the use of Flight Manual Graphs


If you are using flight manual graphs or tables to establish TOOR and LOR in
accordance with Part 135, then the procedure can be summarised as follows:

I. Before using the graphs, establish:


(a) the QNH and elevation of the runway to be used, and calculate the
pressure altitude;
(b) what the ambient temperature is;
(c) the aircraft weight for take-off and landing;
(d) what the headwind (or tailwind) component is, if necessmy using a
wind component graph. (Remember that you should not apply more than
50% of the headwind component, or less than 150% of any tailwind
component in your use of the graphs.)

2. With the above data available and using the flight manual graphs,
establish what the take-off (or landing) distance is at the take-off (or landing)
weight.

3. Correct the flight manual take-off or landing distances as necessary


depending on the runway surface or slope, using the correction factors given at
Part 135.211. These corrected figures now represent the take-off distance
required (TOOR) or landing distance required (LOR).

4. To comply with Part 135, the TOOR and LOR must not exceed 85% of the
take-off distance available (TOOA) and landing distance available (LOR)
respectively. To check this, divide TOOR and LOR by 0·85. The resulting figures
must not exceed the respective TOOA and LOA (which can be obtained from the
appropriate aerodrome chart in the appropriate Volume of the AlPNZ.

5. For landing on a wet runway, the figure obtained by dividing LOR by 0·85,
must be further increased by 15% (ie multiplied by 1·15) and this figure must not
exceed the published LOA.

NOTE: From the foregoing, it can be seen that while it is not particularly difficult,
using flight manual graphs to establish TOOR and LOR in accordance with Part
135 requires several calculations and can be time-consuming. It is preferable to
use the old-style P-charts (based on CASO 4) or, if these are not available, to use
the newer P-charts which can be developed from the aeroplane flight manual
data in accordance with AC 119-3. These cha1ts incorporate the safety, surface
and slope correction factors, and compliance with Part 135 can be much more
quickly and easily established.

We now turn to the use of P-charts, both the old-style, and the new.

17-24 Performance The Commercial Pilot Series


Extract from Civil Aviation Rules Part 135.
Subpart D - Performance

135.211 Runway surface and slope correction factors

Each holder of an air operator certificate shall ensure that, unless performance data is available
that authorises an alternative, the take-ofF distance calculated for a runway surface type under
135.209(b)(5) or 135.229(c)(4) and the landing distance calculated under 135.223(c)(3) and
135.233(c)(3)-

(1) are corrected for use of other runway surface types by applying the factors in
Table 1; and

(2) are corrected for runway slope by-

(i) increasing the take-ofF distance by 5% for each 1% uphill slope to a


maximum of 3% upslope; or

(ii) decreasing the landing distance by 5% for each 1% uphill slope to a


maximum of 3% upslope; or

(iii) decreasing the take-ofF distance by 5% for each 1% downslope to a


maximum of 3% downslope; or

(iv) increasing the landing distance by 5% for each 1% downslope to a


maximum of 3% downslope.

Table 1

Paved X 1.00 X 1.00 X 1.00

Coral X 1.00 X 1.03 X 1.05

Metal X 1.05 X 1.06 X 1.08

Rolled earth X 1.08 X 1.14 X 1.16

Grass X 1.14 X 1.20 X 1.18

Application of the foregoing factors is straightfmward, as shown by the following


examples.

Example 18. You are conducting an air transport operation in accordance with
Part 135 in a twin-engined aeroplane. You are planning a take-off on a grass
runway and under the particular conditions for that day you have calculated a
take-off distance required frorn the flight manual of 450 metres. What is the
corrected take-off distance?

The flight manual figure relates to a paved surface. In accordance with Table I
above, the flight manual TODR must be increased by a factor of x 1.14. Thus
corrected TODR for the grass runway is 450 x 1.14 = 513 metres.

Principles of Flight Performance 17-23


reciprocal runway heading, and then designate that as a tailwind. The
crosswind component remains the same regardless of head or tailwind.
Using Fig. 17-8, follow through on the following two examples.

Example 16. The wind velocity is 160°M/35kt and the runway-in-use is II. What
is the head or tailwind component and what is the crosswind component?
Solution. The difference between runway II (II 0°M) and the wind direction
(160°M) is 50°. Follow the 50° line until you reach the 35 knot arc. From there,
trace a horizontal line to the left which intersects the y-axis at 23 knots and trace
a line vertically down which intersects the x-axis at 27 knots.

Thus the headwind component is 23 kt and the crosswind component is 27 kt.


Example 17. The runway in use is 08 and the wind is 200°M/12kt. What is the
head or tailwind component and what is the crosswind component?
Solution. The difference between 080° and 200° is 120°. Follow the 120° line
until you come to the 12 knot curved line. From the intersection draw a line to
the left which indicates a tailwind of 6 knots and draw a line vertically down
which indicates a crosswind of II knots. Note that you would have obtained the
same answer if you had worked out the headwind component for the reciprocal
runway, then called that a tailwind.
NOTE:

If you are using aeroplane flight manual graphs to work out TODR or LDR, to
comply with Part 135 you must take account of "not more than 50% of the
reported headwind component or not less than 150% of the reported
tailwind component". This means that after working out the wind component
as just described, you should apply not more than half of the headwind
component (or not less than 1·5 times the tailwind component) to the take-off
and landing graphs.

Runway Slope and Surface Correction Factors


The take-off and landing performance data furnished by the manufacturer in the
aeroplane flight manual will normally apply to a level, paved, dry runway. If the
runway to be used is sloping or has a different type of surface, runway slope and
surface correction factors must be applied to the take-off, accelerate-stop or
landing distances which have been calculated using the flight manual figures.

Both the older-style P-charts, and any charts developed in accordance with AC
119-3, have the runway slope and surface factors incorporated in them (so there
is no need to apply these factors separately to the take-off or landing distances
produced by these charts). However, if the flight manual graphs are being
used to calculate take-off and landing performance under Part 135, the factors to
be applied to the flight manual figures are given at Part 135.211 as shown in the
extract on the next page.

17-22 Performance The Commercial Pilot Series


Headwind and Crosswind Component Graph
To operate the take-off and landing graphs (and the P-charts) you need to be
able to determine what the headwind (or tailwind component) is for the runway
in use. To facilitate this, most aircraft flight manuals contain a head/tailwind and
crosswind component graph (or grid) which will be similar to that reproduced in
Fig. 17-8. Alternatively, you can use the wind component grid, or the headwind/
crosswind correction tables which can be found on the reverse of the wind slide
of most navigation computers (e.g. the E6B).

Fig. 17-8.
Wind component graph.

0 10 20 30 40 45
CROSS\VIND COMPONENT KNOTS

The wind component graph is simple to use. It shows a 90' (or greater) sector
on which the wind speed and the angle between its direction and the runway
heading can be plotted. From this plot, the headwind (or tailwind) component
can be measured against the vertical (y) axis of the graph, and the crosswind
component against the horizontal (x) axis.

For example, if the runway designator is 35 (heading 350'M) and the wind
velocity is 330'M/20kt, the angular difference between the wind and the runway
is 20'. On the graph, follow the 20' line until you meet the 20 kt arc. From that
point trace a horizontal line parallel to the grid to the left which gives you the
headwind component (19 knots) on the y-axis and trace a vertical line down to
the x-axis to read off the crosswind component (7 knots).

If the angle between the wind and runway heading is greater than 90', then
clearly a tailwind component will exist. Some graphs may allow for a small
amount of tailwind to be plotted, but others may cover only 90' of wind angle. In
these latter cases, work out what the headwind component would be on the

Principles of Flight Performance 17-21


The method of using the landing graph is the same as for the take-off graph, and
again, dashed lines and arrows have been used to provide guidance in its use.
In this case, the example shows calculation of landing distance given an outside
air temperature of 14"C, a pressure altitude of 2000 ft, a landing weight of 6000 lb,
a headwind component of 5 knots.

Note that this landing graph does not give tailwind guidelines, and that would be
typical of a number of flight manuals for aeroplanes of this class as it is expected
that they will not be landed intentionally with a downwind component. Also, an
indication of the approach speed to use at various weights is given.

General Notes on the use of Flight Manual Graphs


I. The take-off and landing distances calculated using the flight manual graphs
do not meet the requirements of Part 135 for the calculation of the take-off or
landing distance required (TODR or LDR). To do that, the distances calculated
by this method must incorporate an adjusted head/tailwind component; and
have the appropriate runway slope and surface correction factors, and other
safety factors applied to them. These factors inevitably increase the TODR and
LDR by comparison with the flight manual figures and thus provide an increased
margin for safety. We will be discussing how these factors are to be applied
shortly.

2. The graphs and tables in manuals for many of the US-sourced aeroplanes in
New Zealand, will more than likely give the temperatures in degrees Fahrenheit
("F) and speeds in miles per hour (MPH). To be of practical use in New Zealand,
such graphs and tables will need to be converted to show the temperatures in "C
and the speeds in knots (as we have shown in Figs. 17-6 and 17-7).

3. Similarly, take-off and landing distances in these manuals are more likely to
be given in feet, rather than metres. To apply the charts in a practical situation,
you will therefore have to convert the calculated take-off or landing distances to
metres, since this is the unit used in the AIPNZ charts for runway lengths. A
sufficiently accurate and easily remembered conversion factor is:

I metre = 3·3 feet

Thus, divide the calculated take-off or landing distance in feet, by 3·3 to obtain
the equivalent distances in metres.
4. Again, in similar fashion (although we have not done this in Figs. 17-6 or 17-
7) it may be convenient to convert the graphs/tables from pounds to kilograms,
depending on which unit your flying organisation uses for weight and balance
calculations. The conversion factor is easily remembered:

I kilogram = 2·2 pounds

Thus to convert from pounds to kilograms, divide by 2·2, or alternatively to


convert from kg to lb, multiply by 2·2.

5. Some manufacturers use tables of figures in their aircraft manuals, rather


than graphs or charts, for some performance data-including take-off and
landing distances. We provide some notes on the handling of such tabulated
data at the end of this chapter, preferring to concentrate at this stage on the use
of graphs and charts.

17-20 Performance The Commercial Pilot Series


Landing Graph
A typical landing graph is shown in Fig. 17-7 below. It is similar to the take-off graph we have
already discussed, and it enables the landing distance from a height of 50 ft over the
threshold to be calculated, given pressure altitude, temperature, landing weight and
headwind.

Once again, the conditions under which the graph is valid are stated, i.e. flap setting; power
used on the approach; a paved, level, dry runway; and with maximum braking.

Fig. 17-7. Typical Aircraft Manual Landing Graph.

LANDING DISTANCE- FT

Principles of Flight Performance 17-19


If you don't get this answer (or very close to it) something has gone wrong, and you should
go carefully through the example again using a pencil to accurately mark the points,
following the same procedure as indicated by the dashed lines and arrows.
Fig. 17-6. Typical aircraft manual take-off distance chatt.
(Rotate goo right to read)

TAKE-OFF DISTANCE- FT
0 0
0 0
0 0 0
N ~
0
"'t::
'
0::
I.U
0::
0::
a\
0
. 3Nil o3lJ
0
I I N
(/)
\;::
I- I I 0
~o '
LL ' z
0 s:
I!) 0
3Nll o3lJ
0::: 0
w 0
0

"'
6 I I

0
0
ill

(/)
co
_,
o'
8!i:
<OCJ
~
0
0
"'
<0

.....J
<(
2 0
0
0::: 12
0
z I I
oo
<'>o

17-18 Performance The Commercial Pilot Series


Use of Aircraft Flight Manual Data to Obtain TODR and LOR
The aircraft manufacturer provides various performance graphs or tables as a part of
the aircraft flight manual. Typical take-off and landing graphs for a light twin-
engined aeroplane are given at Figs. 17-6 and I 7-7 respectively.

Take-off Graph
The typical take-off graph at Fig. I 7-6 enables the calculation of the length of the
ground roll and the take off distance to clear a barrier height of 50 ft, depending on
pressure altitude and ambient temperature (i.e. density altitude), take-off weight,
and head or tailwind component. You will notice that the nmway and aircraft
operating conditions for the graph to be valid are stated (i.e. required power settings
and engine handling, aircraft configuration, runway conditions, rotate and barrier
speeds).
The use of the graph is straightforward, and the manufacturer will normally provide
an indication of the method of its use, for example, with dashed lines and arrows.
The example used in Fig. 17-6 is for a take-off with an ambient temperature of 22°C,
a pressure altitude of 1000 ft, a weight of 6200 lbs, and into a headwind component
of 10 knots.
The method shown in the example, is for the chart to be entered on the left by
tracing straight up on the line representing 22°C until the point representing I 000 ft
pressure altitude is reached (i.e. the point where an imaginary line half way
between the sea-level and 2000 ft pressure altitude guidelines). From this point,
trace straight across until the reference line is intercepted, and from there, parallel
with the nearest sloping guideline until the vertical line representing a weight of 6200
lb is reached.
From that point, the procedure is much the same as has already been used. Straight
across to the wind reference line, then down along a "line of best fit" between the
headwind guide lines until the component of 10 knots is reached. (Notice that the
guide lines are not necessarily parallel, so that if your tracing on the graph falls
between two lines, you must judge the correct slope to follow-a line which
provides the "best fit" between the two.) Notice also, that had there been a tailwind
component, you would have traced upward from the wind reference line as
indicated by the tailwind guide lines.)
From the I 0 knot headwind point, trace straight across to the next reference line.
The point on this reference line represents the ground roll distance for this take-
off,which can be read against the scale on the left as being 1750 ft. The take-off
distance to 50 ft is given by tracing up the barrier guidelines and reading off against
the scale at 2600 ft.
Using Fig. 17-6, follow through on another example.
Example 15. What is the take-off distance to 50 ft, given that the ambient
temperature is l2°C; QNH 1003 hPa; aerodrome elevation 200 ftASL; take-off weight
5750 lb; and headwind component 15 knots?
Before using the chart, we must first determine the pressure altitude. QNH is below
1013, hence P Alt will be above elevation, and it will be above by (10 x 30ft) i.e. 300
ft. Aerodrome pressure altitude is thus 500ft.
Thus on the chart, trace up the l2°C line until you intersect an imaginary line
representing 500 ft (Y,, way up from sea level toward 2000 ft). From this point,
proceed as before-across to the reference lines, then down as guided by the guide
lines until you first intersect the line representing 5750 lbs, then 15 knots headwind,
then up from the last reference line, to read off the take-off distance to 50 ft of 1900 ft.

Principles of Flight Performance 17-17


Navigation Computer Method

On the circular slide rule (the calculator side) of the navigation computer you
will find three windows. Using only the two windows which are close together
as shown in Fig. 17-5, simply set pressure altitude in the lower window against
ambient air temperature, and then read off density altitude in the upper window.
For example, Fig. 17-5 shows that with a temperature of plus 35°C (ISA+20) set
against a pressure altitude of zero, density altitude should read 2,400ft.

Fig. 17-5. Calculation of density


altitude using an E6B navigation
computer.

Although the computer can provide a quick and easy indication of density
altitude, it is difficult to get the same answer as one obtains using the
mathematical method described previously because of the vety small scales
used, particularly for temperature. For this reason, the computer method should
not be used in critical planning situations (or in the examination room!).

The presentation shown in Fig. 17-5 is for the E6B type of computer. Other types
of computer will probably have a different type of presentation, although the
method of operation will be the same as for the E6B.

From Performance Charts

When using performance charts (either graphs or tables) the starting point is
normally to enter the ambient temperature (or deviation from !SA) against
pressure altitude and then proceed from there. Although it may not be
mentioned, in fact what these first simple steps do is to determine the density
altitude affecting the aircraft. You will use this method when we operate the
take-off and landing graphs later in the chapter, and you will be able to see the
effects of temperature and pressure altitude on the distance required for take-off
and landing. It is all based on the principles you have learned in this section.

Calculation of Take-off and Landing Distance Required


The calculation of take-off and landing distance required in accordance with
CAR Part 135 can be achieved in one of three ways:
• by using the take-off and landing data provided by the manufacturer in the
aircraft flight manual, and applying the correction and safety factors
required by CAR Part 135; or
• by using the older CM P-charts (based on the now-obsolescent CASO 4)
which were issued as a part of the aircraft flight manual prior to October
1995; or
• by using newer style P-charts which have been developed in accordance
with AC 119-3 Appendix C.

We now cover each of the above methods in turn.

17-16 Performance The Commercial Pilot Series


Remember, when the subscale is turned to 1013 hPa, the altimeter reads pressure
altitude. Thus P is 2,SOO feet.
Solution:
E = 2200.
Q =?
p = 2SOO.
T = + 16.
D =?
Calculate QNH first. Pressure altitude is higher than elevation :. QNH is less than
1013 hPa. 2SOO- 2200 = 300:30 = 10 hPa. Thus the QNH is 1013- 10 = 1003 hPa.
Calculate density altitude. At 2SOO pressure altitude, !SA= IS- (2 x 21/,) = IS-S =
10°C. The actual temperature is + l6°C = ISA+6 which is warmer :. density
altitude is high. 6 x 120 = 720. 2SOO + 720 = density altitude of 3,220 feet.

Example 14. You are at a place of unknown elevation. You select 1013 hPa on the
subscale and the altimeter reads 3,800 feet. The outside air temperature gauge
shows +soc and you have been passed the area QNH of I 024 hPa. What is the
elevation of the place and what is the density altitude?
E=?
Q = 1024
p = 3800
T = +S.
D =?

Solution:
First calculate elevation. The QNH is higher than 1013 :. pressure altitude is lower
than elevation, or put in reverse, elevation is higher than pressure altitude. I 024 -
1013 =II x 30 = 330 feet. 3800 + 330 =elevation of 4,130 feet.

Calculate density altitude. At 4000 feet (nearest 3800), !SA= IS- (2 x 4) = IS- 8 =
+ 7°C. Actual temperature is +Soc :. ISA-2. Colder than !SA means density altitude
is down, deduct temperature deviation height from pressure altitude. 2 x 120 =
240. 3800- 240 = density altitude of 3,S60 feet.

Conclusion

Knowing how to calculate density altitude is important because it is critical to


knowing how your aircraft will perform 'on the day' under actual atmospheric
conditions. Although you are required to show an ability to calculate it accurately
at this stage of your training, you should develop an ability to cany out a mental
check before each flight. When you set your altimeter subscale, note the value and
determine whether it is 'up' or 'down' on 1013 hPa. Have a look at the outside air
temperature gauge and, allowing a quick 2°C/l 000 feet, determine whether the
temperature is warmer or colder than !SA. If the subscale is down and the
temperature is up, be aware of reduced performance. And don't forget to make
allowance for humidity which, unfortunately, has to be an approximation at best. If
you consider the day to be humid, anticipate reduced performance from your
aircraft and engine.

There are three ways of determining density altitude:


• by simple calculation (as described above);
• by using the circular slide 111le on the back of the navigation computer (as
follows); and
• from performance charts (also as follows).

Principles of Flight Performance 17-15


You will have noted that examples 8 and 9 involved five elements: elevation,
QNH, pressure altitude, outside air temperature and density altitude. If we
abbreviate these elements it becomes a simple matter to calculate density
altitude and it also makes it possible to work out other element values when
three out of the five are known.
Let us abbreviate elevation withE, QNH with Q, pressure altitude with P, outside
air temperature with T and density altitude with D so that we see:
E
Q
p
T
D

Example 11. You are at an elevation of 4,300 feet, the QNH is 1021 hPa and the
outside air temperature is + 12'C. What is the density altitude?

Solution:
E = 4300
Q = 1021
p = ?
T = +12
D =?

E = 4300.
Q = 1021 (QNH high:. pressure altitude is lower).
P = 1021- 1013 = 8 x 30 = 240. 4300-240 = 4,060 feet= pressure altitude.
T = + 12. At 4000 feet (nearest 4060) !SA= 15- (2 x 4) = 15- 8 = + 7'C.
Actual temperature = + 12'C :. ISA+5 which is warmer:. density altitude is high,
add 5 x I 20 = 600 feet.
D = 4,060 + 600 = 4,660 feet = density altitude.

Example 12. You are at an elevation of 680 feet, the QNH is I 009 hPa and the
outside air temperature is +9'C. What is the density altitude?

Solution:
E = 680
Q = 1009
p =?
T = +9
D = ?
E = 680.
Q = I 009 (QNH low :. pressure altitude is higher).
P = I 0 I 3 - I 009 = 4 x 30 = I 20. 680 + I 20 = 800 feet = pressure altitude.
T = +9. At 1000 feet (nearest to 800), !SA= 15- (2 xI)= 15-2 = 13'C.
Actual temperature = +9 :. ISA-4 which is colder :. density altitude is low,
subtract 4 x I 20 = 480 feet.
D = 800 - 480 = 320 feet = density altitude.
As stated, provided three out of five elements are given, the remaining elements
can be calculated.

Example 13. You are at an elevation of 2,200 feet. You turn the altimeter
subscale to 1013 hPa and the instrument reads 2,500 feet. The outside air
temperature gauge reads +I 6'C. What is the QNH and what is the density
altitude?

17-14 Performance The Commercial Pilot Series


Conversely, if the ISA temperature deviation is - (minus), it indicates that the
temperature is colder than !SA. Remembering that colder temperatures give a
greater density normally found at lower altitudes, we can conclude that density
altitude is lower than pressure altitude, so we deduct the temperature deviation
height from pressure altitude.

The following examples make the direction in which to apply the correction clearer.

Example 7 The temperature on a given day is expressed as ISA+6. What is the


density altitude if the pressure altitude is 3,000 feet?

Step 1. ISA+6 means warmer than !SA by 6°. Thus density altitude must be higher.
Step 2. 6 x 120 = 720 feet.
Step 3. 3,000 + 720 = density altitude of 3, 720 feet.

Example 8. The temperature on a given day is expressed as ISA-8. What is the


density altitude if the pressure altitude is 2,500 feet?
Step 1. ISA-8 means colder than !SA by 8. Thus density altitude must be lower.
Step 2. 8 x 120 = 960 feet.
Step 3. 2,500 - 960 = density altitude of 1,540 feet.

Calculation of Density Altitude


We can now calculate density altitude if we know the QNH; the altitude or elevation
(which enables us to calculate pressure altitude), and the outside air temperature
(which enables us to calculate temperature deviation).

Example 9. You are at an elevation of 3,000 feet, the QNH is I 009 hPa and the
ambient temperature is + 1rc. What is the density altitude?

Step 1. Convert elevation to pressure altitude. 1013 - 1009 = 4 x 30 = 120 feet.


Pressure is less than 1013, thus add to elevation. 3,000 + 120 =pressure
altitude of 3,120 feet.
Step 2. Determine !SA temperature at pressure altitude. (3120 is nearest to 3000 ft).
15-(2x3) = 15-6 = +9°C.
Step 3. Compare !SA temperature ( +9) to actual temperature (+II) and determine
temperature deviation, !SA+ 2. Warmer temperature gives high density
altitude.
Step 4. Calculate deviation height, 2 x 120 = 240 feet and add to pressure altitude.
Step 5. 3,120 + 240 = density altitude of 3,360 feet.

Example 10. You are at an elevation of 5,700 feet, the QNH is 1025 hPa and the
ambient temperature is + 1°C. What is the density altitude?

Step 1. Convert elevation to pressure altitude. 1025 - 1013 = 12 x 30 = 360 feet.


Pressure is greater than 1013, thus subtract from elevation. 5, 700 - 360 =
pressure altitude of 5,340 feet.
Step 2. Determine !SA temperature at pressure altitude. (5,340 is nearest to 5,500
feet). 15- (2 x 5V2) = 15- 11 = +4°C.
Step 3. Compare ISA temperature ( +4) to actual temperature ( + 1) and determine
!SA deviation, ISA-3. Colder temperature gives low density altitude.
Step 4. Calculate deviation height, 3 x 120 = 360 feet and subtract from pressure
altitude.
Step 5. 5,340- 360 = density altitude of 4,980 feet.

Principles of Flight Performance 17-13


These steps are explained in more detail shortly. In broad terms, what we are
doing when we are calculating density altitude is first, making a correction to
actual altitude to take account of any variation in pressure from standard (we
find the pressure altitude); then we correct that pressure altitude to take account
of any variation in temperature to find the density altitude.

Calculation of Temperature Deviation


To calculate temperature deviation we must first establish the pressure altitude
(explained in the previous section) and determine what the ISA temperature is
at that level. We know that ISA sea level temperature is !Soc and that the
temperature lapse rate is 2°C/1 000 ft in practical terms. Thus by deducting 2°C
for every thousand feet (of pressure altitude) from 15°C, (or 1oc for every 500
feet) we can determine the ISA temperature at any pressure altitude. This
principle is simplified by the following formula:

!SA temperature at any pressure altitude = 15 - (2 x altitude*)


(*where altitude is in thousands of feet)

For example, the ISA temperature at 4,000 feet pressure altitude is 15- (2 x 4) =
15 - 8 = + 7°C. Or, !SA temperature at 2,500 feet pressure altitude is 15 - (2 x
2112) = 15 · 5 = + J0°C.

Example 5. What is the !SA temperature at a pressure altitude of 6,000 feet?

Step 1. 15- (2 x pressure altitude).


Step 2. 15- (2 x 6) = 15- 12 = +3°C = !SA temperature at 6,000 feet.

Example 6. What is the !SA temperature at 3,500 feet?

Step 1. 15- (2 x pressure altitude)


Step 2. 15- (2 x 3!12) = 15-7 =+8°C = ISA temperature at 3,500 feet.

Having obtained the !SA temperature at a given pressure altitude we must


compare it to the actual temperature at that level. For example, if the actual
temperature at 6,000 feet pressure altitude is + 12°C when we know that the ISA
value for that pressure altitude is +3°C (as calculated in example 5), we can
conclude that the temperature on this day is (12 - 3) = 9°C warmer than !SA.
This is expressed as ISA+9, where +9 is the temperature deviation.

Similarly, if the actual temperature at 3,500 feet pressure altitude is +6°C when
we know that the ISA value for that pressure altitude is +Soc (from example 6),
we can conclude that the temperature on this day is (6 - 8) = 2°C colder than
ISA. This is expressed as ISA-2, where -2 is the temperature deviation.

Calculations have shown that each degree temperature deviation is equivalent


to a variation in pressure which would be found in a change of height of about
120 feet. Thus ISA+9 corresponds to a variation in height (from standard) of (9 x
120) = 1080 feet and 1SA-2 corresponds to a variation of (2 x 120) = 240 feet.
The question now arises about which way the correction is applied and for this,
we must depend on our knowledge of the properties of the atmosphere. We
know that warmer temperatures cause air to be less dense and we also know
that less dense air is normally found at higher altitudes. Thus if the ISA
temperature deviation is +, indicating that the temperature is warmer than !SA,
and the air will be less dense and the density altitude is higher than pressure
altitude. In this case, we add the temperature deviation height to pressure
altitude.

17-12 Performance The Commercial Pilot Series


Fig. 17-4.

QNH 1013
1005 set

Example 1. Flying at 6500 It While maintaining altitude, set 1013


QNH of 1005 set. to obtain pressure altitude of 6, 740ft.

QNH 1013
1030 set

Example 2. Flying at 3,000 It, While maintaining altitude, set 1013


QNH of 1030 set. to obtain pressure altitude of 2490 ft.

Using the Altimeter to Determine Elevation or QNH

The approximate elevation (altitude) of a location can be determined by setting


the QNH in the subscale window and reading the altimeter indication.

Alternatively, if the elevation of a place is known, the approximate QNH can be


determined by turning the subscale setting knob until the altimeter indicates the
elevation. The approximate QNH will then be shown in the subscale window.

Density Altitude
Density altitude is the altitude in the lSA which has the same density as the
actual altitude with which we are concerned. !SA (or 'standard') density
conditions provide a yardstick by which aircraft performance can be measured.
Thus, in performance planning, if the actual atmospheric density conditions at a
given altitude are poorer than in the !SA, then aircraft performance at that
altitude will also be poorer. Conversely, if at a given altitude density conditions
are better than in the !SA (better than standard) the aircraft performance at that
altitude will be better. How much better or worse, is determined by how much
actual density conditions differ from standard. In other words, on what the
density altitude is at that actual altitude.

Density altitude is calculated as follows:


• determine the pressure altitude (as explained in the previous section);
• determine the temperature deviation from !SA at that pressure altitude;
• on the basis of I 20 feet per I oc variation in temperature, an adjustment to
pressure altitude is made to obtain density altitude.

Principles of Flight Performance 17-11


the aeroplane to be poorer than standard-it will perform as if it was at a higher
altitude. Hence under these conditions pressure altitudes will be above (higher
than) actual altitudes and we add the height variation. In this example, pressure
altitude is 1000 + 390 = 1390 ft.
If sea level pressure (QNH) is less than the standard, add the height
variation.

Example4

Also from example 2 where the airfield elevation is I ,000 ft, if the QNH is I 028
hPa, what is the pressure altitude of the airfield?
Step 1. Determine the pressure variation from !SA. 1028- 1013 = 15 hPa.
Step 2. Apply the formula. 15 (hPa) x 30 ft = 450ft.
Step 3. Apply the variation in height to actual altitude (or elevation). Since the
sea-level pressure (QNH 1028 hPa) is greater than standard we can expect the
performance of the aeroplane to be better than standard-it will perform as if it
was at a lower altitude. Hence under these conditions pressure altitudes will be
below (lower than) actual altitudes and we subtract the height variation. In this
example, pressure altitude is 1000 - 450 = 550 ft.

The effect of higher or lower sea level pressure (QNH) is felt more or less
throughout the atmosphere. Thus in example 2, had the aircraft been at an
altitude of 5,000 feet, the low QNH of 1000 hPa would produce a pressure
altitude of (5,000 + 390) = 5,390 feet, and the aircraft would perform as if it was
at that higher altitude.

Similarly, if in example 4, the aircraft had been at an altitude (or an elevation) of


7,000 feet, the high QNH of 1028 hPa would produce a pressure altitude of (7,000
- 450) = 6390 feet, and the aircraft would perform as if it was at that lower
altitude.

Interim Summary

Pressure altitude is a theoretical altitude used to gauge the effect of different sea
level pressures (QNH) on aircraft performance. The relationship between
pressure altitude and QNH is:
• When sea level pressure (QNH) is less than the standard 1013 hPa, pressure
altitude is higher than altitude (or elevation).
• When sea level pressure (QNH) is greater than the standard 1013 hPa,
pressure altitude is lower than altitude (or elevation).

NOTE. Remember, altitude (or elevation) is the actual height above sea level
indicated when the correct sea level pressure (QNH) is set on the altimeter
subscale. Thus altitude and pressure altitude will only be the same if the QNH
happens to be I 013 hPa.

Using the Altimeter to Determine Pressure Altitude.


While in flight it is possible to quickly dete1mine pressure altitude by simply
setting I 013 hPa in the altimeter subscale window. The altimeter will then read
pressure altitude. It is most important, however, not to forget to re-set the
subscale to the appropriate QNH as soon as pressure altitude has been
determined or when continuous use of QNH is a requirement. Fig. 17-4 provides
an illustration of the method. In each case, the picture on the right shows the
pressure altitude indication with I 013 hPa set briefly in the subscale window.

17-10 Performance The Commercial Pilot Series


conditions, we find the altitude in the theoretical standard atmosphere which has
the same air density as the actual altitude we are concerned with. We do this in
two steps:
• We first determine the pressure altitude which applies to the altitude (or
elevation) we need to know about.
• Then we apply a correction for any variation in temperature from !SA at that
pressure altitude, to find the density altitude.

When we know what the density altitude is, we will be able to determine how well
the aeroplane will perform under the prevailing atmospheric conditions. We now
cover the methods of calculating pressure altitude and density altitude.

Pressure Altitude
Pressure altitude is the altitude in the (theoretical) !SA which has the same
pressure as the real altitude with which we are concerned. Pressure altitude gives
us an indication of the effect of the prevailing atmospheric pressure on the density
of the air at any level. To illustrate pressure altitude in broad terms, consider the
following example:

Example 2
We are planning a take-off at an aerodrome with an elevation of I ,000 ft.
• If the actual pressure at I ,000 ft happens to be the same as at I ,000 ft in the
!SA, then our pressure altitude will be the same as our actual altitude. As far as
the pressure factor is concerned, there is no need to apply any compensation
for a variation in density. In this case, the QNH will be 1013 hPa.
• If the actual pressure happens to be higher than standard, then our pressure
altitude will be below I ,000 ft and we could expect (disregarding temperature
variation for the moment) to have a better performance than at I ,000 ft on a
standard day. In this case, the QNH will be more than I 0 I 3 hPa.
• Conversely, if the actual pressure is lower than standard, then our pressure
altitude will be above I ,000 ft and we could expect (again disregarding the
effect of temperature) to have a poorer performance than at I ,000 ft on a
standard day. In this case, the QNH will be less than 1013 hPa.

Determining the pressure altitude for any given real altitude (or elevation) is a
relatively simple three-step process. Given the prevailing QNH or the forecast QNH
for the location, we (I) determine the variation in actual pressure from standard
then (2) using the formula below, we convert this into a height variation. Finally,
(3) we apply the height variation to the actual altitude to obtain the pressure
altitude. In this final step we must remember the general process outlined in
example I, to get the right direction (up or down) to apply the height variation.
This becomes clearer in the following examples. Remember:
In the lower atmosphere, one hPa equals approximately 30 feet of height.

Example 3

From example 2 where the airfield elevation is I ,000 ft, if the QNH is 1000 hPa,
what is the pressure altitude at the airfield?
Step I. Determine the pressure variation from !SA. 1013- 1000 = 13 hPa.
Step 2. Apply the formula. I 3 (hPa) x 30 ft = 390 ft.
Step 3. Apply the variation in height to actual altitude (or elevation). Since the sea-
level pressure (QNH 1000 hPa) is less than ISA we can expect the performance of

Principles of Flight Performance 17-9


Other Factors
Apart from the foregoing, there are a number of other factors which affect take-
off and landing performance of which pilots should be cognisant. These other
factors will generally be listed as the conditions under which the performance
graphs or tables in the aeroplane manual have been calculated, and will include
such things as what equipment (e.g. air conditioner) is selected; the amount of
flap applied; the engine handling and braking technique; and the rotate, barrier
and approach speeds. If these conditions are different for the actual take-off or
landing, then performance will not be the same as calculated from the data.

Climb Gradient. It must also be remembered that most of the factors which
affect take-off performance will also affect the climb gradient after the take-off.
Thus, if the take-off distance is increased because of a high density altitude, or a
heavy-weight aeroplane, a tailwind component, or reduced power available,
these same factors will result in a poorer climb gradient being achieved in the
climb-out, with obstacle clearances being reduced perhaps to critical levels.

Calculating Pressure Altitude and Density Altitude


To be able to determine how an aeroplane will perform "on the day" we must
first know how much the existing pressure and temperature conditions vary
from the standard conditions of the ISA. Before moving on to the calculation of
pressure altitude and density altitude, we need to recap on the !SA, and commit
to memory some of its basic features.

The International Standard Atmosphere


The International Standard Atmosphere (!SA) provides the yardstick for
measuring the effects of changing density conditions on performance. A
detailed description of the !SA is given in Chapter 2 of this manual. The main
features of the !SA, sometimes referred to as 'standard conditions', which must
be remembered during performance planning are given below. (Note that the
following figures have been 'rounded off for practical use.)
• A standard sea-level pressure of 1013 hPa. (In some US-sourced
performance graphs and charts you may see the standard sea-level pressure
written as I 0 I 3 mb or 29·92"Hg. These measurements are equivalent to
1013 hPa.)
• A sea-level temperature of l5°C.
• A temperature lapse rate of 2°C per thousand feet up to the tropopause.

Two other figures which you need to remember during calculation of pressure
altitude and density altitude (explained shortly) are:
• Each change in pressure of one hPa from standard is equivalent to a change
in altitude (or height) of 30 ft. This is reasonably accurate in the lower
atmosphere (say, to I O,OOOft) but can also be used to give a 'fair' indication
above I O,OOOft. To make it easy to remember, we can write it as:
I hPa change = 30 ft
• Each I oc variation in temperature from standard causes a pressure change
equivalent to 120 ft of altitude. (Note; this is a different thing from the
standard temperature lapse rate of 2°C/I ,000 ft.)
l°C change = 120ft
The performance figures which the manufacturer provides for an aircraft are
based on the !SA. To determine how an aircraft will perform in actual

17-8 Performance The Commercial Pilot Series


Be aware of the dangers of "high, hot and humid"-meaning that when operating
at high altitude in hot and humid conditions, you will be encountering the worst
performance conditions because of low air density and you should be aware of
this particularly during take-off and landing. For example, at a 4,000ft elevation
aerodrome with a temperature of 27"C (ISA +20°C) a typical light-twin aeroplane
will require a take-off run which is about 60% longer than that required for a take-
off under standard conditions at sea level, just because of the reduction in air
density. And, if high humidity is also a factor, TODR will be even longer.

The International Standard Atmosphere (ISA) is used to provide a measure of


how much effect the ambient density conditions will have on performance "on the
day". This is considered in more detail shortly.

Weight
The weight of the aeroplane has a major effect on performance. Take-off
distances required are increased with increased weight because (a) acceleration
at take-off power is slower (with the increased mass) and (b) the safe lift-off lAS
(1·2 Vs) is increased. Similarly, landing distances are increased because, at the
higher weight, touch-down indicated airspeeds are higher and (with the increased
mass) a given amount of braking is less effective in slowing the aeroplane.

Wind
The amount of headwind (or tailwind component) clearly affect the runway
distance required for take-off or for landing. Wherever possible, pilots should use
take-off and landing directions which afford the greatest headwind component,
and should be very wary of accepting any but the lightest of tailwind components,
since take-off and landing distances are significantly increased with a tailwind.

Runway Slope
Runway slope in the direction of take-off or landing will affect take-off and landing
distance required (TODR and LDR). A downslope will decrease TODR but
increase LDR. Conversely, an upslope will increase TODR but decrease LDR.

Runway Suriace and Condition


The type of runway surface and its condition also affect the take-off and landing
distances which can be achieved. The take-off and landing performance data in
the aeroplane manual will normally be based on a paved level d1y runway. (Paved
can be taken to mean a relatively smooth concrete or bitumen surface.)
Generally, any other nmway surface-grass, metal (i.e. gravel), coral or rolled
earth-will result in an increase in TODR or LDR, because of the increase in wheel
drag during take-off, and the reduction of braking effectiveness on landing. Each
of the runway surface types mentioned has been ascribed a surface co1rection
factor in Part 135 which gives the increase in TODR or LDR which must be applied
when taking-off or landing on the particular type of surface. Surface correction
factors will be covered in more detail later.

Wet or Contaminated Runways


When landing on a wet or contaminated runway, there is a requirement to take
into account the longer landing distance required because of the reduction in
braking effectiveness. The specific requirement for the increase in LDA is given by
Part I 35.225, but for most small aeroplanes this will mean that, under these
conditions, LDA must be at least I 15% of the LDR calculated for dry conditions.

Principles of Flight Performance 17-7


Note that operators of medium aeroplanes under Part 125, and large aeroplanes
under Part 121, are required to take account of the same operational planning
factors.
In this part of the chapter we discuss the effects which the foregoing factors
have on take-off and landing performance

Air Density (Pressure Altitude and Ambient Temperature)


The density of the ambient air is a prime factor in determining how well an
aeroplane will perform. The factors affecting air density, its measurement and
general effect on performance, and the International Standard Atmosphere (ISA)
have been already outlined in Chapter 2. To recap:
• Air density affects both airframe (aerodynamic) and engine performance.
Consider the following example:
Assuming all other factors remain equal, if the ambient air density at a
certain aerodrome is lower (less dense) than normal, the take-off distance
required for any given aeroplane will be increased. The reason for this is
two-fold. First, to reach the lift-off lAS, theTAS (and groundspeed) must be
higher than normal and thus a longer ground run is required to accelerate to
and reach this higher TAS. In other words, more runway is required
because aerodynamic performance has been reduced. Secondly, the lower
air density results in a reduction in the power output (and thrust) of the
engine at the take-off rpm setting. Consequently, the power available for
take-off is reduced. So, in addition to having to accelerate to a higher TAS to
achieve lift-off, the thrust available to do it is reduced, resulting in a further
increase in the take-off distance required.
Landing performance is similarly affected in poor density conditions, but not
to the same degree. Touchdown speeds (for the same lAS) are higher, and
so, for the same amount of braking, longer landing distances are required.
• Air density cannot be measured directly. However, an accurate indication of
air density can be gained by measuring two related properties-pressure
and temperature.
• High pressures are an indication of good density conditions and therefore of
good performance. Conversely, low pressures are an indication of poor
density conditions and therefore of poorer performance. It follows then,
that as altitude is gained, both aerodynamic and engine performance must
reduce.
• Low temperatures are an indication of good density conditions and
therefore of good performance. Conversely, high temperatures are an
indication of poor density conditions and therefore of poor performance.
(Note here that since temperatures normally reduce with altitude, one
would expect the density and performance conditions to improve.
Unfortunately for aviation, the effect of the reduction in pressure is the
predominant effect on density as altitude is gained, and this is only partially
offset by the effect of reducing temperatures.)
• Humidity also affects air density and therefore performance. Humidity is
difficult to measure and under normal circumstances no allowance need be
made for the small effects of changing humidity. It is important to· note
however, that if performance has already been reduced by high
temperatures, the additional reduction in performance resulting from high
humidity may become significant and should be allowed for accordingly.

17-6 Performance The Commercial Pilot Series


Wet Runway

Wet, in relation to a runway, means a runway with sufficient moisture on its


surface to cause it to appear reflective but without significant areas of standing
water.

Contaminated Runway

Contaminated, in relation to a runway, means more than 25% of the runway


surface area within the required length and width is covered by surface water,
slush, or loose snow more than 3 millimetres in depth, or ice on any part of the
runway surface area.

Drift Down

Drift down means a gradual descent by an aeroplane with one engine


inoperative to an altitude at which it can comply with the one engine inoperative
enroute climb performance requirements.

Factors Affecting Take-Off and Landing Performance


An operator under Part 135 is required to ensure that for each take-off, the
weight of the aeroplane does not exceed the maximum take-off weight specified
in the flight manual; and the take-off distance does not exceed 85% of the take-
off run available. When calculating the take-off weight and distance, the
operator is required to take account of:
I. the take-off run available;
2. the weight of the aeroplane at the commencement of the take-off run;
3. the pressure altitude of the aerodrome;
4. the ambient temperature at the aerodrome;
5. the type of runway surface and the runway surface condition;
6. the runway slope in the direction of take-off; and
7. not more than 50% of the reported headwind component or not less than
ISO% of the reported tailwind component.

Similarly, operators are required to ensure that for landing, the weight of the
aeroplane does not exceed the landing weight specified in the aeroplane
manual, and that the aeroplane will be able to make a full-stop landing from 50 ft
above the landing threshold within 85% of the landing distance available. When
calculating the landing weight and distance, the operator is required to take
account of:

I. aerodrome elevation;
2. ambient temperature at the aerodrome;
3. the type of runway surface and the runway surface condition;
4. the runway slope in the direction of landing; and
5. not more than 50% of the reported headwind component or not less than
ISO% of the reported tailwind component.

The foregoing factors apply to landing on a dry runway. For landing on a wet or
contaminated runway the landing distance must be increased, in effect, by 15%.

Principles of Flight Performance 17-5


Net flight path means the gross flight path reduced by specified margins. These
margins, which will be specified in the Rules, make allowance for the reduced
performance which could be expected in a real emergency situation in
unfavourable conditions (e.g severe turbulence).

Landing Distance Available (LDA)

Landing distance available (LDA) means the length of the runway that is
declared by the aerodrome operator as available and suitable for the ground run
of an aeroplane.

The landing distance available starts at the landing threshold and in many cases
corresponds to the physical length of the runway. However, the landing
threshold may be displaced from the end of the runway when it is considered
necessary to make a corresponding displacement of the approach area and
surface for reason of obstructions in the approach path to the runway.

For example, grass runway 20 at Nelson is 731 metres long but the threshold is
displaced by 84 metres making the available landing distance only 64 7 metres.

NOTE: The landing threshold means the beginning of that portion of the
runway declared usable for landing.

Landing Distance Required (LDR)

The Landing Distance Required (LDR) is the horizontal distance measured


from from a point 50 feet above the runway threshold to the point where the
aeroplane can be brought to a complete stop with maximum braking. In
calculating the landing distance required, it is assumed that the approach to the
runway is steady and the speed at the 50 ft point (screen height) over the
threshold is not less than 1·3 Vs or the speed published in the aeroplane flight
manual, whichever is the greater.

landing
K\ _ threshold
~~~·. complete
0
.
--
''
I stop

screen heigh;-!~ '


e.g. 50ft
~
landing distance required -------1
1<4------------ landing distance available -----------~

Fig. 17-3. Landing distance available (LOA) and landing distance required (LOR).

The LOR varies for each landing depending on a number of factors. These are
discussed in more detail later.

Dry Runway

Dry, in relation to a runway, means a runway that is neither wet nor


contaminated, and includes a paved runway that has been specially prepared
with grooves or a porous pavement to retain effectively dry-braking action even
when liquid moisture is present.

17-4 Performance The Commercial Pilot Series


The accelerate-stop distance available (ASDA) is promulgated as part of the
runway information for appropriate aerodromes. The accelerate-stop distance
required (ASDR) under varying conditions can be calculated from tables or
graphs given in the aeroplane flight manual.

NOTE: Operators under Part 135 are not currently required to take account of
accelerate-stop distance during· take-off performance planning.

Gradient of Climb

Gradient of climb means the ratio of height gained over horizontal distance
travelled, expressed as a percentage. Note that a percentage climb gradient is
not a climb angle. (Climb angle can be obtained by finding the tangent of the
gradient percentage value.)

If the height gained is 500 feet when you cover a horizontal distance of 8000 feet:
height gained 500
climb gradient = = - - = 0·063
distance travelled 8000
In other words, for each foot travelled horizontally you climb 0.063 ft.

This can be converted into a percentage by multiplying by I 00/1 so that:


500 100
climb gradient = X = 6·3%
8000

Height gained is a function of rate of climb whilst horizontal distance covered is


a function of speed. Owing to a fortuitous combination of conversion factors, a
very close approximation of climb gradient is given by:
rate of climb (in fVmin)
climb gradient = %
speed (in knots)
If you use true airspeed you obtain the climb gradient with respect to the
airmass and using ground speed gives you the climb gradient with respect to the
ground.

Example I. You climb your aircraft at an lAS of 95 kt at sea level under !SA
conditions. The headwind component is 18 kt and the rate of climb (ROC) is
600 ft/min. What is the climb gradient (a) with respect to the airmass and (b)
with respect to the ground?
ROC 600
(a) climb gradient = % = 6·3%
TAS 95

ROC 600
(b) climb gradient = % = = 7·8%
GS 77
This confirms that in a headwind the gradient of climb increases which
improves obstacle clearance and, in a tailwind the gradient decreases which
reduces obstacle clearance.

Gross and Net Flight Paths

Gross flight path means the flight path it is assumed the aeroplane will follow
when flown in a particular configuration in accordance with specified
procedures. The flight path is established from the aeroplane's certification
performance data and can be accepted as the average fleet performance for the
aeroplane type.

Principles of Flight Performance 17-3


Take-off Distance Available (TODA)

The take-off distance available (TODA) is defined as the length of take-off run
available plus the length of any cleaJWay.
Notes: I. The take-off run available (TORA) means the length of the runway
declared by the aerodrome operator as available and suitable for the ground run
of an aeroplane taking-off.
2. A cleatway means a defined rectangular area on the ground or water
at the departure end of the runway selected or prepared as a suitable area over
which an aeroplane may make a portion of its initial climb to a specified height.
3. Runway lengths for aerodromes in New Zealand are published in the
AIPNZ operational data for the aerodrome. These may be stated simply as "take-
off Distance" and "LOG DIST" and can be taken as having the same meaning as
we have defined for TOOA and LOA respectively. These distances are stated for
a given obstacle clearance slope and may be different for take off (or landing)
from opposite directions on a given runway because this obstacle clearance
slope intersects the runway surface at a different distance from each end.

,.__ _ _ _ _ _ _ take-off run available (TORA) - - - - - - - - - - 1 > ! - ' f - - - clearway ____.:


I
I
~----------- take-off distance available (TODA) -----------1~
'
Fig. 17-2. Take-off distance available (TODA).

IMPORTANT NOTE: Under CAR Part 135, the take-off distance required (TOOR)
must not exceed 85% of the take-off run available (TORA). This means that
operators under Part 135 may not plan to use a cleaJWay for the initial take-off
climb to screen height. As we explain later, this 85% factor is incorporated in the
'P' charts.
Accelerate-stop Distance

The accelerate-stop distance available (ASDA) means the distance specified


by the appropriate authority as being the effective length available for use by an
aeroplane executing an abandoned (or rejected) take-off. The effective length
may include a stopway.
A stopway is a defined rectangular area on the ground at the end of a runway in
the direction of the take-off which is designated and prepared by the competent
authority as a suitable area in which an aircraft can be stopped in the case of an
interrupted take-off.
The longer the declared accelerate-stop distance:
• the higher the speed beyond which an aircraft can be brought to a halt in
the case of an abandoned take-off;
• the greater the allowable take-off weight.
A stopway does not necessarily have to be a prepared sealed or level grass
surface, but a reasonably flat area suitable for braking and bringing an aeroplane
to stop in an emergency. Note the difference between a stopway and a
cleaJWay. A stopway must be on the ground, whereas a cleaJWay is provided as
an obstacle-free zone over which the aeroplane can fly as it is reaching screen
height and can thus be either ground or water.

17-2 Performance The Commercial Pilot Series


Performance
Introduction

How well an aeroplane performs at any given time depends on a number of


factors. The prime factor which always has an effect on the performance of any
aircraft is the ambient (surrounding) air density. In addition to air density, there
will also be a number of other factors which will have an effect depending on
the phase of flight. A proper account must be taken of these factors in pre-flight
planning, and during flight, to ensure that the aeroplane will be able to take off
safely, be flown efficiently and safely during its task, and land again safely.

This chapter covers the syllabus requirements of Part II Performance of the CPL
(Aeroplane) Principles of Flight and Performance examination. This
encompasses the performance planning requirements for take-off, landing, and
en-route flight with the critical engine inoperative for a twin-engined aeroplane
being operated VFR under Part 135 of the NZ Civil Aviation Rules.

Hence, this chapter covers:


• the relevant performance terms and definitions.
• the factors affecting take-off and landing performance;
• the calculation of pressure altitude and density altitude;
• determination of take-off and landing performance using aeroplane flight
manual data and performance graphs (P-charts);
• use of single-engine se!Vice ceiling graphs.

Terms and Definitions


An understanding of the following terms and definitions relating to performance
planning is required.

Take-off Distance Required (TODR)

The take-off distance required (TODR) is the distance required to take-off


from a standing start at maximum take-off power and reach a screen height
(usually 50ft) above the runway at the take off safety speed.
Notes: I. Take-off safety speed (TOSS) is a speed which gives an adequate
margin above the stalling speed. It must be not less than 1·2 Vs in the take-off
configuration. TOSS will normally be incorporated in the take-off climb speed
given in the flight manual for the type.
2. Screen height. The screen height for larger aeroplanes and for
commuter operations may be 35ft (or lower under some circumstances).

~
standing
start
maximum take~off power ~!::reen
e.g. 50ft
height
take~off distance required

Fig. 17-1. Take-off distance required (TODR).

The TODR varies for each flight depending on a number of factors. These are
discussed in more detail later.

Principles of Flight Performance 17-1


20. In a divergent duct, the velocity of supersonic flow is ............................. and
the pressure is ............................... .

21. Wing shockwaves form (before/after) the bow shockwave forms.

22. When the upper shockwave forms on the wing and strengthens, CL
(increases/decreases). This is called the .......................... stall.

23. Wave drag is made up from two sources-........................ drag and boundary
layer ............................. .

24. When the wing shockwaves form and strengthen C0 (increases/decreases).

25. Once the wing shockwaves move to the trailing edge and the bow wave
forms, the Co (increases/decreases) again.

26. As most aircraft enter the transonic range they experience a nose (up/down)
change of trim. This is called Mach ................ .

27. To solve the change of trim problem, aircraft may be fitted with a ................. .
............................ which provides a nose (up/down) control input.

28. If the Vc ratio of a given wing is reduced, the Co will be (increased/reduced)


and the CL will be (increased/reduced).

29. A supercritical aerofoil has:


(a) a very low Vc ratio; or
(b) a relatively high Vc ratio, a flattened upper surface, and reflex camber on
the lower rear surface.

30. When the leading edges of the wings are swept back, Merit is at (a higher/a
lower/the same) flight Mach number.

31. List three disadvantages of sweepback.

32. The main purpose of applying area rule is to reduce the


transonic ........................................ .

33. The common feature of all successful supersonic wing planforms is that the
angle of leading edge sweep is within the ....................................... .

16-32 High Speed Flight The Commercial Pilot Series


Review 16
I. The speed of sound depends only on the ................................. of the air in
which it is travelling.

2. Mach number is the ratio between the TAS of an aircraft (or of an airflow)
and the local speed of ............................. .

3. The Mach number of an aircraft travelling at % the speed of sound is


expressed as M ........... , and at I v, times the speed of sound as M ............... .

4. C1itical Mach number (Merit) is the flight Mach number at which the speed
of the flow at any point around the aircraft has reached ...................... .

5. Accelerating through Mdet, the bow shockwave becomes (detached from/


attached to) the leading edge.

6. The transonic speed range falls between ................. and .................... . This is
approximately between M ...... and M ....... .

7. When an aircraft is fully supersonic, all ...................... Mach numbers will be


greater than M 1·0.

8. When a point source of sound is moving at the speed of sound, the sound
waves will coalesce toward the front and form a Mach ............... (or
Mach ................... ).

9. The Mach wave at Mach 1·0 is (normaVoblique).

I 0. If a sound source is travelling faster than M 1·0, a ................... cone will be


formed, and its angle (the wave angle) will be (normaVoblique).

II. The faster the source is travelling above M I ·0, the more (normaVangled
back) the wave angle.

12. A shockwave is a compression wave of (higher/lower) amplitude than a


Mach wave.

13. The velocity of the airflow is (increased/decreased) when passing through a


shockwave.

14. (Compression/expansion) takes place suddenly across a shockwave, i.e. the


pressure, temperature and density all ....................................... .

15. The speed of the flow across a normal shockwave is always reduced
below ....... and its direction is (changed/unchanged).

16. The speed of the flow across an oblique shockwave (may be/is not)
reduced below M 1·0 and its direction is (changed/unchanged).

17. When supersonic flow meets a compressive corner, a (shockwave/


expansion wave) is formed.

18. When supersonic flow is able to turn away from itself (or diverge) it does
this through a series of (shockwaves/expansion lines or fans).

19. In a convergent duct, the velocity of subsonic flow is .............................. and


the pressure is ................................ .

Principles of Flight High Speed Flight 16-31


mach cone
/ '../
/
/
"'
/
/
' '-,

Swept wing (e.g. Lightning) Delta (e.g. Concorde) Variable geometry (e.g. F111)

Fig. 16-44. Examples of supersonic wing planform shapes.

Only the delta planform shape has been used for supersonic transports (the now
obsolete Concorde and Tu 144). This planform shape offers the following
advantages:
• It is the simplest to constmct and, with a long chord, a low Vc ratio can be
used but still leave sufficient depth inboard for structural strength, storage of
fuel, etc.
• By comparison with the swept-wing shape, a delta wing can be made more
rigid and the problems of flexing and twisting under load avoided.
• With a low aspect ratio, the tendency toward tip stalling can be reduced.
• For a given span, it has an increased wing area, and therefore a reduced
wing loading which gives a better high altitude capability and, other factors
remaining equal, a lower stall speed. Against this, skin friction drag is
increased.
Particularly when it has a high degree of leading-edge sweepback, a delta wing
is prone to the formation of large ram's-horn type vortices above the wing at high
angles of attack. However, as with the Concorde, with careful design of the
leading edges, these vortices can be controlled and turned to advantage in
generating lift. The penalty paid is in the disadvantage of requiring long
undercarriages and, in Concorde's case, the fitting of a 'droop-snoot' cockpit
shield to provide sufficient visibility for landing.

16-30 High Speed Flight The Commercial Pilot Series


Although the double wedge is efficient in supersonic flight, because of the sharp
corners involved, it is not a practical proposition for subsonic flight (which
includes, of course, the important business of the approach and landing).

The symmetrical bi-convex section provides a good compromise between the


conflicting requirements of efficient supersonic and subsonic flight (Fig. 16-42).
It can be built with a low t/c ratio for high-speed flight and its curved section
remains suitable for subsonic flight. In supersonic flight, at an angle of attack of
the 'surface angle' (half the angle of the nose) similar flow conditions to the
double-wedge aerofoil can be achieved, with a minimum of two shockwaves
and a smooth expansion of the flow across the upper and lower surfaces.

CP
M>1.0

expansion waves

Fig. 16-42. Flow pattern and pressure distribution of a bi-convex section at optimum a.
The only practical example of a supersonic transport aircraft wing section in use
until recent times was that of the Concorde. The Concorde section (Fig. 16-43)
had a ve1y low t/c ratio of 2·9% at the wingroots, tapering to 2·1% outboard of the
engines, which provided for low wave drag at the design cruising speed in the
region of Mach 2·2. The leading edge was 'permanently drooped' with a reverse
camber which was carefully designed to provide for control of the leading-edge
separation bubble and the leading-edge (ram's-horn type) vortex which
developed as a result. In the Concorde design, the leading-edge vortex was
turned to advantage in improving the lift characteristics at low speed/high angles
of attack.

Fig. 16-43. The shape of the


Concorde wing section.

Supersonic Planform Shapes


A number of planform shapes can be used successfully for supersonic flight.
The more common are swept wing, various delta shapes, and variable geometry
(i.e. 'swing-wing')-Fig. 16-44 refers. In the latter case, the wings are swung
fmward to gain the advantages of a straight, high aspect-ratio wing for slow
flight, and swept back to gain the advantages of sweepback/delta shape (when
combined vvith the tailplane) for high-speed and supersonic flight.

Most, if not all, successful supersonic planforrn shapes have one common
feature. That is, the angle of sweepback of the leading edge is designed to be
within the Mach cone which is formed by the front shockwave (which may be
from the nose of the aircraft or the fmward inboard leading edge at the
wingroot-see Fig. 16-44). When this is the case, the airflow reaching
successive outboard sections of the wings will have been slowed by the
shockwaves from successive inboard sections. As a result, wave drag is reduced
and the performance of the wing is improved.

Principles of Flight High Speed Flight 16-29


c-= I II
I I I II
1 I wings I 11 I !wings II
I I -...,. I I I 11 1~11111
!~A-" tail 1 • -.,.tail
I~ ~~

without area rule with area rule applied

F1g. 16-39. Area rule.

Area ruling has no effect on the Fig. 16-40 without area rule
speed at which the shockwaves with area rule
Co
form (i.e. Merit). Its main purpose
is to reduce the amount of drag
generated as a result of the
shockwaves forming. With area
ruling, the transonic increase in
1 transonic
C0 can be typically halved, as Merit I Mach
' drag rise number
shown in Fig. 16-40. Mcdr

Area ruling may be applied to


reduce the Co in supersonic flight, but the design can only be optimized for one
flight Mach number.

Design for Supersonic Flight


Wing Sections
For efficient supersonic flight, the thickness/chord (tic) ratio must be as low as
possible. The most efficient supersonic wing section is thus a flat plate.
Although it has been used successfully for guided missiles, it is totally impractical
for use as an aircraft wing, which must have depth for structural strength, fuel
storage, controls etc.

The next in order as an optimum shape for fully supersonic flight, is the
symmetrical double-wedge shape. Its advantage for supersonic flight, by
comparison with what we have come to consider as a normal cambered
aero foil section, is that some of the shockwaves can be avoided and the lift/drag
ratio improved accordingly. Fig. 16-41 shows the flow pattern which will exist if
this type of section is flown supersonically at an angle of attack equal to the bi-
sector of the wedge (the most efficient angle), when the number of shockwaves
is reduced to two.

expansion fan
shockwave

M>1.0

shockwave

Fig. 16-41. Flow pattern and pressure distribution of double-wedge aerofoil at optimum a.

16-28 High Speed Flight The Commercial Pilot Series


A further cause of pitch-up in a swept-wing aircraft which is unrelated to the stall
is the washout effect on the outer wings when they are flexed upward under an
increased load-for example, when the aircraft enters a turn. Fig. 16-38 shows
that when a swept wing is flexed upward those points which are generally at
right angles to the spars (along the lines A- AI, B- Bl etc) are lifted by an equal
amount. Since the outer lines are flexed upward a greater distance than the
inner ones, this means that point AI is lifted higher than B which in effect
reduces the angle of attack of the section which lies on the line B- AI. Thus the
greater the upward flexing, the greater the wash-out effect and the more the
wing CP moves forward and inward to cause a pitch-up. A certain amount of
pitch-up due to flexure is acceptable provided sufficient longitudinal control is
available to cope with it. Othe1wise, the designer must take appropriate
measures to control the amount and nature of the flexing under load.

Fig. 16-38. The washout effect of airflow

j
wing flexure.

when wing flexes upward,


A ~ A1 are lifted higher than B ~ B1
resulting in reduced angle of attack
along the line B - A1, i.e. washout
toward the wingtips

Deep Stall

If pitch-up is allowed to occur at high angles of attack it is possible for the aircraft
to be rotated well beyond the stalling angle. When this happens on an aircraft
with a high T-tail, it is also possible for the tailplane/elevator to become
immersed in the turbulent wake behind the wing, rendering the elevator
ineffective in achieving recovery to a lower angle of attack. For this reason, all
swept wing aircraft with this tail configuration are fitted with 'stick pushers'
which provide an uncommanded nose-down control input when the angle of
attack is becoming dangerously high.

Area Rule
'Area rule' is a design philosophy which states, in effect, that to achieve the
minimum transonic drag rise, the cross sectional area of the whole aircraft
should increase and decrease smoothly from nose to tail. In the 1950's-when it
was difficult with the engines then available to obtain sufficient thrust to
overcome the transonic drag rise and achieve supersonic flight-it was
discovered that 'blending' the cross sectional area of the fuselage, wings and tail
section solved the problem. If the cross-sectional shape of an aircraft increases
or decreases suddenly, there is a tendency for strong shockwaves to form,
resulting in a large increase in wave drag. With an 'area-ruled' aircraft, the
shockwaves which form are more dispersed and less intense, giving the
advantage of reduced transonic drag.

Fig. 16-39 illustrates the principle of area ruling. The diagram shows the use of
'thickening' or 'waisting' the fuselage as necessary to achieve a 'coke-bottle'
shape which accommodates the increase in cross sectional area brought on by
the wings and tail section. Other methods which have been used are flattened
fuselage sides combined with highly swept wings, and the fitting of pods to the
trailing edges of the wings.

Principles of Flight High Speed Flight 16-27


vortex generators-+~~.,
Fig. 16-36.

Leading-edge slots. As described in Chapter 6, leading edge slots also have the
effect of re-energizing the boundary layer. They are often combined with
leading-edge flap which is beneficial in reducing any tendency toward leading-
edge separation as well as lowering the geometric stalling angle.
Boundary layer blowing and suction. With blowing, high-energy air is injected
into the boundary layer to increase its energy. With suction, the weakened
boundary layer is drawn off and replaced with faster-moving air from above.
These forms of boundary layer control are not often used because of the high
demand placed on engine power for their operation-and the subsequent
reduction of reserve power available for use in emergency, for example, during
the approach to landing.
Leading-edge 'sawtooth' extension, or notch. These measures have been
used in the past on some fighter-type aircraft to generate a smaller inboard
vortex and retard the development of a larger ram's horn vortex from the
wingtip.
Tip Stalling and Pitch-up

The migration of the boundary layer, pooling and thickening at the wingtips on
swept-wing aircraft also means that they are much more susceptible to wingtip
stalling by comparison with their straight-wing counterparts. In addition, when a
swept-wing aircraft does suffer a wingtip stall, there is a marked tendency for the
nose of the aircraft to pitch up (and not down, as in the case of a straight wing).
The reason for the pitch-up is that, as the wing is stalling progressively from the
tips, its effective centre of pressure moves forward-since the inboard forward
wing sections will remain unstalled (Fig. 16-37 refers). By reducing the normal
nose-down wing pitching moment, this will pitch the nose up and, if corrective
action is not promptly taken, it will tend to take the aircraft rapidly into a fully
developed stall. The use of washout is common on swept wing aircraft to
reduce the likelihood of tip stalling and all of the other measures which have
been mentioned earlier involving the boundary layer control and the reduction
of induced drag are also effective in helping to prevent the wingtip stall.

wing pitching moment


(nose-down) reduced
= nose-up pitch
..,
CG
wing CP moves
forward and in

mean CP position
moves forward

UNSTALLEO WINGTIPS STALLED

Fig. 16-37. A wingtip stall in a swept-wing aircraft will cause pitch-up.

16-26 High Speed Flight The Commercial Pilot Series


Lower CLmax/Higher Stalling Angle

By comparison with a straight wing with the same section, wing sweep will
decrease the CLmax and give a higher geometric stalling angle (Fig. I6-34 refers).
Approach and landing speeds will therefore tend to be higher, thus necessitating
extensive use of high lift devices to keep these speeds reasonable. In addition,
to give better visibility and avoid the use of inordinately long undercarriages,
methods of reducing the body angle on approach and landing (e.g. leading edge
flaps and slats) must also normally be employed.

High Drag at High Angles of Attack

In swept-wing aircraft, because of the 'angled back' nature of the pressure


distribution over the wing, there is a tendency for the boundary layer to drift
outward along the wing. This results in a 'pooling' and thickening of the
boundary layer at the wingtips making the flow more susceptible to separation in
this area.

At high angles of attack, separated flow from the region of the wingtips
combines with the normal wingtip vorticity and encourages the formation of
large vortices which have their origin at the leading edge. This type of vortex is
called a 'ram's-horn' vortex and it is most pronounced on wings which are highly
swept with sharp leading edges (e.g. as seen on some fighter-type aircraft). Fig.
I 6-35 refers.

'ram's horn' vortex


on highly-swept
wing

enlarged vortex

Fig. 16-35. Swept wings: enlarged vortices and increased induced drag.

The larger vortices generated by swept wings leads to greater induced drag at
high angles of attack by comparison with a straight wing. This is the reason for
the higher geometric stalling angle shown in Fig. I 6-34. Design measures used
to reduce this drag are generally aimed at reducing the vorticity by checking the
bounda1y layer outflow and 're-energizing' it to delay separation. They include:

Wing fences. (Fig. I 6-36 refers). Wing fences provide a physical barrier to the
boundary layer outflow, thus reduce pooling, thickening and flow separation.

Vortex generators. As also illustrated at Fig. I 6-36, vortex generators are sets of
small vertical fins placed at an angle of incidence on the top of the wing usually
at about the point of maximum thickness. They generate a series of small high-
speed vortices which drag down faster-moving air from above the wing to mix
with, and re-energize the boundary layer. Vortex generators are sometimes
fitted ahead of control surfaces (e.g. ailerons) to increase their effectiveness (by
speeding up and strengthening the boundary layer flow over the control). They
can also be effective at high speeds in reducing shock-induced bounda1y layer
separation and the effects of the upper shockwave.

Principles of Flight High Speed Flight 16-25


35. Given aerodrome elevation 1340 ft, QNH I 008 hPa, ambient temperature
4°C, what is the aerodrome density altitude?
A 2450 ft;
B 110ft;
c 530ft;
D 2270 ft.

36. The temperature at pressure altitude 8000 ft is expressed as !SA-l (!SA


minus l°C). What is the actual temperature at that altitude?
A -l°C;
B -2°C;
c ooc;
o we.

37. You are planning a take-off on runway II. The wind is 150°M 15 gusting to
23 knots. Using the wind component graph on page 17-21 determine the
average headwind component, and the maximum crosswind component
which could be expected on this take-off.
A average headwind 17 knots; maximum crosswind 15 knots
B average headwind II knots; maximum crosswind I 0 knots;
C average headwind 15 knots; maximum crosswind 18 knots;
D average headwind II knots; maximum crosswind 15 knots.

38. As part of an air transport operation under Part 135, you are planning a
take-off on a grass strip which has a 2o/o downslope. The aircraft flight
manual data indicates that you will have a take-off distance to 50 ft of 490
metres. Using the extract from CAR Part 135 given at page 17-23, calculate
the corrected take-off distance required.
A 503 metres;
B 614 metres;
C 520 metres;
D 530 metres.

39. Using the take-off graph on page I 7-25, calculate the take-off distance
required under the following conditions; pressure altitude 600 ft, ambient
temperature +8°C, take-off weight 2550 kg, hard surface, lo/o upslope, 14
knots headwind component.
A 760 metres;
B 690 metres;
C 880 metres;
D 830 metres.

Principles of Flight Appendix 1-9


40. Again using the take-off graph on page 17-25, calculate the maximum
take-off weight permissible under the following conditions; pressure
altitude 1700 ft, ambient temperature lSA +3, sealed surface, 2%
downslope, 7 knots headwind component, take-off distance available 720
metres.
A 2500 kg;
B 2270 kg;
c 2460kg;
D 2360 kg.

41. Using the landing graph on page 17-33, calculate the landing distance
required under the following conditions; pressure altitude 150 ft, ambient
temperature 2rc, landing weight 2320 kg, grass strip, I o/o upslope, 18
knots headwind component.
A 492 metres;
B 380 metres;
C 468 metres;
D 500 metres.

42. Again using the landing graph on page 17-33, calculate the maximum
permissible landing weight under the following conditions; pressure
altitude 1700 ft, temperature lSA-4, grass strip, 2% upslope, 12 knots
headwind component, landing distance available 460 metres.
A 2370 kg;
B 2250 kg;
c 2290 kg;
D 1980kg.
43. Using the take-off chart on page 17-37, calculate the take-off distance
required for an air transport operation under the following conditions;
pressure altitude 900 ft, ambient temperature I8°C, take-off weight 1560
kg, paved level runway, II knots headwind component.
A 950 metres;
B 750 metres;
C 910 metres;
D 570 metres;

44. Using the landing chart on page 17-38, calculate the landing distance
required for an air transport operation under the following conditions;
pressure altitude 300 ft, landing weight 1490 kg, paved runway, I o/o
downslope, light and variable wind.
A 570 metres;
B 485 metres;
C 450 metres;
D 530 metres.

1-10 Appendix The Commercial Pilot Series


45. Using the single-engine service ceiling graph on page 17-40, calculate the
maximum all up weight at which the aeroplane can maintain a pressure
altitude of 17,000 ft (FL170) in the event of an engine failure, if the forecast
temperatures are as follows; FL120 minus 2oC (M2); FL140 (M8); FL160
(M14).
A 6700 lb;
B 6900 lb;
c 7100lb;
D 7300 lb.

Principles of Flight Appendix 1-11


1-12 Appendix The Commercial Pilot Series
Appendix 2
Answers to Review Questions
and Sample Examination

REVIEW1 26. You should have a vector REVIEW3


I. newton; (N); 9·8IN. to the 1ight. I. static.
2. watt; (W); horsepower. 27. distance. 2. dynamic.
3. feet; nautical miles; 28. rate; 3. freestream.
knots. work done/time taken; dynamic
4.
4. magnitude; force x distance/time. p =density
magnitude and direction. 29. motion; KE = 1/2m\f2. V = velocity (TAS).
5. direction. 30. potential; kinetic; speed. 5. static + dynamic.
6. velocity. REV/EW2
6. dynamic; knots.
7. kilograms; I. 1·225; I ,000.
7. !AS.
amount of matter. 2. pressure, temperature.
8. lAS.
8. inertia. 3. hectoPascals (hPa).
9. aerofoils.
9. inertia, inertia. 4. higher . 10. refer to Fig. 3-6.
10. accelertion. 5. lower
II. is.
II. rest; motion; force. 6. (a) 1013·2 hPa;
12. refer to Fig. 3-7.
12. mass x acceleration. (b) J5°C.
13. constant.
13. force; gravitational. 7. (a) I hPa every 30ft;
14. (a) decreased
14. 9·81. (b) I·98°C (2°C); 36,090ft.
(b) increased
15. three. 8. (b).
IS. streamlined.
16. momentum; 9. (a) 75%
16. turbulent (or separated).
mass x velocity. (b) 50%
I 7. refer to Figs 3-11 a,b,c.
17. action; reaction. (c) 25%
18. stagnation.
18. toward; centripetal. 10. 30,000 ft; 60,000 ft.
19. critical or stalling.
19. Wv2/gr. II. (a) low; poorer
20. exceed.
20.

~
(b) high; better
21. total reaction;
(c) high; better
centre of pressure
21.

22. c/b.
i2J (d) low; poorer
12. high; low.
13. low; high.
22. forward; rearward.
23. little if any.
24. lift; drag.
23. 2,000 kg-mm. 14. !SA.
24. 30 Newton-metres. IS. dry
25. c. not accelerating.

Principles of Flight Appendix 2-1


REVIEW4 15. fmward; increases. REVJEW7
I. refer to page 4-2. 16. separation; wake. I. power plus attitude.
2. angle of attack; airspeed. 17. 5%. 2. refer to Fig. 7-2.
3. capability 18. length; depth. 3. (a) pitch
4. refer to Fig. 4-1. 19. all factors listed affect (b) roll
5. (a) low; low form drag. (c) yaw
(b) high; high 20. interference. 4. (a) pitch
6. (a) lower 21. lift. (b) roll
(b) lower 22. effective. flightpath. (c) yaw
(c) increased. 23. significant. less. 5. yaw.
7. suddenly. 24. high. 6. roll.
8. greater. 25. inversely propotional. 7. airspeed.
9. lower; loss. 26. low; high. 8. opposite direction to.
10. low; high. 27. washout; taper; wing 9. frise ailerons; differential
fences; winglets; drooping ailerons;coupling aileron
II. effective. wingtips; tip tanks. to rudder; use of spoilers.
12. decreases. 28. refer to Fig. 5-16. 10. effective; higher.
13. wingspan; chord. 29. Drag = Co '/2 pV2 S. II. (a) increases
14. less.
30. refer to Fig. 5-17. (b) rise
15. higher; lower.
31. lift; drag. (c) left
REVIEWS
32. refer to Fig. 5-18. 12. opposite direction to.
I. parallel and opposite to.
33. induced drag; 13. centre of pressure.
2. profile drag (skin friction induced drag, form, skin-
and form drag) and friction, and interference 14. inset hinges; horn
drag. balance;
interference drag. balance tabs.
3. boundary layer. REVJEW6
15. too light.
4. laminar flow and I. camber.
16. same direction as.
tubulent flow. 2. (a) increased; lower
17. closer to.
5. thicker; more. (b) increased
REVJEWB
6. transition. (c) reduced
I. exceeds.
7. maximum thickness. 3. reduced.
2. fixed.
8. adverse. 4. refer to Figs.6-8 to 6-10.
3. reducing airspeed;
9. forward; increases. 5. (a) delay reducing control
(b) reduce effectiveness; buffet;sink.
10. all factors listed affect
skin-friction drag (c) increase 4. down.
II. form. 6. higher. 5. 15 - 20%; rearward.
12. separation. 7. disturbing; decreases; 6. reduced.
13. reverse. increases 7. (a)
14. thickens; forward; 8. square root of the 'g'.
increases 9. stays the same; increases.
10. lower.
II. lower; lower.

Appendix 2-2 The Commercial Pilot Series


12. likely. 5. less. 12. larger; lower.
13. washout. 6. nose attitude. 13. no difference.
14. refer to page 8-13. 7. power; power. 14. design manoeuvre.
15. less; more. 8. available; required. 15. directly downwind.
16. low fluctuating airspeed; 9. excess. 16. reduce; becomes zero.
turn in the spin direction. 10. decrease; lower. 17. roll into.
17. refer to page 8-17. II. thrust. 18. increase
REV/EW9
12. tangential; top. 19. higher; roll out of.
I. lift; weight; thrust; drag.
13. lower. 20. (a) flap operating;
2. lift; thrust.
14. better cooling; (b) normal operating;
3. lift/weight; thrust/drag. better visibility; (c) caution; still;
4. down. reduced time enroute.
(d) maximum speed.
5. up. 15. time; distance.
21. will not; likely.
6. tail plane. 16. poorer.
22. lower.
7. down. 17. poorer.
REV/EW12
8. down. 18. lower.
I. thrust.
9. up. 19. (a) up and right
(b) down. 2. twisted.
10. lower. 20. decreases; decreases. 3. plane of rotation.
II. drag. 21. reduced. 4. hub.
12. drag; lift/drag ratio. 22. less than; unaffected. 5. tip.
13. drag x TAS. 23. weight. 6. (a) thrust
14. lower than. 24. lift/drag ratio. (b) torque
15. (a) 25. lower;steeper; higher. 7. propeller.
16. (b) 26. does not affect; lower. 8. increase.
17. maximum and minimum 27. does not. 9. 70%.
speeds for level flight
28. shallower; increased; 10. one.
18. (a) decreased the same.
II. fine.
(b)increased 29. nose attitude; power
applied. 12. coarse.
(c) decreased
REVIEW11 13. TAS.
19. increases.
I. centripetal. 14. zero; zero thrust.
20. up and to the right.
2. wV2/gr. 15. experimental mean.
21. down.
3. weight. 16. slip.
22. absolute.
4. lift 17. zero degrees.
REVIEW10
5. drag. 18. constant speed unit.
I. speed for altitude.
6. power. 19. propeller or pitch.
2. drag.
7. lift; weight. 20. throttle.
3. thrust; lift.
8. 1·15; 2; 2'g'. 21. coarsen.
4. (a) rearward component
plus drag 9. load factor. 22. fine off.
(b) perpendicular 23. FULL FINE.
component. 10. decreases.
II. 20° 24. FULL FINE.

Principles of Flight Appendix 2-3


25. rpm; then MP. 4. failed; yaw; slipstream. M 0·75 and M 1·2.
26. MP; then rpm. 5. left. 7. local.
27. negative; same. 6. the left engine. 8. wave (or line).
28. net. 7. toward; right; left engine. 9. normal.
29. negative; increased. 8. toward; 10. mach; oblique.
30. fine. closing the throttle on II. angled back.
31. coarse; fine. failed engine;
12. higher
32. faster; higher; thrust; engine instruments.
13. decreased
asymmetric blade effect. 9. centred; failed.
14. compression, increase
33. solidity; increasing the 10. toward.
number of blades; 15. M 1·0; unchanged.
II. (a) toward
increasing the chord of (b) toward; so 16. may be; changed.
the blades. (c) live
17. shockwave.
REVIEW13 REVIEW15
18. expansion lines or fans.
I. initial.
I. distance; fuel.
2. 19. increased; reduced.
dynamically.
2. air; unit quantity.
3. tail plane. 20. increased; reduced.
3. hour (or unit time).
4. greater. 21. before.
4. airframe; engine.
5. more. 22. decreases; shock.
5. drag/max. lift drag ratio.
6. insufficient; more difficult; 23. energy; separation.
6. does not.
may. 24. increases.
7. improve.
7. improved. 25. decreases.
8. reduced. 8. ground.
26. down; tuck.
9. sideslip. 9. increases; decreases.
27. mach trimmer; nose up.
10. (a) low rpm/high MAP
10. dihedral; shielding; (b) lean; cold 28. reduced; reduced
high wing; high fin; (c) at. 29. (b).
swept wings.
II. time; fuel. 30. a lower
II. spiral.
12. increases; low. 31. lower CLmax!higher stall
12. Dutch roll (or snaking).
angle.
13. stable; unstable. 13. reduces. high drag at high angles of
14. (a) lowest attack.
14. asymmetric blade effect;
(b) MP Prone to tip stall and pitch
slipstream effect; up
torque reaction; (c) the mixture
gyroscopic effect; 32. drag rise.
15. drag; power.
crosswind. 33. mach cone.
REVIEW16
REVIEW14 REV/EW17
I. temperature.
I. normal. I. does.
2. sound. 2. run, cleatway.
2. failed.
3. M 0·75; M 1·5. 3. 85%
3. (a) high
(b)long 4. Mach 1-0. 4. is not, may.
(c) high 5. attached 5. may.
(d) aft; low.
6. Merit and Mdet; 6. lower, lower.

Appendix 2-4 The Commercial Pilot Series


7. (a) 7.2%, (b) 9%. landing. The actual LDA 14. B. 7-1, 7-6
(b) is 670m x 3·3 = 2211 ft.
8. IS. A. 7-10
reflective, standing. 2S. 22 kt headwind; 13 kt
9. 16. D. 7-11
crosswind.
10. 2So/o, ice. 17. A. 8-S, 8-7
26. 6 kt head/tailwind; I 7 kt
II. 1013 hPa, +!Soc, 2° crosswind. 18. D 9-9
C/IOOOft, 36,090 ft.
27. "700 metres (accept 670- 19. B. 10-3
12. moisture, higher. 730).
20. D. 10-S
13. 30ft, below. 28. 2380 kg (accept 2360-
2400). 21. A. 10-11
14. press. alt. 960 ft;
QNH I 003 hPa; 29. 370 metres (accept 360- 22. c. 11-8
altitude 7700 ft; 380).
press. alt 270 ft;
23. B. 11-19
30. 21SO kg (accept 2130-
QNH 102S;
2170).
24. B. 12-4
altitude S81 0 ft.
31. 06 needs 87S metres (86S 2S. B. 13-6
IS. press. alt. I 070 ft.
- 88S); 24 needs 8SO 26. A. 7-6, 13-10
16. set I 013 on the subscale metres (840- 860).
27. A. 14-4
17. !SA+ 10; 32. 4 70 metres (4SS - 48S
ISA-6; acceptable). 28. c. IS-2
ISA+7; 29. D. 9-3
+I JOC; 33. 6!S metres (600- 630 a
cceptable). 30. B. I 0-4, I 0-9,
+4°C.
34. 74SO lb (73SO - 7SSO 11-11
18. 20SO ft press. alt, 3130 ft
dens. alt; acceptable). Part II Performance
1027 QNH, 2700 ft dens. 3S. IS,OOO ft p. alt, 14,640 ft 31. D. 17-1
alt; altitude.
6940 ft elev., 6S80 dens. 32. c. 17-3
alt; 33. B. 17-S
4SO ft press. alt, 1410 ft SAMPLE EXAMINA T/ON
dens. alt; 34. A. 17-9
reference
99S QNH, ISA-3 temp; page 3S. c. 17-13
4780 elev., S260 dens. alt; Part I Principles of Flight
I 023 QNH, ISA-4 temp. 36. B. 17-12
I. c. 1-3
19. add 1200 ft to pressure 37. D. 17-21
alt. 2. B. 1-S
38. A. 17-23
20. 2300 ft (22SO - 23SO ft 3. A. 1-6
39. A. 17-26
acceptable).
4. A. 3-3
21. 2873 ft (using 2400 ft as 40. A. 17-29
the basis).
s. C. 3-7
41. c. 17-34
22. No, the TODR must not 6. B. 3-12
42 B. 17-3S
exceed 8So/o of the TORA. 7. A. 3-13
For this take-off a TORA 43. D. 17-36
of at least (2873 + 0·8S) 8. B. 4-2
44. D. 17-36
3380 ft must be available. 9. A. 4-10
4S. C. 17-39
23. 16S2 ft (1700 X 1·08 = 10. B. S-6
1836; 1836 less I Oo/o =
16S2). II. c. 8-11

24. Yes, a minimum LDA of I 12. C. S-12, S-14


6S2 + O·SS = 1944 ft 13. C. 6-2, 6-3
must be available for this

Principles of Flight Appendix 2-5


Appendix 2-6 The Commercial Pilot Series
INDEX

climb 10-1
A forces acting I 0-2, I 0-3,
accelerateMstop distance available 17-2 performance 10-4
acceleration 1-3 maximum rate 10-5
adverse yaw 7.5 maximum angle 10-5, 10-6
aerodynamic balance (of controls) 7-11 speeds (VX, VY, normal) 10-7
aerofoil nomenclature 3-5, 3-6 factors affecting 10-8 to!0-10
aerofoils 3-5 coefficient of drag (CD) 5-13
aileron reversal 7-15 CD curve 5-13
aircraft attitude 7-I coefficient of induced drag (Cdi) 5-11
in pitch 7-3 coefficient of lift 4-2,
in roll 7-4 variation with angle of attack 4-3
in yaw 7-6 CL curve, 4-3
aircraft axes 7-1 effect of camber 4-5
airspeed, measurement of 3-2, 3-3 effect of roughness 4-5
angle of attack 3-6 effect of high CLmax 4-4
geometric 3-6, 4-8 shape of CL curve 4-4
effective 4-8 combined high-lift devices 6-7
anti-balance tabs 7-13 compressibility,
aspect ratio 4-8 effect on lift 16-15
effect on lift 4-9, 4-10 effect on drag 16-18
asymmetric flight modes, compressive comers 16-10
all rudder 14-5 constantMspeed unit 12-8
all bank 14-6 contaminated runway 17-5
combined rudder and bank 14-7 control at high speed,
asymmetric flight, longitudinal 16-19
yawing moment 14-2 lateral 16-21
rolling moment 14-2 directional 16-21
propeller torque reaction 14-2 control on the ground 13-12 to 13-14
asymmetric blade effect 14-3 couple 1-10
critical engine 14-4 critical angle 3-11
immediate actions 14-4 crosswind landing 13-17
identification of failed engine 14-5 crosswind takeMoff 13-16
atmospheric pressure 2-2
autorotation 8-15
D
density (of air) 2-1
B density altitude 2-7, 17-11
balance tabs 7-13 descending, power on 10-14
BernoulJils Theorem 3-7 (see also gliding)
boundary layer 5-2, 5-5 dihedral 13-8
laminar flow 5-3 downwash 3-11
turbulent flow 5-3 drag (definition of) 3-13
transition point 5-3 drag curve 5-14
pressure gradients 5-4 drag, classification 5-2
boundary layer blowing/suction 16-26 alternative classification 5-16
coefficient 5-13
c parasite
skin friction
5-2
5-3
Calibrated Airspeed (CAS) 3-3
form 5-5
camber 3-6
interference 5-9
centre of gravity (CG) 1-11
induced 5-9
centre of pressure (CP) 3-11
total 5-13
movement of 3-12
drift down 17-5
centripetal force 1-7, Il-l
dry runway 17-4
chord 3-5
Dutch roll 13-10
chord line 3-5
dynamic energy 3-2
clearway 17-2

Index 1
dynamic pressure 3-1,3-2
dynamic stability 13-1 H
high-speed design,
thickness chord ratio 16-22
E supercritical wing section 16-23
endurance flying, best speed 15-9 sweepback 16-23
effect of weight, altitude 15-9 disadvantages of sweepback 16-24
engine considerations 15-9 area rule 16-27
practical application 15-9 horn balance 7-12
energy 1-12 humidity (of air) 2-7
equilibrium 1-10
excess power curve 10-5
expansion waves 16-12
I
expansion fans 16-13 Indicated Airspeed (lAS) 3-3
expansive corners 16-12 inertia 1-4
inset hinges 7-11
International Standard Atmosphere
F (]SA) 2-5, 2-6, 17-8
fairings 5-9 !SA temperature deviation 17-12
fineness ratio 5-8
flap, effects on pitching moments 9-4
Flap, effects of 6-2 to 6-4 K
simple or plain 6-2, 6-4 kinetic energy 1-12
split 6-5
slotted
Fowler
6-5
6-5
L
landing distance available (LDA) 17-4
leading edge 6-7
landing distance required (LDR) 17-4
Kmeger 6-7
lateral and directional stability together 13-10
flexural aileron flutter 7-14
lateral stability 13-7
flight controls 7-2
factors affecting 13-8, 13-9
elevator 7-3
leading-edge sawtooth or notch 16-26
aileron 7-4, 7-5
lift (definition of) 3-13
rudder 7-6
lift coefficient 4-2
primary and further effects 7-6
(see also coefficient of lift)
effect of airspeed 7-7
lift formula 4-2
effect of slipstream 7-7
lift, factors affecting 4-1
canard 7-9
vee (or butterfly) tail 7-9 lifVdrag ratio 5-15
load factor 11-4
elevens 7-9
load factor limitations 11-18
tailerons 7-9
flight manual take-off and landing graphs longitudinal stability 13-3
wing pitching moment 13-3, 13-4
17-18,17-19
factors affecting 13-5, 13-6
force 1-4
looping 11-15, 11-16
freestream static pressure 3-1
Frise ailerons 7-5
M
G Mach trimmer
Mach wave (or Mach line)
16-21
16-7
gliding, forces acting 10-11
normal Mach wave 16-7
angle I 0-11
oblique Mach wave 16-8
performance 10-11 to 10-14
Mach angle 16-8
gradient of climb 17-3
Mach number 16-4,
gross flight path 17-3
freestream Mach number (Mfs) 16-5
ground effect 13-18
local Mach number (ML) 16-5
ground roll stability,
critical Mach number (Merit) 16-5
nosewheel aircraft 13-11
detachment Mach number
tailwheel aircraft 13-12
(Mdet) 16-5
groundloop 13-12
critical drag rise Mach number.
(Mcdr) 16-6
manifold pressure (MP) control 12-9
mass 1-4
mass balancing 7-15

Index 2
maximum operating
Mach number (MMO) 16-20
R
maximum rate/ ram's horn vortex 16-25
minimum radius turns 11-9toll-Il range flying,
moment 1-9 specific air range (SAR) 15-1
momentum 1-6 specific fuel consumption (SFC) 15-1
gross fuel consumption (GFC) 15-1
range: airframe considerations 15-2
N best range speed 15-2
effect of altitude, weight, wind 15-2
net flight path 17-4 specific ground range (SGR) 15-2
Newton's First Law 1-3 recommended range speed 15-3
Newton's Second Law 1-5 range: engine considerations 15-4
Newton's Third Law 1-6 factors affecting SFC 15-4
range: practical application 15-5 to 15-8
p Rectified Airspeed (RAS)
relative airflow (RAF)
3-3
3-6
Reynold's Number 5-5
p-charts runway slope and surface correction factors
older style 17-25 to 17-35 17-22
new-style 17-36 to 17-38
pitching moments in flight 9-3 s
pitot pressure 3-2 scalar quantity 1-2
pressure altitude 17-9 scale effect 5-5
potential energy 1-13 separated flow 5-6
power 1-12 separation point 5-6, 5-7
power available curve 9-10 shock stall 16-18
power required curve 9-8, 9-9 shockwaves,
propeller (pitch) control 12-9 normal shock 16-9
propeller asymmetric blade effect 12-15 oblique shock 16-9
propeller modes, bow shockwave 16-10
windmilling 12-11 wing shockwaves 16-11
feathering 12-12 single-engine seiVice ceiling graphs
reverse thrust 12-12 17-39to17-43
propeller twisting moments, SI units 1-1
centrifugal (CTM) 12-13 slats and slots 6-6
aerodynamic (ATM) 12-14 snaking 13-11
propeller, speed 1-3
blade face 12-1 speed of sound 16-1, 16-4
blade back 12-1 speed ranges,
disc 12-2 subsonic, transonic, supersonic 16-6
helix angle 12-2 spin characteristics 8-16,
pitch angle 12-2 confirmation 8-16,
angle of advance 12-2 recovel)' 8-17
blade twist 12-3 spoilers 6-8
forces acting on 12-3 stability and controllability 13-2
thrust 12-4 stagnation point 3-9
torque 12-4 stall,
rpm/airspeed relationship 12-4 symptoms of 8-2, 8-3
effective blade sections 12-5 warning devices 8-4
performance 12-6 recoveJY 8-5
experimental mean pitch 12-7 wing drop at 8-10, 8-1 I,
slip 12-7 use of aileron near and during 8-12
geometric pitch 12-8 recoveJY after wingdrop 8-13
zero-thrust angle 12-8 recovety at onset 8-13
solidity 12-16 stalling angle 3-11
propellers, constant-speed 12-8 stalling in turns I 1-5
operation of 12-10, 12-11 stalling speed,
basic 8-2
factors affecting 8-6 to 8-10
'standard' conditions 2-5, 2-6
standard rate (or rate 1) turns I 1-7
static pressure 3- I

Index 3
static stability
steep turns
13-I
II -9 w
stopway 17-2 washout 5-12
streamline flow 3-7 wave drag,
streamlining 5-7 energy drag 16-18
supersonic flow, nature 16-13, 16-14 boundary layer separation 16-18
supersonic planform shapes 16-29 weight 1-5
supersonic wing sections 16-28 wet runway 17-5
swing on takeoff, causes 13-14, 13-15 wetted area 5-5
work I· 12

T
tabulated performance data 17-43 z
take-off distance available (TODA) 17-2 zoom climb 10-1
take-off distance required (TODR) 17- I
take-off run available (TORA) 17-2
take-off safety speed 14-8,17-1
temperature (of air) 2-3
thickness/chord (tic) ratio 3-5
'thrust required' cutve 9-8
tip stalling and pitch-up 16-26
torque 1-10
torsional aileron flutter 7-15
total pressure 3-2
total reaction (TR) 3-7
transition point 5-3, 5-7
transonic drag hump 16-19
trigonometric functions 1-9
trim controls/tabs 7-9, 7-10
True Airspeed (TAS) 3-3
turbulent flow 3-7
turning, effect of wind 11-12

u
units 1-2
upwash 3·11
v
V code (speeds) 11-17
vso (stall speed landing
configuration) 11-17
VSl (stall speed clean) I].] 7
VFE (max. speed flap extended) II-17
VA (design manoeuvre speed)ll-17, 11-18
VNO (normal operating limit
speed) 11-17
VNE (never exceed speed) Il-l 7
VX (best angle of climb speed) I 0· 7
VY (best rate of climb speed) I 0-7
V-n (orV-g) diagram 11-18
vector quantity 1-2
velocity 1·3
venturi effect 3-8
Vmca (minimum control speed
asymmetric) 14-8
vortex generators 16-25
vortices 4-6
wingtip 4-7
trailing edge 4-7

Index 4

You might also like