FM 1-203 Fundamentals of Flight

Download as pdf or txt
Download as pdf or txt
You are on page 1of 448

FIELD MANUAL '''FM 1-203

NO 1-203 HEADQUARTERS
DEPARTMENT OF THE ARMY
3 October 1988
Washington, DC,
FUNDAMENTALS OF FLIGHT

CONTENTS

Page

PREFACE ix
..
CHAPTER 1. BASIC AERODYNAMICS

Newton's Laws of Motion. 1-1


I-I. . . . . . . . . . .

1-1
1-2. Fluid Flow and Airspeed Measurement. . . . . . . .

1-3. Vectors and Scalars. . . . . . . . . . . . . . . .


1-9

CHAPTER 2. AERODYNAMICS AND FLIGHT MECHANICS

Section 1. AIRFOILS

2-1. Airfoil Characteristics and Functions. 2-1


2-2. Airfoil Terminology. . . . . . . . . . . . . . . .
2-3

FORCES ACTING ON AIRCRAFT


Section II.
2-3. In-Flight Forces. . . . . . . . . . . . . . 2-5
2-4. Airfoil Airflow . . . . . . . . . . . . 2-7

Section DEVELOPMENT OF AERODYNAMIC FORCES


III.
2-5. Total Aerodynamic Force. . . . . . . . . . . . . . 2-12
2-6. For,ce Equations. . . . . . . . .
2-13

DISTRIBUTION RESTRICTION. Distribution authorized to US government agencies


and their contractors to protect information and technical data that address
current technology in areas of significant or potentially significant mili-
tary application. This determination was made on 12 February 1988. Other
requests for this document will be referred to Commander, US Army Aviation
Center. ATTN: ATZQ-DAP-SS. Fort Rucker, AL 36362-5035.

DESTRUCTION NOTICE. Destroy by any method that will prevent disclosure of


con~ents or reconstruction of the document.

*This publication supersedes FM 1-203, 9 September 1983.


i
Page

2-7. Lift and Drag Equations. . . . . . . . . . . . . . 2-14


2-8. Lift and Drag Influences. . . . . . . . . . 2-16
2-9. Angle of Attack. . . . . . . . . . . . . . . 2-19
2-10. Aerodynamic Pitching Moments. . . . . . . . 2-20

Section IV. DRAG

2-1l. Drag Considerations. . . . . . . . . . 2-22


2-12. Types of Drag. . . . . . . . . . . . . . . . 2-22
2-13. Total Drag. . . . . . . . . . . . . . . . . . . . 2-23
2-14. Lift/Drag Ratio. . . . . . . . . . . . . . . 2-24
2-15. Total Drag Curve. . . . . . . . . . . . . . . . . 2-27
2-16. Parasite Drag. . . . . . . . . . . . . . . . . . . 2-30
2-17 . Induced Drag. . . . . . . . . . . . . . . . . . . 2-34

Section V. PERFORMANCE CURVES

2-18. Curve Construction . . . . . . . . . . 2-39


2-19. Power Requirements . . . . 2-40
2-20. Power Equations. . . . . . . . . . . . . . . . . . 2-41
2-21. Power-Required Curves. . . . . . . . . . . . . . . 2-42
2-22. Power-Available Equation . . . . . . . . . . 2-43
2-23. Power-Available Curves. . . . . . . . . . . 2-44
2-24. Maximum Endurance. . . . . . . . . . . 2-44
2-25. Max imum Range. . . . . . . . . . . . . . . . 2-44

Section VI. FORCES AFFECTING PERFORMANCE

2-26. Variation Factors. . . . . . . . . . . 2-45


2-27. Effect of Weight on Performance. . . . . . . . . . 2-46
2-28. Effect of Configuration on Performance. . . . . . 2-46
2-29. Effect of Altitude on Performance 2-47
2-30. Effect of Wind on Performance. . . . . 2-48

CHAPTER 3. STABILITY

Section FIXED-WING STABILITY


I.
3-l. Motion Sign Conventions. . . . . . . . . . . 3-1
3-2. Static Stability. . . . . . . . . . . . . . 3-2
3-3. Dynamic Stability. .'. . . . . . . . . . . . . . . 3-3
3-4. Pi tch Stability. . . . . . . . . . . . . . . . . . 3-5
3-5. Directional Stability. . . . . .
3-13
3-6. Lateral Stability. . . . . . . . . . . . . . 3-17
3-7. Cross Effects and Stability. . . . . . . . . 3-20

ROTARY-WING STABILITY
Section II.
3-8. Considerations. . . . . . . . . . . .
3-22
3-9. Speed Stability. . . . . . . . . . . . . . . . . . 3-25

ii
Pa~e

3-10. Rotor Angle of Attack Stability. . . . . . . . . .


3-26
3-11. Static Stability During Hovering Flight. . .
3-28
3-12. Dynamic Stability During Hovering
Flight. . . . . 3-28
3-13. Static Stability During Forward Flight. . . . . .
3-30
3-14. Dynamic Stability During Forward Flight. . . . . .
3-34
CHAPTER 4. AIRCRAFT DESIGN, STRUCTURE, AND ENGINE OPERATION

Section DESIGN AND STRUCTURE


I.
4-1. Design Considerations. . . . . . . . .
4-1
4-2. Static Strength of Metals. . . . . . . . . .
4-1
4-3. Flight Envelope. . . . . . . . . . . . . . .
4-3
Section PROPULSION
II.
4-4. Power Sources. . . . . . . . . . . . . . . .
4-6
4-5. Propulsion Principles. . . . . . . . . . . . . . .
4-7
Section GAS TURBINE (TURBOJET) ENGINES
III.
4-6. Operating Characteristics. . . . . . . . . .
4-7
4-7. Gas Turbine-Propeller Combination. . . . . .
4-13
4-8. Gas Turbine Operating Limitations. . . . . . . . .
4-14
Section IV. RECIPROCATING ENGINES

4-9. Operating Characteristics. . . . . . . . . . . . .


4-17
4-10. Operating Limitations. . . . . .
4-25
CHAPTER 5. HELICOPTER DESIGN AND MECHANICAL CHARACTERISTICS

Section I. HELICOPTER CONTROL

5-l. Basic Single-Rotor Design. . . . . . . . . . . . .


. 5-1
5-2. Rotor System Terms. . . . . . . . . . . . . . . .
5-2
5-3. Rotor System Characteristics. . . . . . . . . . .
5-3
5-4. Rotor Blade Movements. . . . . . . . . . . .
5-6
5-5. Con ing . . . . . . . . . . . . . . . . . . . . . .
5-9
5-6. Rotor Head Control. . . . . . . . . . . . . . . .
5-11
5-7. Blade Lead and
5-8.
Lag. . . . . . . . . . . . . . . .
5-18
Torque. . . . . . . . . . . . . . . . . . . . . .
5-21
5-9. Antitorque Rotor. . . . . . . . . . . . . .
5-22
5-10. Heading Control. . . . . . . . . . . . . . . . . .
5-23
5-11. Translating Tendency . . . . . . . . .
5-24
5-12. Fuselage Hovering Attitude. . . . . . . . . . . .
5-25

Hi
Page

MECHANICAL CHARACTERISTICS AND REQUIREMENTS


Section II.
5-13. Effect of Rotor System Design on Weight and
Balance Limitations. . . . . . . . . . . . . . . . 5-28
5-14. Danger of Exceeding Center-of-Gravity Limits 5-30
5-15. Pendular Action. . . . . . . . . . . . . . . . . . 5-31
5-16. Fuselage Add-Ons, Fixes, and Modifications. . . . 5-33

CHAPTER 6. ROTARY-WING PERFORMANCE

BALANCE OF FORCES
Section I.
6-l. Newton's Second Law of Motion. . 6-1
6-2. Balanced and Unbalanced Forces. . . . . . . . . .
6-2

DEVELOPMENT OF AIRFLOW ON THE ROTOR SYSTEM


Section II.
6-3. Relative Wind. . . . . . . . . . . . . . . .
6-4
6-4. Rotational Relative Wind . . . . . . . . . . 6-6
Induced 6-7
6-5. Flow. . . . . . . . . . . . . . . .

6-10
6-6. Resultant Relative Wind. . . . . . . . . . . . . .

6-7. Airflow in Forward Flight. . . . . . . 6-10

DIFFERENTIAL AIRFLOW IN THE ROTOR SYSTEM


Section III.
6-8. Causes of Dissymmetry of Lift in the Single-Rotor
Helicopter. . . . . . . . . . . . . . . . . . . . 6-14
6-9. Compensating for Dissymmetry of Lift in the
Single-Rotor Helicopter. . . . . . . . . . . 6-15
6-10. Tandem-Rotor Helicopter Dissymmetry of Lift. 6-20
6-20
6-1l.
6-12.
Tail Rotor Dissymmetry of
Transverse Flow Effect.
Lift. . . . .
.

.
.

.
.

. . . . . . . 6-21

Section IV. HOVERING

6-13. Forces at Hover.


a . . . . . . . 6-23
6-14. Airflow During Hovering. . . . . 6-23
6-15. Ground Effect. . . . . . . . . . . . . . . . . . . 6-24

Section V. AIRFLOW PATTERNS IN FORWARD FLIGHT

6-16. Translational Flight. . . . . . 6-28


6-17 .
Effective Translational Lift . . . . . 6-29

Section VI. COMPRESSIBILITY EFFECTS

6-18. Compressible and Incompressible Flow. . . . . . . 6-30


6-19. Transonic Flow Patterns. . . . . . . . 6-34

iv
Page

Section VII. EMERGENCY SITUATIONS

6-20. Retreating Blade Stall. . . . .


6-39
6-21. Settling With Power. . . . . . . . . . . . . . . .
6-43
6-22. Resonance. . . . . . . . . . . . . . . . . . . . .
6-48
6-23. Dangerous Combinations of Height and Velocity. . .
6-50

Section AUTOROTATION
VIII.
6-24. Aerodynamics of Vertical Autorotation . . . . . . .
6-53
6-25. Aerodynamics of Autorotation in Forward Flight 6-55
6-26. Glide and Rate of Descent in Autorotation . . . . .
6-61
6-27. The Last 100 Feet. . . . . . . . . . . . . . . . . 6-67

Section IX. PERFORMANCE CURVES

6-28. Performance Factors. . . . . . . . . . . . . . . .


6-71
6-29. Power Requi rements . . . . . . . . . . . . . . . .
6-71
6-30. Power Available. . . . . .
6-73
6-31. Forward Flight Performance . . . . . .
6-73

Section X. MANEUVERING FLIGHT

6-32. Acceleration. . . . . . . . . . . . . . . .
6-79
6-33. Turning Performance. . . . . . . . . . . . .
6-82
6-34. Turning Flight. . . . . . . . . . . . . . .
6-82
6-35. Radius of
Turn. . . . . . . . . . . . . . .
6-83
6-36. Vertical Turns . . . . . . . . . . . . . . .
6-86
6-37. Rate of Turn. . . . . . . . . . . . . . . .
6-87
CHAPTER 7. ROTARY-WING FLIGHT TECHNIQUES

Section I. PLANNING CONSIDERATIONS

7-l. Training. . . . . . . . . . . . . . . . . . . . .
7-1
7-2. Attitude Flying. . . . . . . . . . . . . . . . . .
7-2
7-3. Attitude Control and Coordinated Turns. . . . . .
7-3
7-4. Attitude Control and Airspeed. . . . . . . .
7-5
7-5. Heading Control and Antitorque Pedals. 7-6
7-6. Power Control and Resulting Altitude, Climb, or
Descent. . . . . . . . . . . . . . . . . . . . . .
7-13
7-7. Traffic Pattern. . . . . . . . . . . . . . . . . .
7-14

FIELD OPERATIONS
Section II.
7-8. Basic Considerations. . . . . . . . . . . .
7-18
7-9. Reconnaissance. . . . . . . . . . . . . . .
7-20
7-10. Confined-Area Operations . . . . . . . . . . 7-21
7-11. Pinnacle and Ridgeline Operations. . . . . . . . .
7-22
7-12. Terrain-Flight Operations. . . . . . . . . . . . . 7-23

v
Page

7-13. Slope Operations. . . . . . . . . . . . . . 7-24


7-14. General Precautions. . . . . . . . . . . . . . . . 7-28

CHAPTER 8. FIXED-WING PERFORMANCE

Section AIRCRAFT DESIGN


I.
8-l. Classification of Aircraft. . . . . . . . . . . . 8-1
8-2. Aircraft Structure. . . . . . . . . . . . . . . . 8-2

Section HIGH-LIFT DEVICES


II.
8-3. Purpose. . . . . . . . . . . . . . . . 8-4
8-4
8-4.
8-5.
Increasing the Coefficient of
Types of High-Lift Devices.
Lift. . . . .
.

.
.

.
.

.
.

.
. . .

8-7

Section STALLS
III.
8-6. Characteristics. . . . . . . . . . . . . . . 8-11
8-12
8-7. Aerodynamic Stall. . . . . . . . . . . . . .

8-14
8-8. Stall Warning. . . . . . . . . . . . . . . . . . .

8-9. Stall Recovery. . . . . . . . . . . . . . . . . . 8-17


8-10. Stall-Speed Equation . . . . . . . . . . . . 8-17

Section IV. MANEUVERING FLIGHT

8-11. Climbing Performance. . . . . . . . . . . . . . . 8-22


8-12. Climbing Flight. . . . . . . . . . . . . . . 8-22
8-13. Steady-State Climb. . . . . . . . . . . . . . . . 8-23
8-14. Angle of Climb. . . . . . . . . 8-24
8-15. Rate of Climb. . . . . . . . . . . . . . . . . . . 8-26
8-16. Climbing Stall Speed. . . . . . . . . . . . . . . 8-26
8-17. Performance of an Aircraft in a Climb or Dive 8-26

Section V. GLIDES

8-18. Performance. . . . . . . . . . . . . . . . . . . . 8-29


8-29
8-19. Power-On Glide. . . . . . . . . . . . . . . . . .

8-30
8-20. Power-Off Glide. . . . . . . . . . . .

8-21. Ma~imum-Glide Distance. . . . . . . . . . . . . . 8-31


8-22. Wind Effect on Glides. . . . . . . . . . . . . . . 8-32

Section VI. TURNS

8-23. Performance. . . . . . . . . . . . . . . . . 8-32


8-24. Turning Flight. . . . . . . . . . . . . . . . . . 8-33
8-25. Radius of Turn. . . . . . . . . . . . . . . 8-34
8-26. Vertical Turn. . . . . . . . . . . . . 8-37
8-27. Rate of 8-38
Turn. . . . . . . . . . . . . . . .

vi
PaRe

Section VII. TAKEOFF AND LANDING PERFORMANCE

8-28. Procedures and Techniques . . . .


8-38
8-29. T akeo f f . . . . . . . . . . . . . . . . . . .
8-39
8-30. Land ing . . . . . . . . . . . . . . . . . . . . . .
8-41
Section VIII. FLIGHT CONTROL SYSTEMS

8-31. Development. . . . . . . . . . . . . . . . . . . .
8-44
8-32. Control Surface Operation Theory. . . . . .
8-44
8-33. Longitudinal Control. . . . . . . . . . . .
8-47
8-34. Directional Control. . . . . . . . . . . . .
8-49
8-35. Lateral Control. . . . . . . . . . . . . . .
8-51
8-36. Control Forces. . . . . . . . . . . .
8-51
8-37. Control Systems. . . . . . . . .
8-55
8-38. Propellers. . . . . . . . . . . . . . . . . . . .
8-57
CHAPTER 9. FIXED-WING FLIGHT TECHNIQUES

9-1. Taxi ing . . . . . . . . . . . . . . . . . . .


9-1
9-2. Trim Tab Use. . . . . . . . . . . . .
9-4
9-3. Takeoffs. . . . . . . . . . . . . . . . . .
9-4
9-4. Air Work . . . . . .
9-7
9-5. Banks. . . . . . . . . . . . . . . . . . . .
9-13
9-6. Attitude Flying . . . . . . . . . . . . . . .
9-14
9-7. Slow Flight. . . . . . . . . . . . . . . . . . . .
9-15
9-8. Stalls. . . . . . . . . . . . . . . .
9-17
9-9. Sp
ins. . . . . . . . . . . . . .
9-19
9-10. Traffic Patterns. . . . . . . . . . .
9-21
9-11. Approaches. . . . . . . . . . . . . .
9-22
9-12. Slips. . . . . . . . . . . . . . . . .
9-23
9-13 .
Lan,dings . . . .

9-23
.........
. . . . .

9-14. Go-Arounds . . . . .
9-26
CHAPTER 10. FIXED-WING MULTIENGINE OPERATIONS .........
10-1. Light-Twin Aircraft Performance . . . .
1D-1
10-2. Asymmetric Thrust . . . . . . . . . . . . . . . . . 1D-2
10-3. Critical Engine . . . . . . . . . . . . . . . . . . 1D-3
10-4. Minimum Single-Engine Control Speed . . . . . 1D-4
10-5. Single-Engine Climbs . . . . . . . . . 1D-7
10-6. Single-Engine Level Flight . . . . . .
1D-7
10-7. Single-Engine Descents . . . . . . . . . . .
lD-7
10-8. Single-Engine Approach and Landing . . . . . 1D-8
10-9. Propeller Feathering . . . . . . . . . . . .
lD-8
10-10. Accelerate-Stop Distance . . . . . . . . . .
lD-9
10-11. Engine-Out Procedures . . . . . . . . . . . . . . . 1() -10
10-12. Pressurized Aircraft . . . . . . . . . 1() -14

vii
Page

GLOSSARY

Section I. Terms and Definitions. . . . . . . . . .


Glossary-l
Section II. Abbreviations and Brevity Codes . . . . .
Glossary-23
Section III. Symbols. . . . . . . . . . . . . . . . .
Glossary-28
Section IV. V-Speed Abbreviations. . . . . . . . . . . . .
Glossary-29
Section V. Aerodynamic Equations. . . . . . . . . . . . . Glossary-31

REFERENCES . . . . . . . . . . . . . . . . . . . .
References-l

INDEX. . . . . . . . . . . . . . . . . . . . . . . . . . . .
Index-l

viii
PREFACE

the
prepared specifically for Army aviators. It presents
and rotary-wing
This manual is and the principles
of fixed-wing
fundamentals of aerodynamicsa guide and reference for--
flight. also serves as
It and
aviation students during the primary
0
Fixed-wing and rotary-wing
advanced stages of training.
presenting instructional material.
0
Academic instructors in
the students' fundamental knowledge of
reinforcing
0
Instructor pilots in
flight. knowledge of
the students' fundamental
evaluating
0
Flight evaluators in
aerodynamics and flight principles.
or aircraft
when undergoing instructor pilot training
0
Rated aviators
qualification training.
flight opera-
the aerodynamics that relate directly to of the
This manual presents practical experience, and applicationand general
the theory, transition training,
tions. It covers to flight training, mathematical
aerodynamics that apply specialized
For simplicity and clarity, effective flight
flight operations. considered unnecessary to
detail was omitted where it was
not need a working
knowledge of the aero-
probably mission
operations. Aviators will this manual for the performance of normal
dynamic formulas given
in
flight tasks.
HQ TRADOC. Submit changes for improving
publication is and
The proponent of thisDA 2028 (Recommended Changes to Publications
the
this publication on
Form commander to
and forward it through the aviation unit ATTN: ATZQ-
Rucker,
Blank Forms), Aviation Center and Fort
Army
Connnander, United States 36362-5163.
TDP, Fort Rucker, Alabama

of the following
of this publication are the subject Standard Visual
The provisions STANAG 3554 (Edition Two),
international agreement: R/T Phraseology.
and Heliport Traffic
Patterns and Associated
Airfield

operations security
been reviewed for
This publication has
considerations.

gender is used, both men


whenever the masculine
Unless otherwise stated,
and women are included.

ix
CHAPTER 1

BASIC AERODYNAMICS

other gaseous fluids and the


air and
Aerodynamics concerns the motion of
motion through the air (gases). In effect,
forces acting on objects in object (aircraft), the movement (relative
aerodynamics is concerned with the laws of
and the air (atmosphere). This chapter reviews certain basic
wind), and scalars
acting on an aircraft. Vectors
motion, fluid flow, and forces
the explanation of aerodynamics.
are also discussed to simplify

1-1. NEWTON'S LAWS OF MOTION

acceleration, and action-reaction.


Newton's three laws of motion are inertia,
These laws are applicable to the
flight of all aircraft. By knowing these
understand the aerodynamics presented
laws of motion, aviators can better
later.
body at rest will
first law, inertia, states inthat
a

Inertia. The
a.
body in motion will remain motion at the same speed
remain at rest and
a
Nothing
affected by some external force.
and in the same direction until about or prevent motion.
stops without an outside force to bring
starts or of motion is
The resistance that
a body offers to a change in its state
called inertia.
The second law, acceleration, asserts
that the force
b. Acceleration. body is directly proportional
chaRge in the motion of
a
a
required to produce refers to
and the rate of change in its velocity. Accelerationcommonly used
to its mass
deceleration is
an increase or
a decrease in velocity, although
is a change in the magnitude or
to indicate a decrease. Acceleration respect to time. Velocity refers to
with
direction of the velocity vector motion of an object.
the direction and the rate of linear
that for
Action-Reaction. The third law, action-reaction, states
c. and opposite reaction. If an interaction
every action there is an equal directions will be
forces in opposite
occurs between two bodies, equal
imparted to each body.
MEASUREMENT
1-2. FLUID FLOW AND AIRSPEED
a Swiss mathematician,
Bernoulli's Principle. Daniel Bernoulli, between internal fluid
a. the relationship
stated a principle that describes essentially a statement of the
His principle,
pressure and fluid velocity. somewhat why an airfoil develops an
conservation of energy, explains
aerodynamic force.

1-1
(1) All forces acting on a
surface over which there is a flow of
are the result of skin friction or air
of viscosity and are confined to a pressure. Friction forces are the result
very thin layer of air near the
surface.
(2) As an aid in
visualizing what happens to pressure as
over an airfoil, it is helpful to air flows
consider flow through a tube. The
of conservation of mass states that concept
mass cannot be created or
goes in one end of a tube must come destroyed; what
out the other end. In Figure 1-1, the
flow through a tube is neither
accelerating
the mass of flow per unit of time nor decelerating at the input;
at station 1 must equal the mass of
per unit of time at station 2 and then flow
at
unit area (cross-sectional area of tube) station 3.
The mass of flow
per
is called the mass flow rate.

I
I I I
STATION STATION STATION
0 Ð Ð

Figure 1-1. Airflow through a tube


(3) At low speeds,
air experiences relatively small changes in
pressure and negligible changes in
density. This airflow is termed incom-
pressible since the air may undergo changes
in pressure without apparent
changes in density. Such airflow is similar
to the flow of water, hydraulic
fluid, or any other incompressible fluid. As the
tube is reduced, the velocity of the cross-sectional area of the
fluid must increase to maintain a
constant mass flow rate; this suggests that
the velocity varies inversely
with the cross-sectional area. Venturi
this phenomenon. Figure 1-2 effect is the name used to describe
illustrates
flow through the tube when the
what happens to the speed of the
area size changes.
(4) The total energy in a
given closed system does not change, but
the form of the energy may be
altered. The pressure of flowing air may be
likened to energy in that the
total pressure of flowing air will always
remain constant unless energy is added
to or taken from the flow. In the
examples in Figures 1-1 and
1-2, energy is not added or subtracted;
therefore, total preSsure remains constant.

1-2
,
,
, STATION
STATION
STATION
e Ð p = 2,074Ib/ft2

2,116 Ib/ft2
0 p = 2,014lb/ft2
q = 761b/ft2
q= 1361b/ft2
=
P
H =
2,150 Ib/ft2
q = 34 Ib/ft2 H = 2,150 Ib/ft2
H = 2,150 Ib/ft2

Figure 1-2. Venturi effect

made up of twO
components--static
(5)Fluid flow pressure is pressure is defined
ss the force per
pressure. Static component of total
pressure and dynamic a surf see. Dynamic pressure 1s that
on
unit area acting the motion of the air.
pressure due to
barometer placed
pressure is measured by an aneroid multiplying the
(a) Static is computed by
but not moving with the flow. It column of air.
(Table 1-1
in the flow heigbt of a

weight of air by the


a

average specific Atmospheric static pressure on


shows the ICAO
standard atmosphere chart.) as 14.7 pounds per square inch,
standard day at sea
level can be expressed of mercury, or 1,013 millibars.
foot, 29.92 inches
2,116 pounds per square following equation:
expressed by the
Static pressure is (Equation 1.1)
p = wh

static pressure pounds per cubic foot)


weight of air (in
where P =

average specific
of air (in feet)
w =

h = height of a column

1-3
Table 1-1. lCAD standard
atmosphere

Pressure Temp Density


All Ratio Ratio Ratio
It ð (f 1/2 film' Pressure Temp Density Speed 01 Sound
8 (f IbIIP Ib/II' of P slugs/It' , IVsec mph 1
All
0 knots It
1.000000 1.0000 1.00000 1.0000 1482.524
1000 0.964387 0.9931 2116.216 59.00
0.97106 0.9854 1429.727
0.00237690 1116.9 761.5
2000 0.929809 2040.852 55.43 661.7 0
0.9862 0.94277 0.00230812 1113.0
3000 0.9710 1378.464 1967.676 758.9 659.5
0.896241 0.9794 51.87 0.00224087 1000
0.91512 0.9566 1109.2
4000 0.863661 1328.698 1896.639 756.3 657.2 2000
0.9725 0.88808 48.30 0.00217514
0.9424 1280.398 1105.3 753.6
1827.693 44.74 654.9 3000
5000 0.832047 0.00211088 1101.4 751.0
0.9656 0.86167 0.9283 652.6 4000
6000 0.801377 1233.530 1760.791
0.9587 0.83586 41.17 0.00204810 1097.5
7000 0.9143 1188.060 1695.886 748.3 640.3
0.771628 0.9519 37.60 5000
0.81064 0.9004 0.00198675 1093.6 745.6
8000 0.742780 1143.957 1632.932 647.9 6000
0.9450 0.78601 34.04 0.00192681
0.8866 1101.190 1089.7 743.0
9000 0.714812 0.9381 1571.884 30.47 645.6 7000
0.76196 0.8729 0.00186827 1085.7 740.3
1059.726 1512.697 643.3 8000
10000 26.90 q.00181110
0.687703 0.~312 1081.8 737.6
0.73848 0.8593 640.9 9000
11 000 0.661432 1019.536 1455.328
0.9244 0.71555 23.34 0.00175528
12000 0.8459 980.589 1077.8 734.9
0.635980 0.9175 1399.733 19.77 0.00170078
638.6 10000
0.69317 0.8326 1073.8 732.2
13000 0.611327 942.855 1345.871 636.2 11 000
0.9106 0.67133 16.21 0.00164759
14000 0.8193 906.307 1069.8 729.4
0.587453 1293.699 12.64 633.8 12000
0.9037 0.65002 0.8062 0.00159568 1065.8
870.913 1243.178 726.7 631.5
9.07 13000
15000 0.564340 0.00154503 1061.8 723.9
0.8969 0.62923 629.1 14000
16000 0.7932 836.648
0.541969 0.8900 1194.266 5.51 0.00149562
0.60896 0.7804 1057.7 721.2
17000 0.520322 803.482 1146.924 1.94 626.7 15000
0.8831 0.58919 0.00144743 1053.7
18000 0.7676 771.389 718.4 624.3 16000
0.499380 0.8762 1101.113 -1.62
0.56991 0.7549 0.00140043 1049.6 715.6
19000 0.479126 740.342 1056.795 621.9 17000
0.8694 0.55112 -5.19 0.00135461 1045.5
0.7424 710.315 712.8 619.4 18000
1013.933 -8.76 0.00130995 1041.4 710.0 617.0 19000

(b) Dynamic
pressure is computed by multiplying
air density by the square
of the one-half the
the following velocity of the airflow.
equation: It is expressed by
q =
1/2p V2
(Equation 1.2)
where q =
dynamic pressure
P (rho) =

air density (in slugs


V = foot) per cubic
velocity of the airflow (in
feet per second)

1-4
Dynamic pressure is difficult to measure
directly, but a pitot-static tube
measures it indirectly. The sum of the dynamic and static pressures is
total
pressure. Total pr~ssure is measured by allowing the flow to impact against
an open-end tube which is vented to an aneroid barometer. The incompressible
or slow-speed form of the Bernoulli equ~tion is expr~ssed as follows:

H =
p +
1/2p V2; or H =
P + q (Equation 1.3)
H
where =
total pressure (in Ib/ft2)
p =
static pressure (See Equation 1.IJ
p(rho) =
air density (in slugs/ft3)
V =
velocity (in ft/sec)
q =
dynamic pressure (See Equation l.~)
(6) Static pressure decr~ases as velocity increases. On the bottom
half of the venturi tube shown in Figure 1-3, the shape of the c011stricted
area at station 2 resembles the top surf~ce of an airfoil. Even when the top
half of the venturi tube is taken away, the air still accelerates over the
curved shape of the bottom half. This happens because the air layers
restrict the flow just as the top half of the venturi tube does. As a
result, acceleration causes decreased static pressure above the curved shape
of the tube. This is what happens to air passing over the curved top of an
aircraft wing. The local variation of static and dynamic pressures on the
curved s~rface generates a pressure differential force. This force is
explained in Chapter 2. .;

UPPER LAYERS ACT TO RESTRICT FLOW

I
.,

.~ I. .
. þ
.

~~ ~~
~
--~ fffffffff f f f f f f
;: f f
III"~:
1/FIlllln.:UII;;
STATION
-: STATION
I
I
STATION
0 e Ð

Figure 1-3. Venturi flow

1-5
(7) simple illustration of water flowing through a garden hose
A

demonstrates how changes in the diameter of the hose affect flow speed.
Water moving through a hose of a constant diameter exerts a uniform pressure
on the hose. However, if the diameter of a section of the hose is increased
or decreased, it is certain to change the pressure of the water at that
point. If the hose is pinched, the area through which the water flows is
thereby constricted. Assuming that the same volume of water flows through
the constricted portion of the hose in the same period of time as before the
hose was pinched, it follows that the flow speed must increase at that point.
If a portion of the hose is constricted, the flow speed not only increases a
but static pressure decreases at that point. Like results are achieved if
streamlined solid (airfoil) is introduced at that same point in the hose.
This principle is the basis for measuring airspeed (fluid flow) and for
analyzing the airfoil's ability to produce lift.
.

b. Airspeed Indicator. The Bernoulli equation is the basis of the


construction of the airspeed indicator. It is also the principle of the
pitot-static system for measuring airspeed.
(1) In Figure 1-4, a symmetrically shaped object has been placed in
the airstream; this results in the flow pattern shown by the dashed lines.
The airstream ahead of station 1 has a certain total pressure due to the
velocity, density, and static pressure of the airstream. At the point where
the airstream strikes the forward end of station 2, the relative velocity of
the airstream is reduced to zero. This is referred to as the forward stag-
nation point. Since the static pressure increases as the velocity decreases,
the static pressure at the forward stagnation point must increase until it is
equal to the total pressure of the airstream. When there is no velocity,
there can be no dynamic pressure; therefore, the static pressure at the stag-
nation point is equal to the total pressure. Dynamic pressure is determined
by subtracting the static pressure in the airstream at station 1 from the
total pressure. The pitot-static pressure system is constructed to determine
dynamic pressure.

1-6
POINT
FORWARD STAGNATION

1
,
j. ----~--_.-.-
/--~---+--
-----..... --~-
_l.---~
- --
\-- ... ---+-----"
-
-
---+-
/~ ~

---~--~--
~--=- =- -..... ~
- + ---+-
-
-

~ ~ ~ .::.:::::-~
~ , ::::.~~
~-.-
-

t-- ..-=--=
- -

--..
-
---+-
\ ""!
------~-~--
\----~_.-.-
-

.1
\
STATION

0
\
STATION

e
---~----
Figure 1-4. Flow pattern in airstream

schematic drawing of
a pitot-static system.
1-5 shows a
tube (total pressure
port)
(2) Figure the pitot-static open
end of the
At station 2, the
open
The velocity of the
airflow impacting
Since
projects into the airflow. to zero, forming a stagnation point. of the
end of the tube
decreases
is equal to the total pressure
the static pressure present in the air-
velocity is zero, The static pressure that was that altitude.
airstream at this point.
1 is equal to the
atmospheric pressure at The static
stream at station perpendicular to the airflow.
This pressure must
be measured be any"dynamic pressure
designed so tbere will not measured
pressure port must
be
increase in the static pressure
cause a slight passes the static
influences tbat could The velocity of the airstream as it
at station 1.
at the static port. be equal to the velocity of the airstreamstation 1, then
pressure port must than at
higher velocity at the pressure port atmospheric pressure of the
If there is pressure will be lower than the true
a
The port is
the static of the static port is very important. of the fuselage.
airstream. The location of the pitot tube or on the side free-stream
the side must be the
either located on the velocity passing the port be accomplished.
Wherever it is located, this cannot always
airspeed. However, of the static pres-
velocity or the true introduced because of the location with an instrument
Therefore, any error a along
called position error. This error, and the calibrated
sure port is the indicated airspeed most
difference between charts found in
error, is the can be corrected by using corrective
airspeed. It
operator's manuals.

1-7
TOTAL PRESSURE PORT

INSIDE OF
DIAPHRAGM

STATIC PRESSURE PORT

DENSITY SPRING

Figure 1-5. Measuring pressure


at stagnation point
(3) The airspeed
pi tot tube and the indicator receives the
total
Since it is a
free-stream static pressure from the pressure from the
differential pressure-measuring static pressure port.
ference of the two device, it indicates the dif-
pressures (dynamic
in either knots or miles pressure) on a dial that is
standard sea-level
per hour. When an
airspeed indicator is
calibrated
density conditions are calibrated,
neglected, the indicated used.
airspeed is equal to the
If the position error is
for incompressible true airspeed at sea level
flow.
(4) At altitude, the
level and the IAS is lower air density is less than the
than the actual density at sea
is true because the IAS aircraft velocity or TAS. This
is actually the dynamic
velocity. As altitude pressure and the TAS is the
airspeed decreases increases, density decreases.
if
ship between IAS and TAS
true airspeed is held Therefore, indicated
constant. There is a
that
altitude to the density of the involves the ratio of the density of relation-
the air at
the difference between air at sea level. The phenomenon
IAS and TAS which causes
constant IAS will be is the density. An
increasing its actual aircraft climbing at a
decreasing air density. velocity, or TAS, because of

1-8
VECTORS AND SCALARS
1-3. tWO typea of
enhanced hy understanding
flight is further described by
A study of aircraft Scalar quantities are those must be
and vectors. Vector qusntities
quantities--scalars volume, time, and mass.
size alone such as area,and direction.
described by magnitude
and drag
Velocity, acceleration,
weight, lift, qnantities
a. Vector Quantities. The direction of vector
of vector quantities. from whatever
are commoo examples as the size or magnitude. All forces, by two or more
is just as important being acted upon
When an object is represented by the ose witb
of
source, are vectors. effect of these forces may bea segment
combined directed line
forces, the graphicallY represented by which the force.is
vectors. Vectors are
indicateS the direction in the mag-
an arroW at tbe
end. The arrow to a given scale represents
segment length in relation a reference line.
acting. Line is drawn in relation to
The vector a specific
nitude of the force. whatever scale is most convenient to
Magnitude is drawn to
problem.
useful in analyzing
Individual force vectors are the resultants
b. Vector Solutions. the the chief concern is with an airfoil or
In air,
conditions of flight. the several component forces acting
on
of sre the parallelo-
or combined effects of solving for
resoltants
Three methods
on an aircraft.
and triangle methods.
gram, polygon, two vectors,
A parallelogram contains
(1) Parallelogram method.
plane. The other vector
known reference The
parallel to
a

with one vector drawn from the tail of the first vector.
appropriate angle
is drawn at the completed by drawiog opposite sides
parsllel to the known
poiot of
psrsllelogram is by drawing
a vector from the two
is determined Figure 1-6,
vectors. The resultant corner of the parallelogram. In
berge moves fotwsrd
origin to the opposite s barge with equal force; the The
pnshing of both tugboats.
tngboats are shown the direction
that is a mean to
in a direction method most often used in
aerodynamics.
is the vector solntion
parallelogram

1-9
-------
~
<I:
0
III
CJ
~
~

0 TUGBOAT

il~
Figure 1-6. Resultant by parallelogram
method
(2) Polygon method. When
more than two
different directions, the forces are acting in
resultant
solution. One solution is shown may be found by
using a polyzon
in Figure 1-7. A force vector
angle of 090 degrees with a is acting
180 at an
grees with 90 pounds; and a force of pounds; a second
force, at 045
determine the
third force, at 315 degrees de-
resultant, the with 120 pounds. To
and the remaining first vector is drawn from the
vectors are drawn point of origin
mined by drawing a consecutively. The resultant
line from the point of is deter-
vector. origin to the end of the
final

1-10
c

~
~
~
~':J B
.c..~~
-\)'V ;::::~
{v'? q,o
q,.; ~q
'),

0900
0 A
180 POUNDS

method
Figure 1-7. Resultant by polygon
simplified and
A triangle of vectors is
a

method. and their


(3) Triangle solution involving only two vectors used in
vector
special form of polygon method is the vector solution most often
The triangle vectors and con-
resultant. of vectors is formed by drawing two can
navigation.
A triangle
line of vector. In this way, calculations
heading
necting them with
a resultant an aircraft is
ground speed. In Figure 1-8, from the north-
100 knots and the wind is
and
be made for drift of
true airspeed and a con-
078 degrees with each known velocity
a

By drawing
a vector for
east at 30 knotS. determined.
between the ends,
a resultant velocity is
necting line
used to depict the various
Throughout this manual,
vectors will be The forces shown in
NOTE:
and on aircraft in flight.
forces acting on airfoils chapters.
discussed in the following
Figures 1-9 and 1-10 are

1-11
N

þ.ÑO
e.Þ.D\ÑG
if\\.I!:e.Þ.\f\S~~OiS
e.D
f,,1? ~
in07801'\00 ~ ~O~,O
~{:. f<"CJ
r<,.Q
RESULTANT
o\'!) ~ ~f(;
~~ S)Q ~ Gj
0900/77 KNOTS
TRUE COURSE AND ~~,f
GROUND SPEED

Figure 1-8. Resultant by


triangle method

TOTAL AERODYNAMIC
FORCE
----. -

--
lÑ~
f\O'" -
RESULTANT
C~~
-

DRAG --- ..-INDUCED


FLOW
ROTATIONAL RELATIVE WIND
(FREE STEAM AIR)

AIRFOIL

Figure 1-9. Force vectors on


airfoil

1-12
RESULTANT
LIFT LIFT

LIFT

1
-4
DRAG ~ THRUST
~

1
WEIGHT
--
-

RESULTANT OF WEIGHT
DRAG AND WEIGHT

Force vectors on aircraft in flight


Figure 1-10.

1-13
CHAPTER 2

MECIl1\ti11CS
AEROOYNMUCS AND FLIG3T

to use the power


flight depends on their ability forces.
in thrust, and drag
rtoW well aviators perform the lift, weight,
changing
and flight controls for the balance betIJeen
these forces.
control
Aviators must always

SECTION I

AIRFOILS

AND ~UNCTIONS
2-1. AIRmIL CllARACTEíUSTICS
usefully upon
or body designed to react
An airfoil is structure,
a piece,
usually has
a crosS section care-
through the air. It but it may be
itself in its motionsuit its intended applicetion or function, missiles, or
fully contoured to to aircraft,
flat plate. Airfoils are applied (wing or rotor blade).
to develop lift
a
no more than
or projectiles and thrust or
otber aerial vehicles stability (fin), control (elevator),
such as rotor
They are also used for blade). Certain airfoils,
rotor
propulsion (propeller or
some of these functions.
blades, combine that
A helicopter flies for
the same basic reason
Designs. necessary to keep it
a. Airfoil the aerodynamic forces blades have
conventional airplane flies;
The
a
when air passes
about the rotor blades. As
are produced of flight characteristics.
aloft specific set this
designed for through the air;
a

airfoil sections when passing


surfaced bodies, airfoils produce lift are less efficient
under
Some airfoil designs of
makes flight possible. airspeeds. Other combinations
higher they may
s pedf ic condi dons,
yet permit may generate more lift,
but
lower-surface designs must compromise
upper-surface and travel. Usually the designer the
have ~ide center-of-pressure
a
the best flight
characteristics for
section ~ith
to obtain an airfoil perform.
~ission the aircraft ~ill
basic types--
Airfoil sections are of t~o 2-1) have
b. Airfoil Sections. (Figure
Symmetrical airfoils (cambered)
symmetrical and nonsymmetrical. designs; nonsymmetrical
and lower-surface
identical upper- in design.
(Figure 2-2) may vary ~idely
airfoils more suited to
symmetrical airfoils are
(1) Symmetrical airfoils. no center-of-pressure
because they have almost constant under a
rotary-wing applications travel ramains relatively
produces less lift
center-of-pressure than
travel. The A symmetrical airfoil
has relatively
varying angles of attack. a given angle of attack. It also
nonsymmetrical airfoil
at

2-1
undesirable stall
to a wide range of characteristics. The helicopter blade (airfoil) must
airspeeds and angles of adapt
the rotor. Even with attack during each
those alternating revolution of
can deliver acceptable conditions, the symmetrical
foil, when
performance. Other benefits airfoil
of the symmetrical
compared to the
cost. nonsymmetrical, are ease of construction air-
and lower

4lI1t1/-7~-
MEAN CAMBER LINE
CHORD LINE

Figure 2-1. Symmetrical airfoil section


(2) Nonsymmetrical (cambered)
airfoils, nonsYmmetrical airfoils have airfoils. Compared to
symmetrlcal
desirable characteristics. increased
used on fixed-wing Until recently, these lift-drag ratios and more
aircraft. They were not used onairfoils were primarily
aircraft because the center-Of-pressure earlier rotary-wing
angle of attack changed. location moved too much when
When the center
of pressure moves, a the
is exerted on the
would withstand airfoil. Rotor system components had to be
twisting force
the twisting designed that
materials used to manufacture force. Recent design processes and
rotor systems have the new
problems. Nonsymmetrical partially overcome these
airfoils are currently used on most
copters; for example, the OH-58,
AH-1S (Modernized), Army heli-
They are UH-60, AH-64 and CH-47.
increasingly being used on newly ,
designed aircraft.

c1#~-
MEAN CAMBER LINE
CHORD LINE

Figure 2-2. Nonsymmetrical (cambered)


airfoil section

2-2
2-2. AIRFOIL TERMINOLOGY

The terminology associated with airfoils must be understood


before
aerodynamic forces can be understood. Figure 2-3 illustrates a typical
airfoil section and the various airfoil terms.

MAXIMUM THICKNESS
MAXIMUM CAMBER

LEADING EDGE RADIUS

AERODYNAMIC
CENTER CHORD LINE

Figure 2-3. Typical airfoil section

a. Certain terms describe the shape of an airfoil. Some of these terms


are defined below.

(1) Chord line. A straight line intersecting the leading and


trailing edges of the airfoil.

(2) Chord. The length of the chord line from leading edge to
trailing edge; it is the characteristic longitudinal dimension of the airfoil
section.

(3)Mean camber line. A


line drawn halfway between the upper and
lower surfaces. The chord line connects the ends of the mean camber
Camber refers to the curvature of the airfoil and may be considered
line.
the
curvature of the mean camber line.
(a) The shape of the mean camber is important for determining
the aerodynamic characteristics ot an airfoil section. Maximum camber
(displacement of the mean camber line from the chord line) and its location
help to define the shape of the mean camber line. The location of maximum

2-3
camber and from the chord line are expressed as fractions or
its displacement
of maximum
percentages of the basic chord length. By varying the point
camber, the manufacturer can tailor an airfoil for
a
specific purpose. The
airfoil shown in Figure 2-3 is a positive-cambered airfoil because the
mean
camber line is above the chord line.

(b) The profile thickness and thickness distribution are


important properties of an airfoil section. The point of maximum thickness
and its location help to define the airfoil shape; the quantities are
expressed as a percentage of the chord length.

(4)
Leading-edge radius. The radius of curvature given th~
leading-edge shape.

b. Several terms are used to describe the development of aerodynamic


properties. Some of these terms are defined below.
(1) Flight-path velocity. direction of the airfoil
The speed and
FPV is equal to true
passing through the air. For fixed-wing airfoils,
airspeed. For helicopter rotor blades,
FPV is equal to rotational velocity
plus or minus a component of directional airspeed.

Relative wind. The air in motion that is equal to and opposite


(2)
FPV may be modified by an
the flight-path velocity of the airfoil. Because
induced flow of air, the RW experienced by the
airfoil may not be exactly
opposite its direction of travel. In this manual, this modified relative
winà is referred to as resultant relative wind.

(3) Angle of attack. The angle measured between the resultant


relative wind and the chord.

(4) Angle of incidence (fixed-wing aircraft). The angle between


the airfoil chord line and the longitudinal axis or other selected reference
plane of the airplane.

Angle of inci.dence (rotary-wing aircraft). The angle between


(5)
the chord line of a main or tail rotor blade and the plane of rotation (tip-
path plane). It is usually referred to as the blade pitch angle. For fixed
is the
airfoils, such as vertical fins or elevators, thea angle of incidence plane
angle between the chord line of the airfoil and selected reference of
the helicopter.

(6) Center of pressure. The point along the chord line of an


airfoil through which all aerodynamic forces are considered to act.

(7) Aerodynamic center. The point along the chord line where all
changes to lift effectively take place. If the center of pressure is located
behind the aerodynamic center, the airfoil experiences a nosedown pitching
moment. A change in the center of pressure changes the magnitude of the
located at the 25 percent
pitching moment. The aerodynamic center is usually
chord.

2-4
SECT ION
II
FORCES ACTING ON AIRCRAFT

2-3. IN-FLIGHT FORCES

a. Lift. The term lift describes two distinct forces acting on an


aircraft. One lift force acts on an airfoil, and the other lift force acts
on the entire aircraft.
(1) Lift acting on an airfoil. This force is the component of the
total aerodynamic force that acts at right angles to the resultant relative
wind. Resultant relative wind is the direction of the airflow with respect
to the airfoil. This lift
force is considered to act through the center of
pressure of the airfoil. The magnitude of this lift
varies proportionally
with the square of the velocity (V2) of the airflow across the airfoil, air
density, shape and size of the airfoil, and angle of attack. Figure 2-4
shows lift acting on an airfoil section.

AIRFOIL DIRECTION AND SPEED

Figure 2-4. Lift force acting on an airfoil section


(2) Lift acting on the entire aircraft. This force, often referred
to as t~e vertical component of lift, is the sum of all lift forces produced
by the lift-producing surfaces of the aircraft. These surfaces include the
wings, rotor blades, elevators, and horizontal stabilizers. This lift force
acts through the center of lift, which is the mean of all the centers of
pressure of all lift-producing surfaces. In straight and level flight, it is
equal to and opposite weight (Figure 2-5).

2-5
LIFT

THRUST
~ ~ .

~
ÆpARASrrE DRAG

t
WEIGHT

Figure 2-5. Forces acting on entire aircraft in flight


b~ Weight. This force is exerted by an aircraft from the pull of
gravity. Weight acts on an aircraft through the center of gravity; its
direction is straight down toward the center of the earth (Figure 2-5). The
magnitude of this force chanp,es only with a change in gross weight.

c. Thrust. This force drives an aircraft forward through the air


(Figure 2-5). Thrust is produced by a rotating propeller, helicopter rotor
system, jet engine, or some other propulsive device.

d. Drag. This force is always parallel to the relative wind. Drag is


produced by the resistance of the air to an object passing through it (Figure
2-5). In unaccelerated flight, drag is equal to and opposite thrust. Addi-
tional drag forces are produced from pressure distribution on the airfoils
and from friction caused by the movement of air across the airfoils. Sec-
tion IV contains a more detailed explanation of drag forces.

e. Centrifugal Force. Centrifugal force is the component of apparent


force acting on a body moving in a curved path that is directed away from the
center of curvature or axis of rotation. This force results when a body
attempts to maintain linear (uniform) motion according to Newton's first law
of motion and the centripetal force changes the linear velocity. In the
study of physics, centrifugal force is not recognized as a true force; in
this manual, it is considered equal in magnitude to centripetal force but
opposite in direction.

f. Centripetal Force. Centripetal force is the accelerative force


acting on a body moving in a curved path. It is the component of force that
is directed toward the center of curvature or axis of rotation. Centripetal
force changes the direction of the linear-velocity vector of a body in motion
and results in an acceleration of the body. Since centrifugal force is only
an apparent force, centripetal force is the out-of-balance force that causes
an aircraft to turn. 'fhen an aircraft turns, centripetal force is the
horizontal component of lift that is directed toward the center of the turn.

2-6
Centripetal force is created by tilting the total lift vector toward the
center of the turn, as shown in Figure 2-6.

CENTRIFUGAL FORCE
CENTRIPETAL
FORCE

VERTICAL VERTICAL
TOTAL TOTAL
LIFT LIFT
LIFT LIFT

WEIGHT WEIGHT

CENTRIFUGAL
CENTRIFUGAL
FORCE
FORCE
CENTRIFUGAL FORCE

Figure 2-6. Centrifugal and centripetal forces in a


normal turn

2-4. AIRFOIL AIRFLOW

The production of lift


can be analyzed with two different theories--the
airfoil section, or blade element, theory and the momentum theory. These
theories are mutually supportive ways to explain airfoil lift. As an airfoil
moves through the air, it disrupts the natural flow of air; this changes the
direction and velocity of the airflow around the airfoil. These changes,
shown in Figure 2-7, are described below where they are keyed to the figure.

2-7
STRAIGHT-LINE DISTANCE 7

-8

v+
G) p-

Figure 2-7. Airflow around the airfoil

a. Airfoil Section, or Blade Element, Theory. The airfoil section, or


blade element, theory is based on Bernoulli's principle of airflow. The
following discussion of airfoil lift analyzes how the aerodynamic conditions
that exist on a cross section of a wing or rotor blade affect airfoil lift.
It also explains how variations occur on the airfoil from root to tip as the
aerodynamic environment changes. The total lift can then be determined by
adding together the effects for all wings or blades.

(1) Asthe relative wind CI) strikes the leading edge of the
airfoil, its velocity is reduced to zero. This point of impact is called the
stagnation point~. The pressure at this point on the airfoil is equal to
total pressure (dynamic pressure plus static pressure). This area of high
pressure creates pressure waves which are propagated ahead of the airfoil at
the speed of sound. These pressure waves cause the wind moving toward the
airfoil to deflect over and under the airfoil beginning some distance ahead
of it CD. This air flows smoothly around the airfoil following its sha.pe ø
and defl~s at a slight angle downward off the trailing~d
e of the
The air following the shape of the airfoil 6 has a greater
airfoil~.
distance to travel than if it had not changed direction 7 ;. therefore, the
air must speed up or a partial vacuum would be created after the airfoil
passes. Nature does not allow the existence of a vacuum in the earth's
atmosphere. To prevent this vacuum from forming, the airflow over and under
the airfoil increases in velocity. As the velocity of the airflow increases,
the dynamic pressure above and below the airfoil also increases.

2-8
(2) According to the law of conservation of energy, an increase in
one component of energy must be accompanied by a corresponding decrease in
another component of energy. Total energy (total pressure) cannot change.
To accommodate the increase in dynamic pressure around the airfoil, static
pressure decreases above and below the airfoil. The air has a greater
distance to travel over the upper surface of the airfoil than over the bottom
surface. This causes a g~ter velocity (dynamic pressure) increase and
static pressure decrease ~ over the top surface than over the bottom
surface of the airfoil~. This static pressure differential across the
airfoil produces the total aerodynamic force ~
created by the airfoil.
The pressure differential between the upper and lower surfaces of the airfoil
is about 1 percent. Even a small pressure differential produces a
substantial force when applied to a
large area.

(3) Because of the viscosity of the air, some of it sticks to the


surface of the airfoil. This thin layer of air near the airfoil surface is
called the boundary layer QJþ. Because of the stickiness, the layer of air
molecules against the skin of the airfoil has no movement with respect to the
airfoil. Each succeeding layer of air molecules increases in velocity as the
distance away from the airfoil increases. The boundary layer is considered
to end where the velocity of the air molecules reaches 99 percent of the
velocity of the airflow around the airfoil. This distance is about the.
thickness of a playing card. This boundary layer forms the transition
between the air and the airfoil. Without it, the airfoil would produce very
little lift.
b. Momentum Theory. Another method of analyzing the production of lift
is the momentum theory. According to Newton's second law, the production of
lift requires A
that a mass of air be accelerated to a final downward direction
and velocity. brief discussion of this theory is in Chapter 6.

c. Pressure Differential. Bernoulli's principle states that as


velocity increases, dynamic pressure also increases; this causes static
pressure to decrease. Figure 2-8 shows how this principle works as flow is
constricted in a stream tube. Thus static pressure on the upper surface of
an airfoil decreases when velocity increases. The static pressure on the
upper surface is less than the static pressure on the lower surface. Even if
the static pressure on both surfaces is less than the atmospheric pressure,
the airfoil will still produce lift. The important point here is the
pressure differential that develops across the airfoil.

2-9
~~~~ ~~~~
FREE FLOW
CONSTRICTION
(As dynamic
pressure (velocity)
increases, static
pressure decreases.)

Figure 2-8. Constricted flow in a stream tube

(1) Figure 2-9 shows the pressure differential by using vectors.


Those that point away from the surface of the airfoil represent pressures
below atmospheric) and those that point toward the airfoil represent
pressures above atmospheric.

PRESSURE PATTERN IN
POSITIVE-CAMBERED
AIRFOIL PRODUCING LIFT

Figure 2-9. Pressure differential across an airfoil


(2) An example of how to compute pressure differentials is given in
Figure 2-10. The equation for free-stream pressure (total pressure) is
H =
P + 1/2p V2 or H p +
=
q (Equation 1.3).

2-10
.

~900 ft/sec
--ní71771/I7lÞ .. 800 ft/sec
850 ft/sec
~
Air density =
.002377 slug/ft3 (sea level, standard day, from
Table 1-1)

Air pressure =
2,116 lb/ft2 (static pressure, sea level,
standard day, from Table 1-1)

Free stream pressure H =


P + 1/2p V2
(Total pressure)
H =
2,116 + .002377/2 x 8002
H + .001189 640,000
=
2,116 x

H =
2,116 + 761
H =
2,877 1 b/ft2
Total pressure of free stream =
total pressure on both upper and lower
surfaces of airfoil

Upper surface pressure =


2,877 =
P + .002377/2 x 9002
(Solve for p.) 2,877 =
p + .001189
810,000 x

2,877 =
p + 963.09
p =
2,877 963.09
-

p =
1,913.91 lb/ft2
Upper surface static pressure =
1,913.91 lb/ft2

Lower surface pressure =


2,877 =
p + .002377/2 x 8502
(Solve for p.) 2,877 =
p .001189 x 722,500
+

2,877 =
p + 859.05
P =
2,877 859.05
-

p 2,017.95 lb/ft2
=

Lower surface pressure =


2,017.95 lb/ft2

Difference between upper and lower surface pressure =


2,017.95 -

1913.91
=
104.04 lb/ft2 =
Lift

Figure 2-10. Computation of pressure differential on an airfoil

2-11
SECTION
III
DEVELOPMENT OF AERODYNAMIC FORCES

2-5. TOTAL AERODYNAMIC FORCE

As air flows around an airfoil, a differential develops between the


pressure
upper and lower surfaces. differential, combined with the resistance of
The
the air to the passage of the airfoil, creates a force on the airfoil. This
force, known as total aerodynamic force, is represented by vector in a

Figure 2-11; TAF acts at the center of pressure on the ai doil and is
normally inclined up and to the rear.
TOTAL AERODYNAMIC FORCE

"n"E 'l'I"'~O
þ.t.~1" ~E\.Þ.
~ES\.l\.1"
... INDUCED FLOW

DIRECTION AND VELOCITY


OF AIRFOil

Figure 2-11. Development of total aerodynamic force

a. Components. The forces and moments acting on an airfoil can more


easily be described and predicted by resolving the TAF into two components.
The force perpendicular to the resultant relative wind is called the lift;
force parallel to the resultant relative wind is called drag (Figure 2-12).

INDUCED FLOW
DIRECTION
AND VELOCITY
OF AIRFOIL
DRAG

Figure 2-12. Development of airfoil lift and drag

2-12
b. Factors. The aerodynamic forces produced by an airfoil depend on
many different factors, to include--
0
Air density.
0
Air viscosity.
0
Airfoil surface area.
0
Shape of the airfoil.
0
Airfoil angle of attack.
a
Airflow veloci ty across the ai rfoi 1.
To predict the forces acting on an airfoil under certain flight conditions,
these factors can be related to equations.

2-6. FORCE EQUATIONS

The aerodynamic force acting on an airfoil can be represented as the product


of three quantities. Together these quantities contain the factors given
above.

a. The general form of the force equation is expressed as follows:

F =
CF 1/2p S v2; or F =
CF q S (Equation 2.1)

where F =
the aerodynamic force
CF =
the force coefficient, which might be considered the
efficiency with which the airfoil develops the force in
question
p (rho) =

air density (in slugs/ft3)


S the surface area of the airfoil (span times chord in square
feet)
V =
velocity (in ft/sec)
q =
the dynamic pressure of the stream of air flowing around
the airfoil (See Equation 1.2.)

b. Equation 2.2 can be rearranged to solve for the coefficient of


force. The CF equation is expressed as follows:
F
CF =

1/2p S V2; or CF =~ q S (Equation 2.2)

Equation 2.2 shows that the CF is a dimensionless parameter which relates the
aerodynamic force to the product of the dynamic pressure and the surface
area. When applied to such forces as lift and drag, this equation is useful
in evaluating lift and drag under varying conditions of flight.

2-13
(1) The force coefficient of the equation varies only with the
shape of the airfoil and angle of attack if other quantities are held
constant. From this single dimensionless number, the relative efficiency of
the airfoil in generating lift
and drag may be evaluated. These coefficients
are easily determined from wind-tunnel experiments as well as from actual
flight. An example of the effect of CF on an aerodynamic force is
experienced in sticking a hand out the window of a moving car at some angle
of attack. A force acting upward and aft is felt; this is the aerodynamic
force. If the hand is rotated, creating a larger angle of attack, the
aerodynamic force increases even though the velocity of the car and the size
of the hand area remain unchanged. The angle of attack is one of the four
variables contained in the CF' When an airfoil is tested to find its
aerodynamic characteristics, the CF is derived.

(2) These tests can be conducted in a wind tunnel by mounting the


airfoil on probe called a stinger (Figure 2-13). The stinger can vary the
a

angle of attack of the airfoil and measure the direction and magnitude of the
aerodynamic force. The density and velocity of the airstream through the
wind tunnel are known; the area of the airfoil can be measured. Therefore,
CF can be determined for each angle of attack.

TOTAL AERODYNAMIC FORCE

æ~ME~ATIO~ STINGER

... AERODYNAMIC CENTER

SCALE /40Oo~
-

~T-
't:i

Figure 2-13. Stinger used to determine coefficient of force

2-7. LIFT AND DRAG EQUATIONS

Lift and drag are influenced by the shape and area of the airfoil, angle of
attack, air density, and airspeed. Figure 2-14 shows some of these factors
acting on an airfoil. A change in any of the factors affects the relation-
ship of liftand drag. This is apparent in solving the lift
and drag
equations.

2-14
a. The lift equation is expressed as follows:
L CL 1/2p S V2; L CL q S (Equation 2.3)
=
or =

where L =

lift force
CL =
coefficient of lift
p(rho)= density of the air (in slugs per cubic foot)
S
total wing area (in square feet)
=

V
airspeed (in feet per second)
=

q =
dynamic pressure (See Equation 1.2.)

LIFT' - -
vAERODYNAMIC FORCE

I
ANGLE OF ATTACK I

~
I
I
I
I

IF
OF TRAVEL
AIRF.OIL DIRECTION

Figure 2-14. Forces acting on an airfoil


b. The drag equation is expressed as follows:

D =
CD l/2p S V2; or D =
CD q S (Equation 2.4)

where D =
drag force
CD =
coefficient of drag
P(rho) =
density of the air (in slugs per cubic foot)
S =
total wing area (in square feet)
V =
airspeed (in feet per second)
q =
dynamic pressure (See Equation 1.2.)

c. The lift
and drag forces can then be determined directly from the
aerodynamic force (Figure 2-13). Equations 2.3 and 2.4 can be rearranged to

2-15
solve for the coeff~cients of lift and drag. The CL and Co equations are
expressed as follows:
L
CL =

1/2p S VL
; or CL
=~ q S
(Equation 2.5)

0
Co =

1/2p S VL
; or CD =~ q S
(Equation 2.6)
The dynamic pressure (q) and the area of the airfoil (8) are known. After
lift and drag are measured, the CL and CD can be calculated for each angle of
attack. Curves similar to those shown in Figure 2-15 will result. This same
procedure can be applied in a wind-tunnel experiment to determine the CL and
CD characteristics of a scaled model.

.14 1.4
-
/
V ./
-.....
.12 1.2

J
.10 1.0
~ V
.08 0.8
CD / v~ V CL
.06 V 0.6
/ /
.04 0.4

.02
1/ V ...V 0.2

0 "L -
0

0 2 4 6 8 10 12 14 16 18
ANGLE OF ATTACK

Figure 2-15. Coefficients of lift and drag plotted


against angle of attack

2-8. LIFT AND DRAG INFLUENCES

a. The two factors that most affect the CL and CD are the shape of the
ai~foil and the angle of attack. The shape is established by the manufac-
turer, and the angle of attack is controlled by the aviator. As the length
of the upper camber increases, lift also increases to a certain point. If
the length increases further, the airflow separates from the airfoil, causing
a
loss of lift. This is the angle of attack at which the airfoil begins to
stall. As lift increases, drag increases. Therefore, the airfoil should
produce the most lift
(highest CL) and least drag (lowest CD) at the weight
and speed for which it is designed. Figure 2-16 shows how and drag lift
increase with the angle of attack. The dashed line shows how the
lift-to-drag ratio (CL/CD) varies with different angles of attack.

2-16
.2000 2.0 20

.1800 1.8 18

.1600 1.6 16

.1400 1.4 14

.1200 1.2 12
CD
CL LID
.1000 1.0 10

.0800 0.8 8

.0600 0.6 6

.0400 0.4 4

.0200 0.2 2

0 0
00 20 40 60 80 100 1~ 140 160 180 200 220

ANGLE OF ATTACK

Figure 2-16. Lift-drag relationship

b. Density is the mass of an object per unit volume; it differs from


pressure in that pressure is force per unit area. Temperature, pressure, and
humidity directly affect air density. Since density affects lift, then
temperature, pressure, and humidity also affect lift. An increase in
temperature or humidity or a decrease in pressure causes density to decrease;
lift.then decreases. To sustain level flight when these conditions exist,
the aviator must increase either the angle of attack or the airspeed.

c. Wing area directly affects lift and drag. If two wings have the
same proportions and airfoil sections, a wing with an area of 200 square feet
lifts twice as much at the same angle of attack and airspeed as a wing with
an area of 100 square feet.

d. the square of the velocity. Therefore, an airfoil


Lift varies with
traveling at 200 knots has four times as much lift as one traveling at 100

2-17

~_.w.._- --...- -"


----
knots when other factors remain constant. If airspeed changes, then some
other factor must change inversely to maintain the same lift. The only other
controllable factor is the angle of attack. For a given airspeed, weight,
and air density, only one angle of attack will maintain level flight.

e. Figure 2-17 shows the CL curves for both a symmetrical a~d a


cambered airfoil. The curve for the cambered airfoil shifts up and to the
left of the curve for the symmetrical airfoil. The CL is zero when the
symmetrical airfoil is at a zero-degree angle of attack. The relative wind
aligns with the chord, and the curvatures of the upper and lower surfaces are
identical; this results in no pressure difference and, therefore, no lift.
The upper surface of the cambered airfoil has more curvature than the lower
surface when the chord line aligns with the relative wind; therefore, the
airfoil develops lift at a zero-degree angle of attack. For a given airfoil,
the value of CL varies only with a change in the angle of attack. Regardless
of the type of airfoil section, the CL curves are similar in appearance. The
curve has a linear portion over which the CL increases directly with the
angle of attack until reaching a point near CLmaximuQ; at that point, the
curve begins to round off. The high point of the CL curve is called
CLmaximum' At angles of attack above CLmaximum' the curve starts to descend,
indicating the airfoil has started to stall. (Aerodynamic stall is covered
in Chapters 6 and 8.) All CL curves have a similar appearance. However, the
values of CLmaximum vary from 0.8 to 1.7, and the angle of attack at which
CLmaximum occurs is usually from 10 to 16 degrees. The slope of the linear
portion of the curve is an approximate 0.11 increase in CL per degree of
increase in the angle of attack; this varies slightly with the type of
airfoil. Thus the exact curve depends on the physical characteristics of the
airfoil section, which is defined by its shape.

1.8

1.6

1.4

1.2

1.0
...I
U
.8

.6

4 6 8 10 12 14 16 18 20
ANGLE OF ATTACK
ZERO LIFT POINT

Figure 2-17. Lift curves for cambered and symmetrical airfoils

2-18
f. In the past, families of airfoils were developed by varY1ng the ratio
of the--
0
Maximum camber to chord length.
0
Maximum thickness to chord length.
0
Location of maX1mum camber to chord length.
0
Location of maximum thickness to chord length.

Recently, new families of airfoils have been developed to minimize the drag
and compressibility effects and to improve the CLmaximum performance. They
combine variations in the thickness distribution and the shape of the mean
camber line to achieve a more favorable pressure distribution across the
chord.

2-9. ANGLE OF ATT ACK

The angle of attack (Figure 2-18) is the aerodynamic angle between the
airfoil chord and its direction of motion relative to the air (resultant
relative win~). It is one of the primary factors that determine the amount
of and drag produced by an airfoil.
lift

RESULTANT RELATIVE WIND


ANGLE OF
ATTACK

CENTER OF PRESSURE &O~-~ INDUCED


"'"

ROTATIONAL RELATIVE WIND OR


...-
- -
FLOW
- -

FREE-STREAM VELOCITY
Þ DIRECTION OF AIRFOIL

'"

Figure 2-18. Angle of attack

a. When aircraft slow down, aviators must make some correction for
the
As used in
decrease in velocity to maintain enough lift to stay at altitude.
the lift equation (L CL 1/2p S V2), CL is the ratio of the lift pressure
=

and a function of the shape of the airfoil and angle of attack. Velocity and
CL are the only variables affecting lift that aviators can control. They
control CL by varying the angle of attack. To maintain constant lift as an
the angle of attack to increase CL;
aircraft slows down, they must increase
this compensates for the decrease in velocity.

2-19
b. In rotary-wing aircraft, several factors may cause .the angle of
attack of rotor blades to change. Aviators can control some factors while
others occur automatically because of the design of the rotor system. By
moving the cyclic and collective pitch controls, aviators can adjust the
angle of attack. Even when they hold these controls stationary, the angle of
attack constantly changes as the blade moves about the circumference of the
rotor disk. Some factors over which aviators have little control are blade
flapping, blade flexing, and gusty wind or turbulent air conditions.

2-10. AERODY[~AflIC PITCHING MOMENTS

a. Pressure distribution over a surface is the source of the aerodynamic


pitching moments as well as the aerodynamic forces. Figure 2-19 shows
pressure distribution acting on the cambered airfoil. The pressures distrib-
uted on the upper surface produce the upper surface lift; those on the lower
surface produce the lower surface lift. The net lift produced by the airfoil
is then the difference between the lifts on the upper and lower surfaces.
The point along the chord where the distributed lift is effectively concen-
trated is the center of pressure. The CP is essentially the center of
gravity of the distributed lift pressure. The camber, angle of attack, and
section lift coefficients determine the location of the CPo

b. Another aerodynamic reference point is the aerodynamic center. The


AC is considered the point along the chord where all changes in lift
effectively take place.
(1) For the symmetrical airfoil, shown on Figure 2-19, this point
can be visualized by noticing how pressure distribution changes with a change
in the angle of attack. At zero lift, the upper and lower surface lifts are
equal and located at the same point. When the angle of attack increases, the
upper surface lift increases while the lower surface lift decreases. The
change in lift takes place with no change in the center of pressure; this is
character~stic at symmetrical airfoils.
(2) For the cambered airfoil to produce zero
lift, the upper and
lower surface lifts must be equal. These lifts are not opposite one another
on the cambered airfoil as they are on the symmetrical. While no net lift
exists on the airfoil, the couple produced by the upper and lower surface
lifts creates a nosedown moment. As the angle of attack increases, the upper
surtace lift increases while the lower surface lift decreases. A change in
lift takes place, but no change in moment takes place about the point where
toe lift change occurs. The moment about the AC is the product of a force
(lift at the CP) and a lever arm (distance from CP to AC). An increase in
lift moves the center of pressure toward the aerodynamic center.
(3) The symmetrical airfoil at zero lift has no pitching moment
aoout the aerodynamic center. This is because upper and lower surface lifts
act along the same vertical line. An increase in
lift on this airfoil
produces no change in this situation; the center of pressure remains fixed at
the aerodynamic center.

2-20
CAMBERED AIRFOIL
DEVELOPING POSITIVE
UPPER
LIFT
SURFACE
LIFT
NET
LIFT

LOWER SURFACE LIFT


~~
SYMMETRICAL AIRFOIL
AT ZERO LIFT CAMBERED AIRFOIL
AT ZERO LIFT

UPPER SURFACE UPPER SURFACE


LIFT LIFT

LOWER SURFACE
LIFT LOWER SURFACE
LIFT

SYMMETRICAL AIRFOIL CAMBERED AIRFOIL


AT POSITIVE LIFT AT POSITIVE LIFT

I
UPPER SURFACE UPPER SURFACE
LIFT LIFT

LOWER SURFACE LOWER SURFACE


LIFT LIFT

t CHANGE IN LIFT

~
~ CHANGE IN LIFT

+ PITCHING MOMENT
AC
AC

Figure 2-19. Deve lopment of aer,)dynamic pitch ing moments

2-21
(4) The camber,thickness, and angle of attack have little effect on
the location of the aerodynamic center of an airfoil. Regaraless of these
factors, the two-dimensional incompressible airfoil theory predicts the
aerodynamic center at the 25 lJercent chord lJoint tor any airtoil. Actual
airfoils, which are subject to real fluid flow, may not have the lift caused
by the anLle of attack concentrated at the exact 25 percent c.hord IJoint.
However, the actual location of the AC for various sections is rarely forward
of the 23 percent or aft of the 27 percent chord lJoint.

SECTION IV

DRAG

2-11. DRAG C()t~:;ID:ERATIONS

The force that retards the motion of an aircraft or airfoil through the air
is referred to as drag. At tiwes, this force decreases the performance of
the aircraft and is a hindrance. At other times, drag can be an advantage.
Because ùrag retards motion and increases fuel consumption, it also affects
To
performance objectives such as range, endurance, and maximum velocity.
reduce drag, aviators will fly with clean aircraft. Drag requirements vary
according to flight conditions; therefore, aviators must understand this
force to obtain the required performance from their aircraft.

2-12. TYPES OF DRAG

(Dp) and induced


Two types of drag are produced at subsonic speeds--parasite
is
(Di)' The drag acting on nonlift-producing surfaces of the helicopter systems
called parasite drag; the parasite drag that acts on helicopter rotor
is usually referred to as profile drag (Do)' Drag that is due to the ~roduc-
tion of lift is called induced drag. At subsonic velocities, total drag (Dt)
is the sum of parasite and induced urag. Both types of drag are briefly
discussed below but will be explained in detail in paragraphs 2-16 and 2-17.

(Dp) results tram the way the air flows


a. Parasite Drag. Parasite arag
around the aircraft. TIle shape, surface, smoothness, size, and design of the
aircraft affect this type of drag. Also contributing to l'arasite drag are
add-ons such as wingstores, exposed weapon systems, and drop tanks. If any
of these cause turbulence or hinder airflow about the aircraft, they increase
the parasite drag. An example of how parasite drag is affected can be seen
A !.louern car offers much
by comparing recent models of cars to oleer ones.
less resistance to the airflow than an older car that has the same cross-
sectional area but has square corners and vertical surfaces. At the same
speeds, the modern car has much less parasite drag than the older car,
(Do) (helicopter
pr imarily because of its streamlined shape. Prof ile drag
only) is incurred from the frictional resistance (form and skin friction) of
the rotor blades passing through the While profile drag rioes not change
air.
significantly with the. angle of attack of the airfoil section, it does

2-22
increase moderately as airspeed ~ncreases. Profile drag is the parasitic
drag of the rotor system.

b. Induced Drag. Induced drag (Di) results from the downward velocities
imparted to the air by the wing as it produces lift and to the vortexes
developed by the' wing or blade tip. If no lift is produced, there is no
induced drag.

2-13. TOTAL DRAG

Total drag (Dt) is that component of the total aerodynamic force parallel to
the relative wind that tends to retard the motion of the aircraft or airfoil
(Figure 2-20). It acts parallel to the flight path or the relative wind but
is not necessarily parallel to the thrust vector.

a. Total Drag Equation. Since drag is a component of the aerodynamic


force, the drag equation is the same as the aerodynamic force equation
(F CF 1/2p S V2) except for the
=

coefficient. The drag equation is


D =
CD 1/2P S V2. The CD equals the CFsinX (Figure 2-20) and, as with CL, CD
is the part of the above equation that is determined from wind-tunnel tests.
The coefficient of drag equation is CD D/qS. =

Df

" TOTAL
AERODYNAMIC
FORCE

Figure 2-20. Drag as a component of aerodynamic force

2-23
b. Coefficient of Drag. Wind-tunnel results can be illustrated by
plotting the CD against the angle of attack, as shoNO in Figure 2-21. The CD
continues to increase even after the started to decr~ase. Since CL curve has
there is always some resistance to motion, drag will never be zero. There-
fore, the CD will never be zero when the airfoil is moving. The two curves
are plotted with different scales along the vertical axis. As with all
airfoil characteristic curves, the CD varies only wit~ the angle of attack
for a given aircraft at a subsonic velocity and with a constant configu-
ration.

.14 .
1.4
-
/
./
........
.12 / 1.2

.10 v' /
1.0
r V
.08 0.8
Co / c..~ V CL
.06 / 0.6
/ /
.04 0.4
/
.02
/ -
"""""
.,V 0.2
0 0

0 2 4 6 8 10 12 14 16 18
ANGLE OF ATTACK

Figure 2-21. Coefficient of drag plotted against angle of attack

2-14. LIFT/DRAG RlITIO

An airfoil is built to obtain always produce some drag. lift, but it will
Drag is the price paid to obtain of lift to drag (L/D) lift. A
ratio
indicates airfoil efficiency. Aircraft with higher L/D ratios are more
efficient than those with lower L/D ratios.

a. Lift/Drag Equation. The L/D


ratio is determined by dividing the CL
by the CD' which dividing the lift equation by the drag
is the same as
equation. All terms except coefficients cancel out. The lift/drag ratio is
expressed as follows:

L CL L CL 1/2p S V2

-;
=
or =
- -

D CD D CD S V
2 (Equation 2.7)
1/ 2p

A
ratio of the coefficients at a certain angle of attack determines the L/D
ratio at that angle of attack. The L/D can be plotted against the angle of
attack; this is the third airfoil characteristic curve. The peak of the

2-24
curves in Figures 2-22 and 2-23 is (L/D)maximumi it occurs at the most effi-
cient angle of attack for the airfoil. The angle of attack for (L/D)maximum
changes with the configuration shown in Figure 2-22.

12

10

8
0
j:
C(
III:
~
6
C(
III:
Q
......
~
::;
4

-80 -40 00 40 80 12D 160 200 240

ANGLE OF ATTACK

Figure 2-22. Typical lift-to-drag ratios (fixed-wing aircraft)

2-25
.2000 20

.1800 18

.1600 1.6 16

.1400 14

.1200 1.2 12
Co CL LID
.1000 1.0 10

.0800 0.8 8

.0600 0.6 6

.0400 0.4 4

.0200 0.2 2

0 0
00 20 40 60 80 100 120 140 160 180 200 220

ANGLE OF ATTACK

60 CL =
.5 LID =
fL
CD
~
60 CD =
.04 LID =

.04
LID =
12.5

Figure 2-23. Lift-drag relationship

2.,..26
b. Maximum Performance. Performance should not be confused with
efficiency. supersonic fighter has a higher performance level than a
A

glider, but less efficiency by comparison. Some maximum performance


maneuvers are flown at the (L/O)maximum angle of attack; these maneuvers
include--
0
Maximum endurance for a
jet aircraft.
0
Maximum range for a
propeller aircraft.
0 a
Maximum angle of climb for jet aircraft.
0
Maximum power-off glide range for both jet and propeller aircraft.
A
jet trainer supersonic jet fighter have their best, or maximum, angle
and a

oi climb at the (L/O)maximum angle of attack (disregarding afterburner).


Because of its greater thrust, however, the jet fighter greatly outperforms
the trainer; its climb is much steeper, even though its (L/D)maximum ratio is
about one-third that of the trainer. Table 2-1 shows the L/Omaximum values
of several aircraft.

Table 2-1. Aircraft (L/D)maximum values


AIRCRAFT (L/D)MAXIMUM

High-Performance Sailplane 25:1-40:1

Typical Patrol or Transport 12:1-20:1

High-Performance Bomber 20:1-25:1

Propeller-Powered Trainer 10:1-15:1

Jet Trainer 9:1-16:1

Transonic Jet Fighter 10:1-13:1

Supersonic Jet Fighter 4:1-9:1

Helicopter 3:1-5:1

1-15. TOTAL DRAG CURVE

a. At subsonic speeds, the total drag curve contains only induced and
parasite drag (including profile drag for helicopters). In Figures 2-24 and
2-25, .the independent curves for the induced and parasite drag are plotted
against the velocity. (The construction of these curves is discussed in
Section V.) As the velocity increases, the induced drag decreases and the .

parasite drag increases.

2-27
b. For an aircraft in level flight to maintain a constant altitude while
iacreasing its airspeed from minimum to maximum flight velocity, the lift
produced by the wings must remain constant. As a car accelerates along a
level road, it does not change weight; the road has to support the total
weight of the car, no matter how fast it goes. This is also true for
aircraft; lift must remain constant or the aircraft will change altitude.
Therefore, the numerator in the L/D ratio is constant, but the denominator
varies. In Figure 2-24, at the velocity where the total drag is minimum
(240 knots at the bottom of the curve), the denominator is minimuQ and the
L/D ratio is at its maximum. The most efficient angle of attack for the
airfoil is the one where the L/D is maximum. At (L/D)maximum' the induced
drag is equal to the parasite drag. At velocities below the lninimum-drag
velocity, induced drag is predominant; at velocities above the minimum-drag
velocity, parasite drag is predominant.

c. The angle of attack is plotted below the velocity scale only as a


reference; this scale can shift either to the left or to the right when the
weight or altitude of the aircraft changes. To understand how weight affects
the angle of attack with reference to the velocity, the (L/D)inaximum point
must be considered. If a pilot takes on 1,000 pounds of fuel and wants to
re~ain at the most efficient angle of attack, the lift force must increase by
1,000 pounds to support the added weight. To stay at the (L/D)maximum angle
of attack, the CL in the lift equation must remain constant. Therefore,
velocity must increase to obtain the lift required by the added weight.
Minimum drag will still occur at the same angle of attack, but at a higher
velocity. This shifts the angle-of-attack scale to the right.

ë 5.0
8-
0-
M
~

-4.0
CJ
C
a:
Q

3.0
\ BUFFET
LIMIT
2.0

1.0

0
100 200 .300 400 500 600
VELOCITY (TAS)

I'
ANGLE OF ATTACK I I
180 50 20 10

Figure 2-:-24. Typical drag plot (fixed-wing aircraft)

2-28
TOTAL
DRAG

1
PROFILE
DRAG

CJ
c
a:
0
B

INDUCED
DRAG

FORWARD SPEED
..
Curve A shows that parasite drag is very low at Curve D shows total d rag and represents the
slow airspeeds and increases with higher air- sum of the other three cur'w'es. Point E identifies the
speeds. Parasite drag goes up at an increasing airspeed range at which total drag is lowest. That
rate at airspeeds above the midrange. airspeed is the best airspeed range for maximum
endurance, best rate of cl imb, and minimum rate
of descent in autorotation.

Curve B shows the profile drag curve. Profile


drag remains relatively constant throughout the
speed range with some increase at the higher A line is drawn from th e point of origin tangent
airspeeds. to the total drag curve. Ttle point of tangency (F),
identifies the airspeed range which results in
maximum glide distance in autorotation and
maximum range.
Curve C illustrates how induced drag decreases
as aircraft airspeed increases. At a hover or at
lower airspeeds, induced drag is highest. It
decreases as airspeed increases, and the heli-
copter moves into undisturbed air.

Figure 2-25. Helicopter drag and airspeed re lationship

2-29
P ARASlI'E DRAG
2-16.

Parasite drag refers to all drag on an aircraft not associated with the
production of the lift force. (Parasite drag is also discussed briefly in
paragraph 2-12.) Parasite drag is created by--
a
Displacement of air by the aircraft (form drag).
0
Hindrance of the airflow as it passes over the surface of the aircraft
(skin-friction drag).
0
Turbulence generated in the airstream (interference drag).

These factors create drag forces that are due to the movement of an object
through the air mass rather than from the production of The airstream lift.
velocity affects these types of drag; the higher the airstream velocity, the
greater the effect. The parasitic drag on main and tail rotor blades is
known as profile drag, which is due to form, skin-friction, and interference
drag acting on the airfoil.

a. Components of Parasite Drag.

(1) Form drag. The portion of the parasite drag that is generated
because of the shape of the aircraft or airfoil is called form drag. Figure
2-26 snows a flat circular disk (part A) and an aerodynamic shape (part B)
placed in an airstream. The streamlines flow smoothly over the aerodynamic
snape with little or no turbulence generated. Around the disk, however, the
streamlines break down and turbulence is generated; this lowers the static
pressure behind the disk. The static pressure is higher on the front of the
disk than on the back. The net result is a force (drag) that tends to retard
motion. In actual tests, the drag force on an aerodynamic shape is about
one-third of the drag on a disk with the same cross-sectional area. This is
because the air is displaced gradually and flows smoothly over the shape
without creating turbulence. To decrease form drag, most aircraft fuselages,
engine nacelles, wing pods, and other components exposed to the airstream are
generally shaped like the teardrop shown in part B, Figure 2-26.

A FLAT PLATE B AERODYNAMIC SHAPE

.~~~
-~.
~/j;/~
\\.V;Br-.r; ')~
~
.

~ .~~ --=--
~//////
~
//~-
~--p---=
Figure 2-26.

2--30
..
Form drag
(2) Skin-friction drag.
(a) Covering the entire "wetted surface" of the aircraft is a
tain layer of air called the boundary layer. Figure 2-27 shows air flowing
over a surface and the velocity profile of the air within the boundary layer.
Air molecules on the s\Jrface nave zero velocity in relation to the surface.
Ho~ever, the layer above moves over the stagnant molecules on the surface,
because it is pulled along by the third layer. Velocities of the layers
increase as distance from the surface increases until the free-stream
velocity is reached. when the velocity of the molecules reaches 99 percent
of the velocity of the flow above the airfoil, the end of the boundary layer
has been reached. At subsonic velocities, the boundary layer is about as
thick as a
playing card. The various layers of air within the boundary layer
slide over one anotner and create a drag force. This force retards motion
because of [he viscosity of the air. This drag force, called skin-friction
drag, is very small per square foot of surface area. When applied to large
areas of a wing or rotor blade, however, the force can become a significant
part oí parasite drag.
(b) Aluminum exposed to the atmosphere develops a coating of
aluminum oxide. This coating causes the surface to become rough and pitted
and offers appreciable resistance to smooth airflow. The boundary layer is
tnicker and, though the viscosity of the airstream is constant, more layers
of air slide over one another, causing a higher skin friction. The best way
to reduce this effect is to smooth the surface with wax and polish. Paint
hel~s to smooth the surface; but if toe painted surface is left unattended,
it can oxidize like aluminum and increase the skin-friction drag. Most
large, high-speed aircraft that cruise with parasite drag as the predominant
drag have highly polished surfaces to reduce the skin-friction effects.

I
I
I
I
I
I
I
IFREE-STREAM
\ VELOCITY

..
If ,.
,.,.....- --
---
,.
'f:".t:'~.:"~.:- ''<'';:;'' :.:~.~
.- ..~. ".'.
'-1------
BOUNDARY

......."'
LAYER

".: "~;:'-AIRFOILSURFACE

Figure t.-27. Boundary layer

2-31
(3) Interference drag. Interference drag is generated by the
collision of airstreams, creating eddy currents, turbulence, or restrictions
to smoot~ flow. For example, the air flowing along the fuselage collides
with the air flowing over the wing in the area of the wing root, as shown in
Figure 2-28. The effects of this collision can be reduced by allowing the
two air currents to merge smoothly. The wing is faired into the fuselage
junction area to reduce interference drag; this allows the airstreams to meet
gradually instead of abruptly, reducing the turbulence formed. With a fuel
tank hung on the wing of an aircraft, the drag produced is actually higher
than the sum of the drag of the components when computed separately. Because
of the interference drag created, the airstream must flow around the tank and
the wing. Again, fairing the area between the wing and the fuel tank or
increasing the distance between the two will lower the effect of interference
drag.

Figure 2-28. Reducing interference drag

b. Equivalent Parasite Area and Parasite Drag Equation. Force is equal


to pressure times area. Parasite drag, as a force, can be computed by
multiplying dynamic pressure by the area. The area used is called the
equivalent parasite area. It is a computed area measured in square feet; it
is not the cross-sectional area of the aircraft. The difference between
these two concepts can be illustrated by comparing an object--a barn door--to
an aircraft. The barn door has a smaller surface area than the cross-
s~ctional area of the aircraft. rlowever, the door may have the same
equivalent parasite area as the aircraft. In this example, the barn door
needs to be visualized with its broad surface facing into an airstream. If
the barn door and the aircraft travel at the same velocity, they will have
tne same parasite drag; the equivalent parasite area will be the same.
Although the aircraft has a larger cross-sectional area than the barn door,
~ts equivalent parasite area in this example is the same.

2-32
The aircraft is streamlined and, therefore, offers less resistance to airflow
than the flat surface of the door. Parasite drag is computed as follows:
V2 (Equation 2.8)
Dp =
1/2Pf
where f =
equivalent parasite area (in square feet)

c. Parasite Drag and Velocity. Parasite drag varies wit~ the square of
A change in the
the velocity and, therefore, predominates at high speeds.
configuration of the aircraft, such as lowering the gear, increases the
equivalent parasite area and increases the parasite drag at a g1ven ~elocity.
Figure 2-29 shows how parasite drag varies with the velocity.

4,000

V
/
3,000

~ /
C)
<
J
~ 2,000
w
~

<
/
a: ,
<

I
Q.

1,000
V
/
'
V
---- 300
0 100 200
VELOCITY (TAS)

Figure 2-29. Parasite drag curve

2-33
2-17. INDUCED DRAG

Ihe portion of the total drag force that is due to the production of the lift
force is called induced drag. This drag is induced as the wing develops
To simplify
lift. A discussion of induced drag have
can become quite technical.
been Qade and induced drag is pre-
the discussion, certain assumptions
sented only as it concerns aviators.

a. Vortex. When a wing produces lift, a pressure d~ffere~tial is


Bound
created across it; this differential causes pressure below atmospheric on the
top surface and a higher static pressure on the bOttOIil surface. This pres-
sure differential induces a circulation about the wing; this circulation is
called bound vortex. The high-pressure air tends to flow to the top surface
ot the wing or to the area of low-pressure, as shown in part A, Figure 2-30.
The wing in the figure has an infinite span and no wingtips for the air to
tluw around. The upwash ahead of the wing is equal to the downwash be~ind
it, so velocities cancel each other. On this infinite span, there is no
induced drag. Part ß, Figure 2-30, shows the plot of the vertical-velocity
vectors io the vicinity of the aerodynamic center. No downwash exists at the
aerodynamic center, the point where the aerodynamic force is generated. For
this reason, there is no induced drag in pure two-dimensional flow.

~RROW

,
~

--_.-'
8

Figure 2-30. Bound vortex

Trailing-Edge Vortexes. Although it would be desirable to design


b. a

wing without induced drag, a wing with an infinite s?an could not be con-
structed. ~ot only does the air flow over the leading edge to create the

2-34
circulation mentioned in a above, but it also flows around and over the wing-
tips. As the wing ,naves through the air mass, the air trying to flow around
the wingtip causes a vortex behind
it. This wingtip vortex induces a span-
wise flow and creates vortexes all along the trailing edge of the wing. As
shown in Figure 2-31, the trailing-edge vortexes are strongest at the tips
and progressively diminish in intensity as they approach the centerline of
the wing. At the ceùterline of the wing, no trailing-edge vortex exists
because (from the cockpit position) the right-wing vortexes revolve counter-
clockwise and the left-wing vortexes revolve clockwise. The tip vortexes
cancel each other at the centerline.

LEFT-WING VORTEXES

/--
TRAILING
EDGE

TRAILING EDGE

- =
negative
+ =
positive

6 ~
~,.)
~ 3
Figure 2-31. Trailing-edge vortexes

c. Downwash.

(1)
The combination of the hound vortex and the trailing-edge
vortexes produces vertical velocities, as shown in Figure 2-32. The dashed
line shows the path of the air mass as it flows over the airfoil. The
downwash velocity is at the aerodynamic center of the finite wing. The
downwash vector (w) added to the FSRW vector results in another relative-wind
vector that is inclined to the actual flight path. From wingtip to wingtip,
the magnitude of the downwash vector (w) varies as the intensities of the
trailing-edge vortexes vary.

2-35
FINAL
DOWNWASH
VELOCITY

(2w)
------ ---

Figure 2-32. Vertical-velocity vectors (finite wing)

(2) Part Figure 2-33, shows the vector diagram, using an average
A,
downwash-velocity vector. The downwash vector has been reversed and added to
the opposite end of the FSRW vector to simplify the diagram. The average
relative wind, the flow that actually affects the wing, is inclined to the
The force that is produced by the
FSRW at an angle of
ai, the induced angle.
wing, labeled airfoil lift (L), is perpendicular to the average relative
wind. The effective lift force is perpendicular to the FSRW. The component
to the FSRw is the induced drag.
af tne resultant force that is parallel

(3) The total aerodynamic force (part B, Figure 2-33) includes


airfoil lift, induced drag, and parasite or profile drag.

2-36
I Dj I AIRFOIL Dt
- Þ
SECTION ANGLE LIFT (L) D' Up
OF ATTACK EFFECTIVE
LIFT I~ TAF

A
FSRW OR
ROTATIONAL RW
TOTAL DRAG
B

Figure 2-33. Drag vector

d. Angle of Attack and Induced Drag. As the angle of attack is

increased, the pressure differential between the bottom and top surfaces of
the airfoil increases. This increases the downwash, which then increases the
induced angle. The result is a greater angle bet~een the effective and
airfoil lift vectors and, therefore, an increase in the induced drag.

e. Velocity and Induced Drag. It is generally accepted that drag


increases as airspeed increases. This is true of total drag. However, total
drag includes both induced drag and parasite drag. The general effect of
speed on induced drag is unusual since low airspeeds are associated with high
lift coefficients which create high induced drag coeff~~ients. The immediate
implication is that induced drag increases as airspeed decreases.
(1) If all other factors are held constant to single out the effect
of airspeed, it is found that induced drag varies inversely with the square
of the airspeed. This effect implies that an airplane in steady flight will
incur one-fourth as much induced drag if the airspeed is doubled or four
times as much induced drag if the airspeed is reduced to one-half the
ariginal value. For example, an airplane in steady, level flight slows from
300 knots to 150 knots; the dynamic pressure becomes one-fourth as great, and
the wing must deflect the airstream four times as much to create lift equal
to that at 300 knots. Therefore, induced drag is of greater importance in
flights at low speeds but is almost insignificant with the high dynamic
pressures associated with higher airspeeds.

(2) In level flight at low altitude and maximum airspeed, a typical


single-engine airplane or a helicopter has an induced drag that is less than
10 percent of the total drag. However, this same airplane in steady flight

2-37
just aDove the stall
speed or the helicopter at a hover could have an induced
drag that is about 75 percent of the total drag. Figure 2-34 shows how
induced drag decreases as airspeed increases when the airfoil is producing a
constant amount of lift.

4,000

3,000
~

~
<(
I%:
C 2,000
C
w

\
U
:)
C
Z

1,000

"- ~
to-..

0 100 200 300 400 500

VELOCITY (T AS)

Figure 2-34. Induced drag plotted against velocity

2-38
SECTION V

PERFORMANCE CURVES

2-18. CURVE CONSTRUCTION

a. An aviator
must know what type of performance he can expect from his
aircraft; for example, maximum velocity, maximum rate of climb, and maX1mum
angle of climb. The performance capability of any aircraft depends on the
relationship between four forces acting on the aircraft--thrust, weight,
lift, and drag. To understand this relationship and the way one force can
change another and thereby affect aircraft performance requires a careful
study of performance curves. Performance capabilities that can be obtained
from such curves are--
0
Hover power.
0
Stalling speed (Vs).
0
Absolute aircraft ceiling.
0
Velocity for maximum range.
0
Obstacle clearance distance.
0
Maximum sustainable G-Ioading.
0
Velocity for maximum endurance.

0
Maximum aircraft velocity (Vne).
0
Velocity for maximum rate of climb (Vy).
0
Velocity for maximum angle of climb (Vx).
A
more detailed discussion on the construction of performance curves for
rotary-wing aircraft is found in Chapter 6; for fixed-wing aircraft, in
Chapter 8. Performance curves for aircraft are found in the specific -10
operator's manual.

b. Any change in weight, altitude, temperature, or drag causes a change


in performance curves. The curves constructed in tnis manual are for a defi-
nite altitude, weight, and aircraft configuration. To simplify the construc-
tion and understanding of the performance curves, a simple force system is
assumed; that is, lift equals weight and thrust equals drag, as shown on
Figure 2-35. The performance of an aircraft is primarily concerned with the
thrust/drag relationship when the weight and lift remain constant throughout
the velocity range of the aircraft.

2-39
L = 10,000 Ib

D ~
1~OOlb

<
W =
1,000 Ib

Figure 2-35. Aircraft in equilibrium

2-19. PO~ER REQUIREMENTS

a. Since the drag force is the basis of performance, a plot of the total
drag curve is the place to start. The total drag at any velocity is de-
scribed as the sum of the parasite drag (including profile drag for heli-
copters) and induced drag at that velocity. The point at the bottom of the
curve (dragminimum) is where the aircraft is operating at its maximum
efficiency and is at an angle of attack for (L/D)maximum.
b. The total power required for unaccelerated flight is the sum of--
0
The power required to produce lift equal to weight (induced power).
a
The power required to produce a
thrust force equal to parasite drag
(parasite power).
a
The power required to turn the rotor systems of helicopters (profile
power).

Some additional power is required to operate auxiliary equipment and aircraft


systems and to overcome friction. Only power required to produce lift and
thrust is discussed in this manual.

c. Parasite drag affects the amount of thrust required to maintain


a

given velocity. If an aircraft with a


level flight velocity of 200 knots
develops a parasite drag force of 750 pounds, the thrust required to maintain
200 knots must be 750 pounds. The thrust required to maintain steady,
unaccelerated level flight is equal to the parasite drag. (The parasite drag
equation was developed in paragraph 2-13.)

2-40
d. To maintain a constant altitude, the airfoils must produce lift equal
to the aircraft weight. Aircraft engines must have enough power to overcome
the induced drag that results from producing this lift.

2-20. POWER EQU ATlONS

The production of power is not directly related to the fuel flow of a

horsepower-prodllcing engine (reciprocating or turbine) as it is of jet a

engine. Power is simply the rate of doing work. The units of power are the
units of work (in foot-pounds) divided by time (in seconds). The term
horsepower was originated by inventor James Watt. Horsepower was calculated
as the amount of work a horse could do in a certain length of time. The
do 550 of work in 1 second.
results showed the horse could foot-pounds

a. The power and horsepower equations are expressed as follows:

Power
Work (in ft-lb) (Equation 2.9)
Time (in see)

(in ft-lb)
Horsepower =
Power (in sec ) (Equation 2.10)
550

b. The product of drag and velocity yields the units of power in the
following equation for power required and horsepower required:
PR = DV (Equation 2.11)

where PR =
power required (in foot-pounds per second)
D =
total drag (in pounds)
V =
velocity (in feet per second or knots)

DV (V in ft/sec)
HPR =

550
-
-

Tv
R550

(or) (Equation 2.12)

DV (V in knots)
HPR =

325

where HPR =
horsepower required
TR =
thrust required equal to drag (parasite, profile, and
induced)

2-41
2-21. POWER-REQUIRED CURVES

An aircraft developing of drag at 325 knots requires 1 horsepower to


1 pound

maintain that velocity (DV/325). The horsepower-required curve can be


plotted simply by multiplying total drag by velocity and dividing by 325.

a. The horsepower-required curve shown in Figure 2-36 was developed from


a
hypothetical thrust-required curve. This curve represents the horsepower
required for this hypothetical propeller-powered aircraft.

b. The fuel flow for the reciprocating engine varies with the power
produced. The propeller converts this power into the thrust required to move
the aircraft. Thrust from the propeller is not constant. In Figure 2-36,
the curve shows that the power required to maintain 200 knots is approxi-
mately 1,000 horsepower. To develop this horsepower, an engine with a
propeller operating at 80 percent efficiency must develop 1,250 horsepower.
This requires a fuel flow of 500 pounds per hour. A fuel-flow scale has been
added to Figure 2-36 to correspond with the HPR curve. Therefore, this curve
can also be considered a fuel-flow-required curve for this propeller-powered
aircraft.
-..--
~
2,000 4,000 f."E.\..l~9
\,;
st.'"
O~t.90
fÙ\..\..9

>I~
c ~

1,500 II 3,000
Q
w
a::
5
0
w
a::
a::
1,000 w 2,000
~ PROP
0 AIRPLANE DATA
0.
w WEIGHT 20,000 LB
I/) WING AREA(s) 500 FT2
a::
0
:2: SEA LEVEL DENSITY a = 1
500 1,000

....
~

'-
..c

~
...J
LL.
...J 100 200 300
UJ 0
:::::>

LL. VELOCITY (T AS)

ANGLE OF ATTACK I
180 120 70 10

Figure 2-36. Horsepower-required curve (fixed-wing aircraft)

2-.42
c.
A
typical horsepower-reqûired curve Íor rotary-wing aircraft is shown
in Figure 2-37. A profile power-required curve is also included.

2,000

1,750

1,500

PR }-o
1,250 }-<f
(,
Po w~
II REQUIRED
1,000

750

500 PROFILE

250

10 20 30 40 50 60 70 80 90 100 110

VELOCITY (TAS)

Figure 2-37. Horsepower-required curve (rotary-wing aircraft)


2-22. POWER-AVAILABLE EQU~rION

Equation 2.12 shows that the horsepower required is equal to the product of
the drag times velocity divided by 325. A similar equation can be written

concerning the thrust developed by the propell~r or rotor. In this case, the
result is horsepower available. The HPA equation is expressed as follows:

T V
HP A =
325 (Equation 2.13)

where HPA =
horsepower available
T =
thrust (in pounds)
V =
velocity (in knots)

2-43
2-23. POV1ER-AVAILABLE CURVES

The three horsepower-available curves in Figure 2-36 represent full power


at sea level, partial power at sea level, and full power at the absolute
ceiling. An important distinction concerns the absolute-ceiling curve. The
decrease in engine-power output resulting from the decrease in air density at
high altitude lowers the HP A curve until it just touches the HPRminimum point
on the curve. This point is not to be confused with the (L/D)maximum angle
of attack (see angle of attack scale). This is the velocity where the horse-
power required is at a minimum, which is a product of drag times velocity.

2-24. MAXIMUM ENDURh~CE

The endurance of a propeller- or rotor-powered aircraft occurs at the


velocity where the fuel flow is minimum. This is the lowest point on the HPR
curve, as shown in Figure 2-38. At this velocity, the aircraft flies at an
angle of attack that corresponds to the equation below. Horsepower for
maximum endurance can be computed by using the following equation:

HPRmax
=
3/2)
(CL
CD (Equation 2.14)
max
where HPRmax =
horsepower required for maximum endurance
CL =
coefficient of lift
CD =
coefficient of drag
2-25. MAXHmM RANGE

The range of propeller- or rotor-powered aircraft is found at the velocity


where the ratio of FF/V is at its Ininimum value. This is shown at the point
of tangency of the horsepower-required curve and the straight line drawn from
the origin. The straight line is the minimum slope of FF/V that can be drawn
and still contact the performance curve. At this point of tangency, certain
relationships, as shown below, hold true. The maximum-range equation is
expressed as follows:

(~R)min
(~F)min -

(Equation 2.15)

where FF =
fuel flow (in lb/hr)
V =
velocity (in knots)
HPR =
horsepower required

Since HPR =
DV, then DV/V must be minimum. The velocity cancels, leaving
only the drag, which must be a minimum. This being true, the angle of attack
for maximum-range performance of a propeller-powered aircraft will yield
(L/ù)maximum.

2-44
2,000 4,000

PROP
AIRPLANE DATA

WEIGHT 20,000 LB
WING AREAS(S) 500 F"P
SEA LEVEL DENSITY Ó= 1

.....
~tl:
1,500 3,000
~
>I~
o~ ~
~
.;:)
II
Q ø
ILl tl:
MAX ENDURANCE
II:
5
0 FF min ~ HPR min 1:1: DVmin. (Cl 312) ~
0
$
.: CD max
ILl t
1,000 II: 2,000 ~
II:
ILl
~
R
0
~
ILl
f/)
II:
0
:%: MAX ENDURANCE

500 1,000
(~ 3/2)
CD max
RANGE MAX
'C'

(~min (H:~min
&.
-- =
g,
3=
0 ~(D~min .: (L)
D max
..I
u.. T.
...I
W
0 100 200 300
::)
L\.
VELOCITY (TAS)
, I I I
10
180 1~ 7"

Figure 2-38. Maximum-endurance curve (propeller)

SECTION VI

FORCES AFFECTING PERFO~~ANCE

2-26. VARIATION FACTORS

Performance curves, as mentioned previously, concern the relationship among


the four forces acting on an aircraft. Factors that vary this relationship

2-45
are weight, configuration, and altitude. These factors also vary the
location and shape of the performance curves. This section covers these
var1ations aad tneir effects on the endurance and range performance of
propeller-driven aircraft. Atmospheric lánds significantly affect range
performance; theretore, a brief discussion of this effect is also included.

ON
2-27. EF.Fr.;Cl' OJ:<'
wElGd'l' PERFOß.l'1ANCE

a. The effect of
a
change in weight on induced drag is easily explained.
Assuming lift is equal to weight, an increase in weight requires the same in-
crease in lift. If tnis is to be accomplished at a constant velocity, the
angle of attack must be increased; this in turn increases the CL. Since
induced drag is directly proportional to CL squared, this increase in the
angle of attack increases the induced drag at the constant velocity.

b. The aviator
needs to know the airspeed that will result in the angle
of attacK for the maximum lift-to~drag ratio--(L/D)maximum; however, a change
in the performance curve will change the airspeed at which the aircraft can
maintain lift
equal to weight at that angle of attack. With an increase in
t~e gross weight of tne aircraft, the (L/D)maximun velocity is still located
at the velocity where drag is at a minimum.

c. The ratio of ,L/D)maximuT:l is still the same. The CL and CD


characteristic curves of the aircraft do not change if weight is increased.
Since the LID ratio is equal to the CL/CO for any angle of attack, the
(L/D)maximum ratio is the sallie for all weights; it occurs at the same angle
of attack, but at a higher velocity. If the weight of the aircraft
increases, the lift must increase.
A weight
d. increase influences all performance points found on the
power-required curves, with the possible exception of Vmaximum. Endurance
performance of propeller aircraft decreases because of the higher drag force
requiring a higher fuel flow to maintain the angle of attack for maximum
endurance.

e. The range and endurance of propeller aircraft decrease with increases


in gross weight. However, these specific points of performance are obtained
at definite angles of attack, which do not change with variations in the
weight of the aircraft.
f. The discussion above also applies to a helicopter hovering in a no-
wind condition. In forward flight, however, a differential exists in the
velocity of the airflow across a rotor system. Because of this differential,
some other means of increasing the total rotor lift must be employed. This
is discussed in Chapter 6.

2-28. EFFECT OF CONFIGURATION ON PERFO~~ANCE

A
a. change in the configuration of an aircraft does not have the same
effect on performance curves as a change in the weight does. The configura-
tion of the aircraft can be changed, for example, by lowering the landing

2-46
A change in con-
speed brakes.
gear, adding external stores, or extending by the
t15uration changes the equivalent parasite area. This area multiplied
dynam1c pressure ot tne airstream is the parasite drag.

b. A change in the configuration of an produces change only in


aircraft
the parasite drag. Therefore, the performance curves will show the greatest
variations at higher velocities where parasite drag is predominant. With
increased drag, the aircraft requires more power or thrust to maintain
a

and
certain velocity. This increased drag also adversely affects range
endurance performance because of the higher fuel consumption associated with
increased power requirements.

'l.-'L.~. EFFë..Cr OF ALnTUD~ ON PERFORM.A1~CE

This means aircraft


a. As altitude increases, air density decreases.
may increase in velocity (true airspeed) without the drag force increasing
(ma1ntaining Cu constant).

b. Figure 2-39 snows the effect of altitude on the range of aircraft


powered by reciprocating engines. A
flight conducted at high altitude has a
greater true airspeed and requires proportionately greater power than
a

tlight conducted at sea level. The drag of an aircraft at altitude a is the


same as the drag at sea level, but the higher true airspeed causes propor-
tionately greater power requirement. The straight line tangent to the sea-
level power curve is also tangent to the altitude power curve.

EFFECT OF ALTITUDE

SEA LEVEL.

CONSTANT
WEIGHT

POWER SPEED(KT)
REQUIRED
(HP)

Figure 2-39. The effect of altitude on range

c. If a change in altitude causes identical changes in speed


and power,
the proportion of speed to power required is unchanged. This fact implies
that the specific range of the aircraft is unaffected by altitude. Actually,
this is true only when specific fuel consumption and propeller or rotor

2-47
efficiency are the principal factors that could cause a variation of specific
range with altitude. If compressibility effects are negligible, any varia-
tion of specific range with altitude is strictly a function of propeller or
rotor performance.

d. At lowlevels, an aircraft with a reciprocating engine experiences


little, if any, variation of specific range with altitude. There is a
negligible variation of brake-specific fuel consumption for values of brake
horsepower below the maximum cruise power rating of the engine; this rating
is the lean range of engine operation. An increase in altitude will produce
a
decrease in specific range only when the increased power requirement
exceeds the maximum cruise power rating. The advantage of supercharging is
that cruise power can be maintained at high altitude; the aircraft achieves
the high-altitude range with a corresponding increase in true airspeed.
Principal differences in high- and low-altitude cruise are the true airspeeds
and the fuel requirements for climb.

e. Anincrease in altitude has a different effect on the performance of


turbine-powered aircraft. The turbine power plant operates best at rela-
tively high power outputs. This may not be possible when it is operating at
low altitudes if the proper velocity is to be maintained for maximum specific
range. Furthermore, the turbine shows improved specific fuel consumption as
inlet temperatures are reduced, which normally occurs at altitude. The aero-
dynamic characteristics of an aircraft indicate that altitude is not desir-
able from a specific-range standpoint. If the aircraft is turbine-powered,
fuel consumption will improve with altitude.

f. If low-altitude or "off-optimum" performance is required, it is


sometimes oecessary to fly at other than optimum-range velocities. The
turbine overbalances the aerodynamic characteristics of the aircraft because
of its preference for high power outputs for minimum-specific fuel consump-
tion. For correct balance, therefore, the aircraft must fly somewhat faster
for maximum range at low altitudes. Multiengine, turbine-powered aircraft
require a different solution. Probably one engine could be shut down and the
remaining engines operated at higher power. This could maintain the correct
balance between power and drag for the aerodynamic design of the airframe and
the high power output that the engine demands.

2-30. EFFECT OF WIND ON PERFORMANCE

Range performance is a basic comparison of fuel consumption to the distance


traveled over the earth; therefore, atmospheric winds must be considered
because of their effect on ground speed. A
plot of the velocity along the
horizontal axis is the true airspeed of an aircraft. The aircraft itself is
not concerned with the atmospheric wind. If an aircraft flies at its maximum
no-wind range airspeed of 130 knots into a head wind of 130 knots, it would
not be moving with respect to the terrain. Obviously, an aircraft flying
into a head wind is never able to attain range equal to its no-wind maximum
range; however, it can minimize the effects of the head wind by increasing
its airspeed to a velocity greater than the velocity of the head wind. ívind
has a significant effect on range when the velocity of the head wind or tail
wind becomes greater than 25 percent of the no-wind range airspeed.

2-48
CHAPTER 3

STABILITY

An aircraft stable to be effective. Stability decreases an aviator's


must be
work load and allows him to direct his attention elsewhere. Rotary-wing
aviators will find the section on fixed-wing stability helpful in under-
standing rotary-wing stability. Rotary-wing stability is discussed in the
second section of this chapter.

SECTION I

FIXED-WING STABILITY

3-1. MOTION SIGN CONVENTIONS

a. An aircraft has six directions of motion about three mutually


perpendicular axes, as shown in Figure 3-1. The three axes are the--
0
Vertical axis about which the aircraft yaws.
0

Lateral axis about which the aircraft pitches.


0
Longitudinal axis about which the airframe rolls.

Sign conventions are assigned to each motion. The right-hand rule


b.
can be applied to remember the signs. A right roll, right yaw, and a pitch-
up are all positive. For example, a positive rolling moment rolls the air-
craft to the right; a
negative rolling moment rolls the aircraft to the left.
VERTICAL
AXIS

POSITIVE ROLLING
MOMENT
(RIGHT ROLL)

POSITIVE PITCHING
MOMENT
(PITCH-UP)
POSITIVE YAWING
MOMENT
(RIGHT YAW)

Figure 3-1. Stability nomenclature


3-1
3-2. STATIC STABILITY

Static stability is the tendency an object possesses after it has been


displaced from its equilibrium. Newton's first law of motion alludes that if
the sum of the forces and moments about the CG of an object are equal to zero
no acceleration will take place. This condition is called equilibrium.

Positive Static Stability. If an object possesses positive static


a.
stability, it tends to return to its equilibrium position after it has been
moved. Point A in part 1 of Figure 3-2 shows the ball in equilibrium. If
the ball is moved to point B, it tends to roll back toward point A. This
tendency demonstrates positive static stability. The ball may not actually
return to point A, but it does tend to return. Therefore, it has positive
static stability.
Negative Static Stability. In part 2 of Figure 3-2, the bowl has
b.
been inverted. Again, point A is the equilibrium position. The ball has
been moved to point B. Now the ball tends to roll away from point A. This
tendency toward movement away from the equilibrium position demonstrates
negative static stability. The ball mayor may not roll away from point A,
but it does tend to roll away. Therefore, it has negative static stability.

I-
Z
w
:E
w
0
<
...I
EQUiliBRIUM
c..
VI
1. POSITIVE STATIC STABILITY is TIME-
4. NEGATIVE STATIC AND NEGATIVE DYNAMIC STABILITY

I-
Z
A w

I --, J :E
w
0
< EQUILIBRIUM
...I
c..
VI TIME~
is

5. NEUTRAL STATIC AND NEUTRAL DYNAMIC STABILITY


2. NEGATIVE STATIC STABILITY

I-
Z
w
:E
A B w
0 EQUILIBRIUM
1"
\ ) :5
c..
VI TIME-
is

3. NEUTRAL STATIC STABILITY


6. POSITIVE STATIC AND POSITIVE DYNAMIC STABILITY

Figure 3-2. Nonoscillatory motion

3-2
c. Neutral Static stability. In part
3 of Figure
3-2, the ball has been
placed on a
flat surface. When the ball is moved to point B, it neither
tends to return to nor roll away from point A. This demonstrates neutral
static stability.

3-3. DYNAMIC STABILITY

The word dynamic refers to motion. stability refers to the movement


Dynamic
of an object with respect to time. the dynamic stability of an object
When
is considered, the static stability of the object must also be considered.

a. Nonoscillatory Motion.

(1) Negative static and negative dynamic stability. As ment~oned


previously, an object that possesses negative static stability tends to move
away from its equilibrium position. As shown in part 4 of Figure 3-2, when
an object does move away from its equilibrium position, it has nonoscillatory
negative dynamic stability.

(2) Neutral static and neutral dynamic stability. As shown in


part of Figure 3-2, when an object that has been displaced does not move
5

toward or away from its equilibrium position, it has nonoscillatory neutral


dynamic stability.
(3) Positive static and positive dynamic stability. As shown in
part 6
of Figure 3-2, when an object has positive static stability and strong
positive dynamic stability, the result is nonoscillatory positive dynamic
stability. This motion is called deadbeat damping. Dynamic stability is
particularly stressful on the structure of an aircraft. This stress would be
similar to a
car on which the springs had been replaced with steel blocks
between the frame and the axle. When the car is driven over
a
railroad cros-
sing, the steel blocks do not give and the energy is absorbed at once. If
this condition were allowed to continue, material failure would eventually
occur.

b.Oscillatory Motion. To have oscillatory motion, an object must


possess positive static stability. The following is a discussion of the
various types of dynamic stability coupled with positive static stability.

(1) Positive static and positive dynamic stability. If an object


that has been displaced tends to return to its equilibrium position, it has
positive static stability. As indicated by the elapsed time that is shown in
Figure 3-3, the object at part A moves toward its equilibrium position but
overshoots it. Because the object has positive static stability, it begins
to turn and move back toward its equilibrium position. This motion continues
but diminishes until the object comes to rest at its equilibrium position. A
decrease in the amplitude of the oscillations indicates that the object has
positive dynamic stability. It will eventually corne to rest in an equilib-
rium position.

3-3
(2) Positive static and neutral dynamic stability. As shown in
B of
part Figure 3-3, the elapsed time indicates neutral dynamic stability.
An object that has been displaced moves toward the equilibrium position and
overshoots it. Positive static stability makes the object move back toward
the equilibrium position. Again, the object overshoots the equilibrium posi-
tion, and its oscillations are equal to the oscillations in the first dis-
placement. As time passes, the amplitude of the oscillations is the same on
both sides of the equilibrium position. The object never comes to rest.
Because the amplitude of the oscillations neither increases nor decreases,
the object has neutral dynamic stability.

(3) Positive static and negative dynamic stability. As shown in


C
part of Figure 3-3, positive static stability makes the object oscillate as
it did in the first two examples. However, the increasing amplitude of the
oscillations as time passes indicates negative dynamic stability. Lines
drawn tangent to the top and bottom of each oscillation diverge.

I-
Z

A ~
ILl
POSITIVE STATIC AND (.)
cc
POSITIVE DYNAMIC STABILITY ~
m
TIME-
C

B
I-
Z
ILl
POSITIVE STATIC AND !
NEUTRAL DYNAMIC STABILITY ~
...I
0-
m
TIME-
C

c I-
Z
ILl
POSITIVE STATIC AND !
NEGATIVE DYNAMIC STABILITY ~ TIME
...I
0-
m
C
--.

Figure 3-3. Oscillatory motion

3-4
3-4. PITCH STABILITY

Pitch, or longitudinal, stability is the stability of the longitudinal axis


of the aircraft about the lateral axis. The following paragraphs discuss the
pitching motion about the lateral axis of the aircraft.

a. Oscillatory Pitching Motion.

(1) An well-designed, complex mass in motion. When


aircraft is a

moved from its equilibrium position, an aircraft develops a moment of iner-


tia. Ifa it has positive static stability, the aircraft becomes oscillatory
unless damping force prevents this motion. Moderate damping forces cause
the oscillation to converge with equilibrium. The period of the oscillation
is a function of the moment of inertia and the damping force.
(2) The period of the oscillation varies depending on aircraft
characteristics at a given airspeed. A long-term oscillation (more than five
seconds) is called phugoid motion (pronounced foogoid). Because the change
in pitch attitude is so slight, the gain or loss in altitude and the forward
movement of the aircraft result in this motion occurring at a relatively
constant angle of attack. An aviator makes pitch attitude corrections almost
subconsciously, just as a person makes minute steering corrections while
driving an automobile down a straight road.

short-term oscillation (0.3 to 1.5 seconds) is called a buzz.


(3) A
A buzz occurs so quickly that the stability of the aircraft dampens out the

motion before the aviator can normally respond.

medium-term oscillation (1.5 to 5 seconds) lasts about as long


(4) A

as the aviator's response time. This cah lead to a sudden and violent diver-
gence in pitch attitude. This divergence in pitch attitude can result in
large positive and negative load factors that the aviator makes larger when
he attempts to control the oscillation. This is called pilot-induced oscil-
lation. During the landing phase, it is called porpoising.
(5) The phugoid motion is relatively unimportant in stability
considerations, since the angle of attack is essentially constant. Shorter
periods of oscillation, however, must be properly dampened in a flyable
aircraft.
b. Pitching Moments About the CG.

controllable aircraft has positive static longitudinal


(1) A

stability positive dynamic longitudinal stability. Dynamic stability


and
that involves motion with respect to time is not fully covered in this field
manual. This section primarily covers static stability. By considering the
pitching moments about the aircraft's center of gravity, any tendency the
aircraft has when it is displaced from its equilibrium position can be
analyzed.

(2) The CM. which is called the pitching moment coefficient, comes
from the pitching moment equation. The sign of the pitching moment

3-5
coefficient indicates whether a pitching moment will pitch the nose of the
aircraft up (+) or down (-). This section of the manual discusses the direc-
tion of the pitching moments created by the various components of the air-
craft. It does not discuss the magnitude of the pitching moment; therefore,
the pitching moment equation will not be explained further in this section.

(3) Pitching moments about the CG of the aircraft are caused by


changes in the total lift as it is distributed among the wings, fuselage, and
tail surfaces. Total lift acts through the AC of the entire aircraft. The
AC of the
aircraft is called the neutral point. To simplify the discussion,
drag changes and compressibility effects are omitted. If the controls are
fixed at the trim position, a constant angle of attack results and a zero
pitching moment exists. As shown in Figure 3-4, this is called the trim
point.

~M

~L .
+CL
.
.
.
.
.
.
.
.
A
......................

~M

Figure 3-4. CM versus CL

(4) Figure 3-4 also shows the variation of CM with respect to


changes in CL. The CL can be used instead of angle of attack because their
relationship is linear, except aSCLmaximum is approached. (At any angle of
attack above CLmaximum' the airfoil begins to stall.) At the value of CL
where the curve crosses the horizontal axis, the pitching moments are zero.
At this angle of attack (trim point), stability considerations-are made.

(5) If the angle of attack is increased to a higher value of CL than


that indicated at the trim point, a negative pitching moment must be present
to return the aircraft to the trim angle of attack. This is shown at point A
in Figure 3-4. The opposite is also true. If the angle of attack decreases

3-6
which tends
from the trim point, a positive pitching moment must be created
to return the aircraft to the trim point. This is shown at point B in
Figure 3-4. Therefore, for an aircraft to exhibit positive longitudinal
stability, the slope of CM versus the CL curve must be negative. The degree
A steeper slope shows
of the slope indicates the degree of stability.
stronger pitching moments with changes in CL; therefore, greater stability
exists.
(6) The location of the trim point is important in aircraft design.
The trim point must occur at some usable angle of attack--between that of
To satisfy the preceding
zero lift and the stalling angle of attack.
requirements, the neutral point, or aerodynamic center of the aircraft, must
be aft of the CG of the aircraft (Figure 3-5). If a sudden gust pitches the
aircraft to a higher angle of attack, the increase in overall a lift
of the
negative
aircraft, which acts through the aerodynamic center, creates
pitching moment. This tends to return the aircraft to its equilibrium posi-
A gust that pitches the nose of the aircraft down causes
a
decrease in
tion.
the angle of attack and an overall net decrease in lift forces. This results
the
in a pitching moment about the CG of the aircraft, which tends to return
aircraft to equilibrium.

CENTER OF GRAVITY
AERODYNAMIC CENTER

Figure 3-5. Fixed-wing aircraft CG and AC

c. Wing Contribution.

(1) The overall static stability of the aircraft depends on the


the
position of the CG in relation to the AC of the aircraft. All parts of
The moments contributed by the
aircraft contribute to its static stability. the other
wing or other parts depend on the location of the
AC of the wing or
parts of the aircraft being considered in relation to the CG. Together,
these moments determine where the neutral point is located. If the airflow
the AC of the wing is approximately at the 25 percent
is incompressible,
chord of the wing. In part A of Figure 3-6, the aircraft's
CG is behind the
AC of the wing; an external disturbance that pitches the wing to
a higher
angle of attack increases the pitching moment toward the stalling angle of
attack. This additional increase in angle of attack also increases the lift
and the pitching moment. Unless another force counters the effect, the air-
and As shown in part B of Figure 3-6,
craft will pitch upward again again.

3-7
when the AC of the wing is aft of the CG
of the aircraft, longitudinal
stability is enhanced. if
Similarly,
the CG of the
aircraft and the AC of
the wing coincide, the wing contributes neutral stability to the
aircraft.

+CM INCREASED LIFT


CG INCREASED LIFT

( A UNSTABLE

-CMCG
B STABLE

Figure 3-6. Wing contribution to longitudinal stability

(2) Lift is normally considered to act through the AC of an


airfoil.
Actually, the aerodynamic force, of which lift is a component, acts through
the center of pressure. As angles of attack change, the center of
pressure
moves back and forth on the
airfoil. Unequal pressure distribution on the
wing creates a moment. For positive-cambered wings, the MAC is in a negative
direction, nose down. The moment about the aerodynamic center of the wing
should not be confused with the moment about the center of
gravity of the
aircraft (Figure 3-7).
TOP SURFACE LIFT

-CM
AC

BOTTOM SURFACE LIFT

Figure 3-7. Negative pitching moment about AC of a

positive-cambered airfoil

3-8
(3) An positive-cambered airfoil and a CG forward of
aircraft with a

its aerodynamic center cannot be trimmed unless the negative moment is bal-
anced by a positive moment about the CG. This balance can be accomplished
only when the value of CL is negative. At any greater value of CL, the net
result of the moments will be negative (Figure 3-8). For a positive-cambered
wing to contribute positive stability to the aircraft, it must be trimmed at
an unusable angle of attack. The. negative moment can be overcome by a posi-
tive moment from the horizontal tail.

+CM
-CMAC

~-~
+CL
-CL

USELESS TRIM POINT


+CMCG
NEGATIVE
SLOPE
-CM STABLE

Figure 3-8. Positive longitudinal stability of a

positive-cambered airfoil

(4) this same wing has the CG of the aircraft located behind its
If
AC, then a positive lift force at the AC creates
a
positive pitching moment.
This positive pitching ~oment balances the negative moment about the AC that
is present in cambered airfoils. The wing can be trimmed at an angle of
0 and below the stalling angle of attack (Figure 3-9).
attack above CL =

However, the wing contributes negatively to the stability of the aircraft.


This can also be overcome by the horizontal tail.

3-9
POSITIVE
SLOPE
UNSTABLE
L

+CL

USEFUL
+CMCG TRIM
POINT

-CM

Negative longitudinal stability of


a
Figure 3-9.
positive-cambered airfoil

(5) When of the aircraft is forward of the AC of the wing, a


the CG

positive-cambered airfoil contributes positively to stability. However, the


aircraft can only be flown in the usable flight mode if horizontal tail
a

balances the moments. If the CG is aft of the AC of the wing, the CG con-
tributes negatively to the stability of the aircraft. However, the aircraft
can be balanced by a horizontal tail.

d. Fuselage and Engine Nacelle Contribution. A symmetrical body in a


perfect fluid at a positive angle develops pressure distributions, but no
resultant force exists. A streamlined fuselage is a good example. The air-
stream is not a perfect fluid, and the fuselage is not perfectly symmetrical.
However, the fuselage produces positive pitching moments at positive angles
of attack and negative pitching moments at negative angles of attack. In-
duced flow from the wing (upwash ahead of the wing and downwash behind the
wing) adds to the unstable contributions of the fuselage. An engine nacelle
located on the leading edge of the wing is also influenced by the wing upwash
and adds to longitudinal instability. The fuselage-engine combination has an
.
AC at about 25 percent of the fuselage length rearward from the nose of the
I

fuselage. Normally, no fuselage resultant force exists. Only


a pitching
moment that must be added to the moments created by the wing about the CG
exists. The fuselage-engine AC is placed in such a
position relative to the
aircraft AC so that it makes negative contribution to aircraft longitudinal
a

3-10
by using the horizontal
stability. This negative contribution is corrected
stabilizer.

e. Horizontal stabilizer Contribution.

(1) horizontal stabilizer, which is usually


The
a
symmetrical
located well aft of the CG of the aircraft. Because the airfoil
airfoil, is
is symmetrical, it can produce either positive or negative lift, depending on
the angle of attack. The entire horizontal stabilizer is located behind the
CG; therefore, its AC is aft of the CG, which creates
a
stable relationship.

(2) The stabilizing moment created by the horizontal stabilizer can


be controlled in two ways. The distance (moment arm) can be increased be-
tween the CG and the AC of the horizontal stabilizer, or the surface area
of
the stabilizer can be increased. Because the stabilizer is an airfoil, it
An increase in the area of the tail
proòuces lift
as the stabilizing force.
CG and the AC of the stabilizer increases the
or the distance between the
tail force and the stabilizing moment.
(3) If the aircraft pitches up to an angle of attack that is higher
than the trim angle of attack, the increased angle of attack on the hori-
This
zontal stabilizer increases the positive lift of the tail (Figure 3-10).
produces a negative pitching moment and returns the aircraft toward the equi-
of attack decreases, the hori-
librium trim point. If the aircraft angle
zontal stabilizer angle of attack also decreases and produces less, CGor even
This positive pitching moment about the and
negative, lift. creates a

returns the aircraft to equilibrium.

f. Wing, Fuselage, and Horizontalstabilizer Combination. Depending on


AC of the wing, the
the location of the CG of the aircraft relative to the
The hori-
wing may be stabilizing or destabilizing to the entire aircraft.
the trim problem of a
stabilizing wing
zontal stabilizer is used to overcome
(CG ahead of the AC of the wing). The horizontal stabilizer is also used to
AC of the
provide the stability needed with an unstable wing (CG aft of the
the
wing). Because the fuselage is destabilizing to the entire aircraft,
this negative contribution. To
ho~izontal stabilizer is also used to solve
of the entire aircraft must be lo-
create positive static stability, the AC

cated aft of the CG of the aircraft.

3-11
NEGATIVE
TAIL LIFT

POSITIVE
TAIL LIFT
+CMICG

-CMICG

CG

Figure 3-10. Lift as a


stabilizing moment of the horizontal stabilizer

g. Thrust Axis Contribution. The line along the thrust force vector is
called the thrust axis. If the thrust axis is located above the aircraft
center of gravity, an increase in thrust creates a negative pitching moment
(Figure 3-11). The horizontal stabilizer must also balance this moment. The
aviator must be able to trim the aircraft at any power setting. If the
thrust is located below the CG, opposite pitching moments are created when
thrust is increased.

ç I

*
~ ~ . T
)-CM
CG

-.. .'!jf~
Figure 3-11. Thrust axis about CG

3-12
3-5. DIRECTIONAL STABILITY

Directional stability involves the motion of the aircraft about the vertical
axis, or the yawing motion of the aircraft. Directional stability also in-
volves sideslip and the yawing moments produced about the CG because of

sideslip.

a. Sideslip angle, which is symbolized by using the


Sideslip Angle.
Greek letter ß (beta), is the angle between the relative wind and the longi-
tudinal axis of the aircraft. When the relative wind is to the right of the
nose of the aircraft, the sideslip angle is positive. Figure 3-12 shows a
positive sideslip angle.

LONGITUDINAL AXIS

Figure 3-12. Positive sideslip angle

b. Yawing Moments Versus Sideslip Angle.

(1)direction of the moments about the CG is discussed in this


The
section; the magnitude of these moments will not be developed. The symbol
for the coefficient of a yawing moment is CN. The sign denotes the direction
of the yawing moments developed by the various components of the aircraft.
The negative (-) sign is used for the left yawing moment, and the positive
(+) sign is used for the right yawing moment.

(2) The aircraft should be in directional equilibrium if the


relative wind is parallel to the nose of the aircraft or along the longitudi-
no sideslip exists, no yawing moment exists. However, if the
nal axis. If positive
positive sideslip angle, as shown in Figure 3-12,
a
aircraft has a

yawing moment is required for static directional stability. If the relative


wind is coming from the right, the aircraft should tend to yaw toward the
right. This yawing motion to the right will return the relative wind to the
nose of the aircraft and reestablish equilibrium.

3-13
c. Aircraft Component Contribution. Several aircraft components
contribute to directional stability. These components and their effects on
the aircraft's directional stability are discussed below.

( 1) Fuselage and engine nacelle contribution.

(a) In considering longitudinal stability, the fuselage and


engine nacelles were influenced by the upwash and downwash of the airstream
as it passed over the wing. This added to the instability created by the
fuselage. In directional stability, wing influences do not affect the yawing
moments created by the fuselage and the engine nacelles. The only consider-
ation is the side area of the fuselage and engine nacelles ahead of the CG
when that area is compared to the side area behind the CG. Most aircraft
have a
larger side area ahead of the CG than behind it. Therefore, a
rela-
tive wind that strikes the aircraft from either side causes a
larger yawing
moment ahead of the CG than behind it. This creates an unstable directional
condition because the aircraft yaws away from the relative wind, as shown in
Figure 3-13. In other words, a positive (right) sideslip angle on the fuse-
lage produces a negative (left) yawing moment.

-13 +/3

VERTICAL
AXIS

Figure 3-13. Directional instability ( ß versus -CN)

(b) To achieve directional stability, the side area of the


aircraft must be larger behind the center of gravity. Therefore, a
fin, or

3-14
vertical stabilizer, must be added to the fuselage to increase the area,
create a desirable yawing moment, and produce positive directional stability.

(2) Vertical stabilizer contribution.


(a) The a
sYmmetrical airfoil~ Like the
vertical stabilizer is
CG of the
horizontal stabilizer, it is located behind the aircraft. There-
fore, the aerodynamic center of the vertical stabilizer is located in posi-
a

tion that produces positive static directional stability.


(b) As with the horizontal stabilizer, the area of the vertical
stabilizer, or the distance from the CG, can be varied to obtain the desired
stabilizing moments. However, increasing the area of the vertical stabilizer
too much increases the height of the tail, which increases the frontal area
and the drag. To decrease the drag, the height of the tail is often de-
creased and a dorsal fin is added (Figure 3-14). This also decreases the
aspect ratio of the vertical stabilizer, which makes the tail effective at
higher angles of sideslip (Figure 3-15). Because the vertical stabilizer is
an airfoil, it is subject to aerodynamic stalls. Aerodynamic stalls can
occur at high sideslip angles. Adding the dorsal fin, which decreases the
aspect ratio of the tail, increases the stalling angle of attack. The tail
is then effective at the larger angles of sideslip. This is of particular
importance to a multiengine aircraft that might be subjected to large side-
slip angles because of aSYmmetrical power conditions; for example, an inop-
erative engine on one wing.

A AND B HAVE
SAME AREA B HAS LOWER EQUIVALENT
PARASITE AREAS (f)

Figure 3-14. Dorsal fin decreases drag

3-15
TAIL WITH
+CN DORSAL FIN

,....-------
TAIL

+(3
-(3

----_..- .-
-CN

Figure 3-15. CN versus ß for dorsal fin

(3) Wing contribution.The wing contribution to the directional


stability of the aircraft is small. However, becomes greater with in- it
creases in swept-back wing design. To be effective, this swept-back angle
must be at least 30 degrees.
(4) Final configuration contribution. Figure 3-16 shows the final
configuration of the aircraft. This aircraft configuration must have a
positive slope of the CN curve. Figure 3-16 also shows how the vertical
stabilizer must produce a stabilizing moment strong enough to overcome the
destabilizing moments generated by the fuselage and the engine nacelles.

VERTICAL STABILIZER
+CN
WITH DORSAL FIN
- -
-

,
- ...
VERTICAL
( STABILIZER

~ COMPLETE
-(3 +(3 AIRCRAFT WITH
DORSAL FIN
"
FUSELAGE
AND ENGINE
NACELLES
-CN

Figure 3-16. Fixed-wing aircraft configuration with +CN

3-16
3-6. LATERAL STABILITY

Lateral, or roll, stability involves the stability of the lateral axisA about
the longitudinal axis. Motion about the longitudinal axis is roll. right
roll is indicated with a
positive (+) sign; a
left roll, with a
negative (-)
sign. Wing design is important to aircraft stability. It is the primary
lift-producing surface and the primary roll-stabilizing surface. As with
directional stability, the aircraft achieves its roll stability through the
sideslip angle. With roll stability only, stabilizing rolling moments are
created by the sideslip acting on the wing.

a.Sideslip Caused by WinK Down. In Figure 3-17, the aircraft has its
right wing down. This tilts the vector of the wing to the right, so a
lift
horizontal component of lift
acts to the right. Because there is no opposing
force, this horizontal force moves the aircraft to the right. This motion to
the right, along with the aircraft's forward motion, produces a positive
sideslip angle. With the right wing down and a positive sideslip angle gen-
erated, the aircraft must develop a negative, or left, rolling moment for
positive static lateral stability. (Cl is the coefficient of the rolling
moment, not the coefficient of lift.) As shown in Figure 3-18,
a
curve of
the rolling moment versus the sideslip angle must have a
negative slope to
indicate positive static lateral stability. The curve must go through the
origin because a rolling moment should not be generated when the aircraft
wings are level and the relative wind is on the nose. The degree of slope,
as with other stabilizing curves, indicates the degree of the lateral static
stability of the aircraft.

./"

Figure 3-17. Horizontal-lift component produces sideslip

3-17
+C1

TRIM POINT

-ß +{3

STABLE

-C1

Figure 3-18. Positive static lateral stability (+ß Vs-Cl)

b. Dihedral.

(1) Definition. Dihedral of the wing is the angle between the wing
and plane that is parallel to the lateral axis (Figure 3-19). This angle
a

creates a stabilizing moment. When the wings droop, they have an anhedral, a
negative dihedral, or a cathedral angle. In other words, the wing produces a
destabilizing rolling moment.

I : DIHEDRAL
ANGLE
-- f
- - - - - - - - - - - - - - - - - - - -
- -

Figure 3-19. Dihedral angle

(2) Dihedral stability.

(a) To understand how dihedral produces a stabilizing moment,


a
three-dimensional picture is required. However, Figure 3-20 shows the
relative wind approaching the airfoil from the left side of the figure. The
aircraft is moving forward in a right sideslip. Therefore, the third
3-18
dimension must be added to the figure. The third dimension is the a relative
wind caused by the aircraft's forward velocity. The low wing
has higher
high This higher angle of attack gives the
angle of attack than the wing.
lower wing
lower wing a
larger coefficient of lift
than the higher wing. The
now creates more lift than the high wing. As a
result, a
negative rolling
between the two
moment is created because of the differential in lift forces
wings.
(b) The anhedral angle produces the opposite reaction. The
positive sideslip angle produces
a
positive rolling moment. Anhedral later-
ally destabilizes the aircraft.

.. RW

Figure 3-20. Dihedral stability

(3) Dihedral effects.

(a) Dihedral was the first in the construction of


method used
aircraft to gain lateral stability. Factors other than dihedral can contrib-
called
ute to the lateral stability of the aircraft. Their contributions are
dihedral effects. They can either stabilize or destabilize; therefore, their
contributions are classified as either positive or negative dihedral effects.
CG
(b) vertical location of the wing ain relation to the
The
affects lateral stability. ,A high wing contributes positive dihedral ef-
low wing contributes a negative dihedral effect. The wing position
fect; a
the large dihe-
has a
significant effect on lateral stability, which causes
dral angle normally used on aircraft with low wings. Wings mounted near the
CG have essentially no effect on the lateral stability of the aircraft.

3-19
(c) The vertical stabilizer can make a
slight, positive con-
tribution to the lateral stability of the aircraft. Because the vertical
stabilizer is a large area above the aircraft CG, the sideward force caused
by the sideslip angle produces a favorable rolling moment and helps stabilize
the aircraft laterally.

Total Aircraft and Positive Static Lateral Stability. The total


c.
aircraft must demonstrate a tendency toward positive static lateral sta-
bility. Some components might produce negative stabilizing moments. How-
ever, they must be overcome by stabilizing moments from some other component
of the aircraft so that the total aircraft is laterally stable.
CROSS EFFECTS AND STABILITY
3-7.
The sideslip angle is the main factor used to achieve directional and lateral
stability. Because yaws and rolls both produce sideslips, cross effects
between directional and lateral stability exist. A sideslip angle produces a
yawing and a rolling moment at the same time. The magnitude of the moments
and the inertia of the aircraft or its resistance to react to the moments
created can produce certain cross effects. Some of these effects are
desirable, some are not.

a. Adverse Yaw.

(1) Adverse yaw is produced by rolling the aircraft with the


ailerons. This is sometimes called adverse aileron yaw. The aircraft in
Figure 3-21 has its right aileron down and its left aileron up. This pro-
duces a
differential in the lift
force which acts on each wing and produces a
left roll. The right wing has a
higher coefficient of lift
because the down
aileron causes increased camber. Therefore, the induced drag is greater on
the right wing than ön the left wing. This increased drag causes the air-
craft to yaw toward the right about the CG. As the aircraft rolls, the rela-
tive wind resulting from the roll on the downgoing wing is upward (opposite
its direction of movement). This relative wind, when added vectorially to
the free-stream relative wind, resolves into an inclined relative-wind vec-
tor. Because the lift force produced by the downgoing wing is perpendicular
to its relative wind, the lift force acts forward. The opposite relative
wind must occur on the upgoing wing; therefore, its lift vector acts in a
rearward direction. The different directions of the lift forces produce a
condition that adds to the adverse yaw caused by the drag differential.

3-20
ROLL

AFT LIFT COMPONENT

UPGOING WING

FREE-STREAM RW
,

'~ ~~
~~...
: RW DUE TO ROLL
--"",
,--~I

RW DUE TO ROLL
-

~~"'~......J ,
-

-- ---
I

FREE-STREAM RW .'

Figure 3-21. Adverse yaw

(2) Modern yaw have controls to


aircraft that are subject to adverse
overcome the problem. They may be equipped with spoilers that extend on the
downgoing wing. They spoil some and add drag lift
to counter adverse yaw.
Differential ailerons are also used to control adverse yaw. The downgoing
wing aileron extends up more than the upgoing wing aileron goes down. This
produces more drag on the downgoing wing and counters the effects of adverse
yaw. Frise ailerons add drag to the downgoing wing. They extend the part of
the aileron forward of the hinge line and down into the airstream, while the
half of the aileron behind the hinge line extends up into the airstream.
Without these or other similar devices, aircraft using only ailerons for
lateral control tend to yaw to the right when they roll left and vice versa.
Even with compensating devices, adverse yaw is present during flight and is
usually corrected with a coordinated application of rudder.

b. Proverse Roll. Proverse roll is encountered when an aircraft yaws.


In this case, an aircraft is put in a right yaw when the aviator applies
right rudder. This creates a left sideslip angle. Since a negative sideslip
angle produces a positive rolling moment, the aircraft rolls to the right.
Because an aircraft wing cannot determine whether it is level, it responds to
a
negative sideslip with a positive roll. Another factor that contributes to
the proverse roll is the difference in the velocities of each wing. In the
example above, the left wing has a greater velocity than the right wing be-
cause of the yawing motion about the aircraft's CG. Increased velocity in-
creases the lift force on the left wing, which causes a
positive roll as long
as the aircraft is yawing.

3-21
c. Directional Divergent Stability.The degree of directional stability
compared with the degree of lateral stability of an aircraft can produce
three conditions. These conditions are directional divergence, spiral diver-
gence, and dutch roll.
(1) Directional divergence. Directional divergence results from
negative directional stability. This cannot be tolerated because directional
divergence allows the aircraft to increase its yaw after only a slight yaw
has occurred. This continues until the aircraft turns broadside to the
flight path or until it breaks up from the high-pressure load imposed on the
side of the aircraft.

(2) Spiral divergence. Spiral divergence will result if the static


directional stability is strong when compared with the dihedral effect. If
an aircraft that has strong directional stability has its right wing down, a
positive sideslip angle is produced. Because of strong directional sta-
bility, the aircraft tries to correct directionally before the dihedral ef-
fect can correct laterally. The aircraft chases the relative wind, and the
resulting flight path is a descending spiral. To correct this condition, the
aviator needs only to raise the wing with the lateral control surfaces and
the spiral stops immediately.

(3) Dutch roll.


(a) A
dutch roll results from a compromise between directional
A dutch
and spiral divergence. roll occurs somewhere between the two. In
this case, the lateral stability of the aircraft is stronger than the direc-
tional stability. The directional tendencies of the aircraft have been re-
duced from the condition that led to spiral divergence.

(b) If the aircraft has the right wing down, the positive
sideslip angle corrects the wing position laterally before the nose of the
aircraft tries to line up with the relative wind. As the wing corrects, a
lateral directional oscillation starts. As a result, the nose makes a
figure-eight pattern on the horizon. The rolling and yawing oscillation
frequencies are the same, but they are out of phase.

SECTION
II
ROTARY-WING STABILITY

3-8. CONSIDERATIONS

Before directly addressing the stability characteristics of a helicopter,


certain properties of rotary-wing aircraft must be understood. These
properties are discussed in the following paragraphs.

a. Developing Vertical and Horizontal Components of Thrust. In this


discussion on stability, total lift of the rotor system is referred to as
thrust. Thrust of the rotor system always acts perpendicular to the
3-22
tip-path plane. When the tip-path plane is parallel to the horizon, thrust
is oriented straight up and no force is created that causes the helicopter to
move directionally. However, when the tip-path plane is tilted in reference
to the horizon, thrust is also tilted. A component of this thrust supports
the weight of the helicopter, while another component of thrust acts in the
direction of rotor tilt. The component that supports the weight is referred
to as the vertical component of thrust (TV). The component of thrust that
acts in the direction of rotor tilt is the horizontal component of thrust
(TH). In equilibrium flight, TV is equal to weight and TH is equal to para-
site drag. Figure 3-22 shows the development of these two components.

T T~

Th~

Figure 3-22. Tilted tip-path plane

b. Developing Control Moments. force times an arm. To


A moment is a

change the fuselage attitude, force must act on the CG of the helicopter.
a
CG of the
A
force acting at some distance (arm) creates a moment about the
helicopter. This moment causes the helicopter to change its attitude about
its lateral (pitch), longitudinal (roll), vertical (yaw) axes. These
and
moments are created by the main and tail rotors. Different types of rotor
systems develop these moments in different ways.

(1) Developing a moment in a semirigid rotor helicopter.

(a) To develop a moment, the semirigid rotor must first develop


thrust. This means that if a control force or moment is needed to change
direction or fuselage attitude, the rotor system must first develop thrust.
Therefore, the control moment becomes a by-product of thrust.

3-23
(b) During hovering flight, the thrust vector passes through
the CG of the helicopter is produced. If the pilot desires to
and no moment
enter forward flight, he moves the cyclic forward. The thrust vector no
longer passes through the CG. Therefore, a moment arm is created with re-
spect to the CG, and a moment is placed on the fuselage in the nose-down
direction. This moment tilts the fuselage until the thrust vector again
passes through the CG. Figure 3-23 illustrates the development of this
pitching moment.

MOMENT ARM

Figure 3-23. Development of a


pitching moment

(2 ) Developing a moment in a
fully articulated rotor system.

(a) A helicopter with a fully articulated rotor system can


develop control moments on the fuselage without developing thrust. While the
semirigid rotor is hinged at the center of the mast, the blades of a fully
articulated rotor system are attached to the rotor hub at some distance from
the mast. This offset hinging allows centrifugal forces to produce control
moments.

With the rotor system turning at normal operating RPM, 10


(b)
to 40 tons of centrifugal force are trying to pull each blade away from the
axis of rotation. Anytime the tip-path plane is not parallel to the hub,
these strong centrifugal forces that are developed by the blades very quickly
align the hub with the tip-path plane. Since the hub is attached solidly to
the mast, any movement of the tip-path plane immediately transmits a pitching
moment to the fuselage. Centrifugal force is independent of thrust; there-
fore, the fully articulated rotor system can produce control moments even
though it is not producing thrust. Development of thrust increases the mag-
nitude of the control moments. Figure 3-24 illustrates the pitching moment
created by centrifugal force.

3-24
HINGE OFFSET
(FROM MAST TO HINGE)

CENTRIFUGAL FORCE
NOTE: When the tip-path plane changes,
the CF acting on the hinge offset quickly
aligns the hub and fuselage with the tip-path
plane.

Figure 3-24. Pitching moment created by centrifugal force

3-9. SPEED STABILITY

a. When a helicopter hovers under zero wind conditions, the velocities


on the advancing and retreating blade sides of the rotor are equal. If a
20-knot wind suddenly develops, the helicopter reacts noticeably.

The relative velocity on the rotor blades moving forward into the
b.
wind will
be greater than the relative velocity on the blades moving with
the wind. Therefore, more lift
exists on the advancing blade side of the
rotor than on the retreating blade side. This additional causes an lift
up-motion that is 90 degrees displaced from the position of the force, and
the rotor will The aft tilting of the tip-path plane causes the thrust
tilt.
vector to and the rotor attempts to take up a velocity equal to that of
tilt,
the wind. In other words, the helicopter attempts to stabilize at zero
airspeed.

c. The helicopter will also attempt to stabilize at a given airspeed if


it is stabilized in forward flight and receives a small force that attempts
to increase forward flight velocity. The velocity increase causes additional
flapping, which tends to decrease the airspeed back to the original airspeed.
This type of reaction is called speed stability. Figure 3-25 shows the
reaction' of the rotor to sudden wind.

3-25
T
T1

...
v

þ-

Figure 3-25. Change in rotor attitude because of wind gusts

3-10. ROTOR ANGLE OF ATTACK STABILITY

When helicopter is in forward flight and a moment is imparted to the


a

fuselage so that the fuselage changes attitude in the nose-up direction,


changes take place in the thrust and moment that act on the entire aircraft.

A pitch to a more nose-up position causes a moment on the helicopter


a.
fuselage and an increase in thrust. As shown in Figure 3-26, as the aircraft
pitches to a more nose-up position, a cyclic åction on the rotor blades
causes an increase in angle of attack on the advancing blade. This always
leads the pitch of the fuselage. As the tip-path plane leads the fuselage in
pitch, a moment develops on the fuselage. This moment tends to pitch the
fuselage to even higher angles of attack. As the tip-path plane tilts to a
more horizontal position, the magnitude of rotor thrust increases. This
causes a
further increase in the moment on the fuselage, which causes a more
rapid pitch increase.

3-26
ðT

T
T

Figure 3-26. Change in thrust because of change in rotor attitude

b. The nose-up pitch of the fuselage is more unstable than the nose-down
moment because of the increase in thrust and the offset of the thrust axis
from the CG. The offset of the thrust axis is due to the pitch of the tip-
path plane ahead of the fuselage. Figure 3-27 shows the moment variation
with fuselage angle of attack.

FUSELAGE NOSE UP
M

Figure 3-27. Helicopter fuselage pitching moments

3-27
c. Figure 3-27 also shows the reaction of the helicopter to fuselage
pitching moments. A
nose-up moment causes a further nose-up moment; a nose-
down moment causes a further nose-down moment. A nose-down condition is not

as unstable as a
nose-up condition because a nose-down condition reduces the
magnitude of the thrust. Therefore, a
further increase in the nose-down
moment is not as large. Although helicopters are unstable with angle of
attack, they are more unstable during a nose-up condition.

3-11. STATIC STABILITY DURING HOVERING FLIGHT

a. previously discussed, static stability is the tendency of an


As
aircraft to return to the trimmed, or equilibrium, position. If a helicopter
is angularly displaced while it is hovering, it is assumed that only the
fuselage is displaced. However, because of the cyclic movement of the blades
that takes place as a result of fuselage rotation, aerodynamic dissymmetry of
The
lift quickly realigns the tip-path plane of the Asrotor with the fuselage. does
thrust axis continues to pass through the CG. long as the helicopter
not develop velocity, it will not experience a restoring moment.

b. If violation of the hovering flight


angular velocity develops, a

restriction this particular condition would exist. Therefore,


imposed on
the helicopter in hovering flight possesses neutral static stability
(Figure 3-28).

Figure 3-28. Neutral static stability during hovering flight

3-12. DYNAMIC STABILITY DURING HOVERING FLIGHT

a. The motion of hovering helicopter and the motion of a fixed-wing


a

aircraft following a disturbance are similar. If the hovering helicopter is


displaced in a roll to the right, the resultant force to the right causes it

3-28
to move to the right. The helicopter then develops
a
velocity relative to
the rotor (Figure 3-29).

T
T

..

v ..

Figure 3-29. Dynamic stability during hovering flight


(1) As the helicopter moves to the right, differential lift develops
on the rotor disk. Because of stability with speed, the tip-path plane
on the fuselage because of the dis-
tilts. This causes a moment to develop
the CG. Dynamic stability in hovering
placement of the thrust axis from
flight is shown in Figure 3-29. The helicopter then moves back with enough
stability and speed to cause it to return to its original position.
(2) As shown in Figure 3-29, when the helicopter moves to the right,
damping mayor may not be present. If damping is present, it tends to delay
the displacement of the rotor cone so that the fuselage catches up to the
thrust axis sooner (Figure 3-30). With damping, the magnitude of the restor-
ing moment is less. Therefore, the restoring time will be somewhat longer.
An increase in speed stability decreases the period of the oscillation, while
an increase in roll damping increases the period of the oscillation.

3-29
T

WITHOUT DAMPING

WITH DAMPING

Figure 3-30. Damping

b. During hovering flight, a single-rotor helicopter is slightly


unstable dynamically. By decreasing the moment of inertia of the helicopter
fuselage and increasing the moment of inertia of the rotor blades about their
flapping hinges, dynamic instability can be reduced. To increase the moment
of inertia of the rotor blades, the vertical height of the rotor above the CG
of the helicopter is increased and the flapping hinges are offset from the
center, or hub, of the rotor.

3-13. STATIC STABILITY DURING FORWARD FLIGHT

a. Static stability of the helicopter in forward flight depends on the


moments produced on the fuselage. These moments can result from a change in
speed from the trim airspeed at a constant angle of attack. The change to
the moment can also result from a change from trim at a constant airspeed.
The aviator must remember that the rotor is stable with respect to speed and
unstable with respect to angle of attack changes on the rotor. In addition
to the rotor, the fuselage and stabilizing surfaces, such as the synchronized
elevator or stabilator, may also contribute to stability in forward flight.

(1) The conventional helicopter fuselage has an unstable variation


of moment with angle of attack which adds instability to the rotor angle of
attack. Therefore, if the fuselage goes to a higher angle of attack during
forward flight, it tends to go to an even higher angle of attack as a result
of the initial pitch-up. A fixed or a movable tail surface can contribute a
stabilizing variation of moment with angle of attack.
(2) During forward flight, conventional helicopters have a nose-down
moment that acts on the fuselage. This happens because the center of drag
that acts on the fuselage is below the center of thrust produced by the
rotor. If the speed of the helicopter increases from the trim condition at a
constant angle of attack, the variation in moment is destabilizing. An
increase in speed causes a further nose-down moment, which tends to cause
further increases in speed. A stabilizing surface that carries a down load

3-30
and contributes nose-up moment in steady trimmed flight will
a compensate for
in speed
this variation in moment. Under these conditions, an increase
stable speed
causes a nose-up moment and the helicopter will
maintain a

(Figure 3-31).

(3) To compensate for the moments present on the fuselage and froma
the tail, the thrust axis is offset from the
CG during steady flight. If
down load in forward flight, the
horizontal tail is present and carryingCG
a

to balance the tail moment. This


thrust axis must be offset aft of the and possibly make
offset can balance rotor instability with angle of attack
3-31 and 3-32 show these
the helicopter stable ~ith angle of attack. Figures
principles.

~v

INCREASE IN DOWN LOAD

Figure 3-31. Speed stability

3-31
T

INCREASE IN FUSELAGE ATTITUDE DECREASE IN DOWN LOAD

INCREASE IN MOMENT ARM

Figure 3-32. Angle of attack stability

3-32
(4) Normally, the helicopter is stable with speed during forward
flight. This implies that an increase in velocity during forward flight will
cause an up moment on the rotor and ultimately on the fuselage. This will
create a
nose-up moment and decrease velocity (Figure 3-33).

(5) If the helicopter is trimmed in level, steady flight" at a


particular fuselage angle of attack and velocity and receives a momentary
a

velocity increase at a
constant angle of attack, it moves to positive mo-
ment, which tends to decrease its velocity. Figure 3-33 shows stability with
speed during forward flight. Unlike fixed-wing aircraft, rotary-wing air-
craft require a
series of curves to show stability because of the moments
produced by a change in speed from the trim point.

CM 0

Figure 3-33. Pitching moments because of speed

The angle of attack reaction of the helicopter during forward


(6)
flight shows one of two types of stability, depending on the amount of down
load being carried by the tail and whether the helicopter has a
horizontal
As mentioned previously, if the down load is sufficient, the helicop-
tail.
ter exhibit stable characteristics with respect to angle of attack during
may
forward flight. Figure 3-34 shows slightly unstable characteristics with
angle of attack.

3-33
+

CM 0

Figure 3-34. Angle of attack stability with speed

b. If the helicopter is stabilized in steady, level forward flight, an


increase in the fuselage angle of attack causes a positive moment about the
CG because of the
effect of the fuselage moment. If an aerodynamically
strong horizontal tail were added to the aircraft and a nose-up moment im-
parted to the fuselage, a stabilizing moment would be created. Figure 3-34
shows instability with speed in forward flight.

3-14. DYNAMIC STABILITY DURING FORWARD FLIGHT

The dynamic stability of the helicopter during forward flight can be


described in terms that have previously been discussed. If the helicopter
has neutral static stability with respect to changes in angle of attack for
hovering flight, the period of the longitudinal oscillation during forward
flight is influenced by the same factors as hovering flight conditions.
These factors are static stability with speed damping in pitch and insta-
bility with angle of attack.
a. Dynamic oscillations during forward flight are described in this
paragraph. A helicopter in trimmed, level flight receives a disturbance that
causes it to pitch nose down. The component of weight along the flight path
causes the helicopter to increase velocity. Stability with speed and damping
with pitch will influence the resulting oscillation as the helicopter at-
tempts to return to straight and level flight under equilibrium conditions.

b. An increase in the amount of speed stability will cause a larger


nose-up moment. A larger nose-up moment will cause larger nose-up angular
velocities and will reduce the period of the oscillation. Increasing sta-
bility with speed reduces the period of the oscillation.

3-34
c.
A
large amount of damping in pitch causes a reduction in the angular
velocity of pitch but increases the period of the oscillation. Angular ve-
locity in pitch is caused by instability with the angle of attack, which is
inherent in all helicopters. Damping in pitch and instability with angle of
attack oppose each other. If a horizontal tail surface is added to the heli-
copter to decrease the instability with angle of attack, the period of the
oscillation increases further because damping in pitch has more effect.

d. The dynamic instability of


a
helicopter during forward flight can be
reduced by increasing the damping in pitch or reducing the instability with
the angle of attack. Instability with the angle of attack can be reduced by
adding a stabilizing device or by sacrificing speed stability.

3-35
CHAPTER 4

AIRCRAFT DESIGN, STRUCTURE, AND ENGINE OPERATION

metals and explains


This chapter discusses the characteristics of aircraft
how a flight envelope is developed to predict operating
limitations on an
of the gas turbine
airframe. It also discusses the operating characteristics
and reciprocating engines used for aircraft propulsion.

SECTION I

DESIGN AND STRUCTURE

4-1. DESIGN CONSIDERATIONS

designed with respect to


Aircraft structures and components are efficiently
performance deteriorates
structural strength versus weight. Because aircraft
as weight increases, aircraft are built of lightweight materials. Structural
and the airframe is constructed to withstand
strength is closely controlled, both the
understand
the loads imposed by normal operations. Aviators must
structural and the aerodynamic limitations on their aircraft.
designed for loads exceeding normal
a. Because the airframe is not to fail. An
requirements, overstress can actually cause the structure loads or exces-
aircraft can be overstressed through excessive acceleration
exceed the structural
sive airspeeds. Aviators must never intentionally discussed
limitations imposed on the aircraft. Structural limitations are
further in paragraph 4-3.
One limitation is set by
b. Aircraft also have aerodynamic limitations.which depends on the maxi-
the maximum lift obtainable at certain airspeeds,
mum value of the CL obtainable (CLmaximum)' In addition, aeroelastic effects
aerodynamic forces acting on the
result from the combined effect of all discussed further in
airframe structure. Aerodynamic limitations are
paragraph 4-3.

4-2. STATIC STRENGTH OF METALS

know the static strength


To design components, manufacturers need to
aircraft
of metals. They can then determine the type of metal to use, how much load
needed to support a specific
the metal can withstand, and how much metal is
are stress, strain,
load. Some of the terms used to explain static strength
and metal fatigue.

a. Stress and Strain.


by giving an
(1) Metal stress and strain can be explained best
10 inches long with a
example. Figure 4-1 shows a metal sample

4-1
cross-sectional area of 1 square inch. If a pulling force, or tension, of
40,000 pounds is applied to the sample, the metal is then under 40,000 pounds
of stress per square inch. Stress, the force per unit area, is found by
dividing the total force by the applied cross-sectional force. If the cross-
sectional area is 2 square inches, then the stress is only 20,000 pounds per
square inch. The stress can be reduced either by decreasing the force or by
increasing the area.

/
F =
40,000 LB

11"t
K,
~ "
10"
/1
~<::J<::J"~
~.-..
o.<::J'
V
Figure 4-1. Metal stress and strain
(2) Thestrain is the deformation of the sample by unit (in inches
per inch). If the sample stretches a total length of 0.1 inch, the strain
would be 0.01 inch per inch. This means that each of the 10 inches of the
original sample stretched 0.01 inch, making the total elongation of the
lO-inch sample 0.1 inch.

(3) "Jhen a sample is given a tension test to determine


its static
strength, the stress applied is plotted against the strain. In designing a
wing main spar, for example, aircraft manufacturers need to know what type of
metal to use and how much load it can withstand. They also want to know how
much metal is needed to support.a specific load and how much the wing
spar
bends when subjected to the load.

(4) As tension stress is applied to metal, the metal starts to


stretch. The strain remains proportional to the stress applied; if the
stress is doubled, the strain is doubled. This proportionality continues to
a
stress equal to a point called the proportional, or elastic, limit. If the
material is not stressed beyond the elastic limit, the material returns to
its original shape and size when the stress is removed. An example is a bar
that is bent slightly and then released; it returns to its original state.
The metal in the bar has not been stressed past its elastic limit, so
it
returns to its original shape and size with no permanent deformation. If the
bar is bent again and a stress greater than the elastic limit is applied,
minute portions of the metal fail. When stress is removed this time, the bar
does not return to its original shape. The bar is now permanently deformed

4-2
because of overstress. These minute portions of the metal are still elastic
but the bar is permanently bent.

(5)Metal can be stressed somewhat beyond the elastic limit without


The maximum stress that can be applied is
failing completely or breaking.
called the ultimate stress.
A
stress greater than the ultimate stress will
metals
break the sample. When designing an aircraft, manufacturers select
that will not be stressed beyond the elastic limit during normal operations;
they use a design limit of about 0.3 of the elastic limit.

b. Fatigue.

(1) Metal fails not only from overstress but also from repeated
This is called metal
applications of loads well within the elastic limit. The lid
by using a
tin can as an example.
fatigue, which can be visualized way around its
is removed by cutting the can about three-quarters of the
forth until it finally breaks. If
perimeter. Then the lid is bent back and
more cycles
the lid is bent only slightly in each direction, it takes several
it will The
for the metal to fail than if it is bent as far as go.
the higher
manufacturer must consider this characteristic of metal fatigue;
the stress applied to a
part, the shorter its service life.
(2) Stress can sometimes be repeated over and over again, and the
this is
materials will withstand the millions of cycles without failure; causes
This type of stress never
referred to as the metal's endurance limit. a
For example,
the metal to fail regardless of how many times it is applied.
would never fail. When
wing spar designed using the endurance limit criteria
a definite load is to be applied, stress is reduced by increasing the cross-
added metal adds weight to the
sectional area of the part. However, the
as the design
aircraft. For this reason, the endurance limit is not
used

limit.
how
(3) obtain the design stress limit, manufacturers calculate
To
the design loads are to
long aircraft are to be in service and how many times
be applied. If a safety factor has not been included ain safetycalculations,
the
factor. Manu-
the maximum design stress limit is lowered to include
facturers use this information to build airframes that fulfill the
the operating
requirements outlined in aircraft specifications. Aviators see results
recommended
of these structural limitations in several forms; for example,
and G-force
landing techniques and power settings and imposed speed
limitations.

4-3. FLIGHT ENVELOPE

the specifications
Manufacturers guarantee aircraft will perform according to
set forth by the certifying agencies. They also guarantee aircraft To
will have
achieve
a
service life equal to the requirements in those specifications. a prescribed
this service life, the aviator must operate each aircraft within
flight envelope (operating limitations).

~3
a. Structural Limitations.
(1) When an aircraft is undergoing positive G-loads, a bending
moment is created at the junction of the wing and the fuselage, as shown in
Figure 4-2. Because of the design of the aircraft, only a certain lòad
factor can be applied without damage. For example, a 6-G aircraft has a
design limit load of 6 Gs at gross weight. The stresses incurred in a 6-G
maneuver with this aircraft are not greater than the elastic stress on any
component of the airframe. When the load factor is removed, the aircraft
returns to its original shape and the airframe is not damaged. The negative
load limit is usually less than the positive load limit. Because the
aviator's body cannot withstand excessive G-forces, there is no reason to
build the airframe to support a load that it will never receive. This
weight-saving factor increases performance.

(2) Any negative or positive G-force greater than the limit load
(load factor times weight) may cause damage, thereby decreasing the service
life of the airframe. The damage may not be visible and might be detected
only in an internal inspection. Therefore, when an aircraft is flown past
its design limit load, maintenance personnel perform an inspection.
(3) The
ultimate limit is usually about 1.5 times the design limit
load. If the aircraft is subjected to G-loads above this limit, the airframe
structure will probably fail.
(4) The structural limitations that the manufacturer imposes on the
airframe should not be exceeded intentionally. If an aircraft is operating
at high G-loads, gust loads may sometimes push an aircraft past the limit
load. However, aviators can avoid this by decreasing the airspeed when tur-
bulent air is encountered. Most operator's manuals recommend a penetration
speed for severe turbulence.

% \\t\L
~"~~/
~- ,I~
14-~ ~

'-.--/
~
-------
-

nW~

Figure 4-2. Bending moment

4-4
b. Aerodynamic Limitations.

(1) Redline speed denotes the maximum design speed of the aircraft,
which is usually marked with a redline on the airspeed indicator. Above this
speed, high dynamic pressure and aeroelastic effects overstress the airframe
structure. The red line speed should never be exceeded because the airframe
will probably be damaged.

(2) Aeroelastic effects are the interactions between aerodynamic,


dynamic, and elastic forces of the structure. One form of these effects is
wing divergence, which causes the wing structure to fail immediately. If an
aircraft is flown at velocities above the redline, a change in the lift force
can produce a
positive twisting moment on the wing because of the high
dynamic pressure striking the leading edge. The twist increases the angle of
attack of the affected wing sections, which then produces an additional lift
force that twists the wing even more. This twisting continues until the wing
.

structure fails.
(3) Wing occurs at excessive airspeeds. Because of
flutter often
the elasticity of the wing, the normal frequency of the wing occurs at
velocities above the designed redline speed. Wing flutter does not occur as
quickly as wing divergence. Therefore, the aviator should have enough time
to reduce the airspeed and thus stop the flutter.

(4) When the airspeed of a compressibility-limited aircraft exceeds


its redline speed, the aviator must be concerned with the effect on compress-a
ibility. A high airspeed can create a shock wave on the wings which causes
buffet, or flutter, that overstresses the airframe. This flutter feels
similar to the flutter encountered before an aerodynamic stall.

Figure 4-3 is a chart showing the flight envelope (or Vg) for a
(5)
T-42A aircraft. The information is typical of all Army fixed-wing a1rcraft.
The velocities shown are indicated airspeeds, so conversion to true airspeed
is unnecessary. The G-limits shown are applicable at gross weight.

4-5
(225 KIAS)
(NEVER EXCEED SPEED)
12
I I I I I
11
NORMAL OPERATING CONDITIONS
- -

c:J
10 OPERATION WITH FLAPS DOWN
)-
-

c:J
<;' CAUTION RANGE
91- -
c:J
~ PROHIBITED OPERATION
c:J
0
I-
8 ........
-

.-
~ 7
/'
~--
u. I
~ 6
0
5
DESIGN CRUISING SPEED (Vd
,STRUCTURAL
~ \
-I
4
NORMAL STALL SPEED 44 '\..
.
í',~,~~~~~~~,~A. r--~ ~
(FLAPS DOWN)
m
...I
0 3
\ I05~~~1
p..~~
'
::I
-I

/
0
LEVEL FLIGHT \ G~\.~"'- I
'\. u. \
:ri- 2 -

AT 1G,
\ , ~ i )
-I
c(
~
\
u.
0
I-
~
1

0
-'
-j )(.
I
I
::I
I-
Ü
::I
J
~
W
3= 1
-
-I--..:!.ELOCITY 0
FLAPS EXTENDED I
I I-
en
I
-

(Vte)
en ~ I \
~ -2 I

NORMAL STALL SPEED


'I'~ J tr-
a: -.......... """"'- I
CI
z
-3 (FLAPS UP) 3.0
~ ~STRUCTURAL
DAMAGE AREA
0
-4
0
~ -S
I
I ~//I W//~
4.S

c(
m
MANEUVERING SPEED (Va) ~ ,, -'
-6 I
I
-7 I
I
-8 I I I I I I I I I I I I I I
0 10 20 30 40 SO 60 70 80 90 100 110120130 140 1S0 160 170 180 190200 210 220230240 2S0 260
(KIAS)

Figure 4-3. Flight envelope chart (Vg diagram)

SECTION
II
PROPULSION

POWE R SOURCE S
4-4.

a. Early attempts at flight showed that heavier-than-air aircraft needed


power to both attain and sustain flight. It was also obvious that the human
body was inadequate as a source of propulsion.

For heavier-than-air aircraft to get off the ground in a controlled


b.
powered flight, an adequate power source had to be developed. In the search

4-6
for a
satisfactory engine, rocket, steam, jet, and reciprocating engines were
tried. It was a reciprocating engine that first pushed the Wright brothers
al of to

c. Modern are propelled by propulsion systems. This manual


aircraft
limits discussion to only those power plants used on Army aircraft--gas
turbine and reciprocating engines. Gas turbine engines are discussed in
Section III
and reciprocating engines, in Section IV. To complete power
production, each type of engine uses an aerodynamic propeller to convert
shaft horsepower to thrust.

4-5. PROPULSION PRINCIPLES

The principal items of flight performance involve steady-state flight


conditions and aircraft equilibrium. For an aircraft to remain in steady and
level flight, a balance of the forces acting on the aircraft must be
obtained. The lift must equal the weight of the aircraft, and the power
plant thrust must equal the total drag of the aircraft. Because drag
increases or decreases with velocity, a power plant must produce sufficient
thrust to overcome total drag. Without this capability, an aircraft is
unable to accelerate for takeoffs or climbs. At maximum thrust, an aircraft
generally reaches the maximum airspeed it can achieve in level or climbing
flight.

SECTION
III
GAS TURBINE (TURBOJET) ENGINES

4-6. OPERATING CHARACTERISTICS

The gas turbine engine is widely used in aircraft propulsion because of its
relatively high power output for its power plant weight and size. Few air-
craft power plants can compare with the high output, flexibility, simplicity,
and small size of an aircraft gas turbine. The continuous steady-flow
feature of the gas turbine allows it to process greater airflow and to use
fuel energy more efficiently. If the entire gas turbine engine is examined
with respect to airflow through it, the components would align as follows:
0
Air inlet, or diffuser section.
0
Compressor section.

0
Combustion chamber.

0
Turbine section.
0
Exhaust nozzle.

4-7
A schematic of a typical turbine engine is shown in Figure 4-4. The
operating characteristics of the eng1ne components are discussed below.

GAS
GENERATOR

FREE-STREAM
AIRFLOW ..-

~ ,
EXHAUST
NOZZLE

Figure 4-4. Turbine engine components

a. Air Inlet, or Diffuser. Air entering the engine passes through the
inlet, or diffuser, to reach the compressor. The inlet 'is designed to
deliver a smooth flow of air with even pressure distribution to the face of
the compressor. The inlet takes air of varying velocities and then delivers
it to the first stage of the compressor at the lowest possible velocity and
highest possible pressure.

b. Compressor Section. The compressor section furnishes the combustion


chamber with large quantities of high-pressure air. A
centrifugal compres-
sor, an axial-flow compressor, or a combination of the two types of
compressors can be used for this function.

(1) Centrifugal compressor. This type of compressor is more


desirable than the axial~flow compressor because of its simplicity and
limi ted susceptibili ty to stall or surge. The impeller rotating at high
speed receives the inlet air and provides high acceleration from the
centritugal force. As a result, the air leaves the impeller at a very high
velocity and with high kinetic energy. When kinetic energy is converted to
static pressure energy, the subsequent expansion in the diffuser manifold
produces a pressure rise. The manifold then distributes the high-pressure
A
discharge to the combustion chamber. double-entry impeller allows a given
diameter compressor to process a
greater airflow. However, the operation of
the centrifugal compressor requires relatively low inlet velocities; a plenum
chamber or expansion space must be provided for the inlet. t1ajor components
of the centrifugal compressor are shown in Figure 4-5. The centrifugal
compressor provides pressure ratios of 9:1 per stage; however, the use of
more than one or two stages is usually not feasible for aircraft turbine
engines. The single-stage centrifugal compressor is capable of producing
pressure ratios of about 3 or 4 with reasonable efficiency. Pressure ratios

4-8
greater than 4
require high impeller tip speed that compressor effi-
such a

ciency decreases rapidly. High-pressure ratios are necessary for low fuel
consumption. For that reason, the centrifugal compressor is best suited to
small engines where the principal requirements are simplicity and flexibility
of operation rather than high efficiency.

þ- CENTRIFUGAL COMPRESSOR
INLET

~ DISCHARGE

DIFFUSER AND MANIFOLD

Figure 4-5. Centrifugal compressor components

(2) Axial-flow compressor. This type of compressor is capable of


providing the high pressure ratios necessary for low fuel consumption. The
axial-flow compressor provides high airflow with a smaller compressor
diameter than the centrifugal compressor. Compared to the centrifugal
compressor, the axial-flow compressor is complex and costly in design and
much narrower range
construction. Also, high efficiency is sustained over
a

of operating conditions. The axial-flow compressor is best suited for


engines where efficiency and output are more important than cost, simplicity,
and flexibility of operations. Major components of the axial-flow compressor
are shown in Figure 4-6. The axial-flow compressor consists of alternate
rows of rotating airfoils (rotor blades) and stationary airfoils (stator
blades). A pressure rise occurs through the row of rotating blades, because
the airfoils cause a decrease in velocity relative to the blades. An addi-
tional pressure rise occurs through the row of stationary blades, because
these airfoils cause a decrease in the absolute velocity of flow. The
decrease in relative or absolute velocity causes a compression of the flow
and results in an increase in static pressure. While the pressure rise per
stage of the axial compressor is relatively low, the efficiency is high.
High pressure ratios can be obtained efficiently by successive axial stages.
Pressure ratios from 5 to 10 (or greater) can be obtained with the
multistage, axial-flow compressor.

4-9
STATOR BLADES COMPRESSOR CASE

DISCHARGE

AIR INLET
.. COMPRESSOR
ROTOR BLADES

AIR INLET ~ SHAFT

DISCHARGE

Figure 4-6. Axi al-fl ow compr essor components

(3) Axial-centrifugal compressor. This compressor, shown in Figure


4-7, combines the advantages of both types of compressors by giving moderate
pressure ratios and a means of minimizing engine length. It also provides
the means to cope with the increased drag effects of small blades i~ small
engines. Because the diameter of the turbine engine has decreased, the chord
of the axial-compressor blades has also decreased. As the stages in the
small engine are increased, the chord of the blade becomes even smaller. The
smaller blade chord reduces the maximum lift
coefficient that the individual
blades can develop. For this reason, it is better to increase the pressure
ratio of small turbine engines by adding a centrifugal compressor rather than
more axial stages.

INLET
. SHAFT

Figure 4-7. Axial-centrifugal compressor

4-10
c. Combustion Chamber. The chemical energy of fuel is converted into
heat energy in the combustion chamber, causing a large increase in the total
energy of the engine airflow. The combustion chamber has one principal
limitation--the discharge from the combustion chamber must be at temperatures
that the turbine section can tolerate. Combustion of turbine fuels can
produce gas temperatures that exceed 2,000 degrees Celsius. However, the
maximum continuous operating temperatures of the turbine blade must be kept
between 800 and 2,000 degrees Celsius. About 75 percent of the total airflow
is used for cooling, and the remaining 25 percent is used for combustion.

(1) The design of a combustion chamber may take various forms; the
main features of a typical chamber are illustrated in Figure 4-8. High-
pressure air is discharged from the compressor into the combustion chamber.
About one-fourth of this air is introduced into the immediate area of the
fuel spray. This primary combustion air has relatively high turbulence and
low velocity to maintain a nucleus of combustion in the chamber. During
normal combustion, the speed of flame propagation is quite low. Poor
combustion results when local velocities are too high at the forward end of
the chamber. High velocities of airflow may cause the flame to blowout.

(2) The remaining air, which bypassed the area where fuel was
injected, is called secondary, or cooling, air. This air is reintroduced
downstream from the combustion nucleus. It dilutes the combustion products
and lowers the discharge gas temperature so the turbine blades do not over-
heat. The mass flow of this cool air produces most of the thrust in a
turbojet or turbofan engine.

SECONDARY AIR
OR COOLING FLOW

~~ -...
PRIMARY
COMBUSTION
AIR
.
c . DISCHARGE
TO TURBINE
NOZZLES

FUEL
SPRAY
NOZZLE

COMBUSTION NUCLEUS

Figure 4-8. Typical combustion chamber

4-11
d. Turbine Section. The turbine section, sometimes called the hot
section, is the most critical element of the gas turbine engine. Power to
drive the compressor, its accessories, and the aircraft propeller or rotor
systems is extracted from the combustion gases as they pass through the
turbine section.
(1) The combustion chamber delivers high-energy combustion gases to
the turbine section at a high pressure and a tolerable temperatur~. Turbine
nozzle vanes consist of a stationary row of blades located immediately behind
the combustion chamber, as shown in Figure 4-9. These blades form the nozzle
that discharges the combustion gases as high-velocity jets onto the rotating
turbine. The high-pressure energy of the combustion gases is converted to
kinetic energy, and the pressur~ and temperature drop. The turbine blades
operating downstream of the jets develop a rotating force along the turbine
wheel, extracting mechanical energy from the combustion gases.

(2) There are two types of turbines in the turbinesection--the gas-


generator turbine and the power turbine. These turbines are usually mounted
on separate shafts and are only connected aerodynamically. The gas-generator
turbine is on the shaft with the compressor. This turbine extracts energy
from combustion to drive the compressor and any accessories that are driven
from that shaft. The power turbine is on a separate shaft that is connected
to the aircraft propeller through a speed-reduction gearbox or to
a

helicopter rotor system through the transmission. Because the power turbine
is located downstream from the gas-generator turbine, the gases that turn it
have been partially spent. This leads to a slower rotational speed for the
power turbine than for the gas-generator turbine. The speed of the
gas-generator turbine is called Nl speed, while the speed of the power
turbine is called N2 speed; both speeds may also be expressed in RPM.

TURBINE NOZZLE
VANES

~
CHAMB~ER
COMBUSTION

'- :-
.- :-
-
i
-
-

-
-

- -
-

-~
-

-
-

~
"~

GAS-G'N~RATOR
COMPRESSOR
:7 ~ POWER TURBINE
TURBINE

Figure 4-9. Typical turbine section

4-12
the engine through the exhaust
e. Exhaust Nozzle. Exhaust gases leave
the hot exhaust gases
nozzle. Some exhaust nozzles are curved to deflect This
away from aircraft surfaces or into propeller or rotor slipstreams.
helps to reduce the infrared signature of the aircraft.

GAS TURBINE-PROPELLER COMBI~ATION


4-7.
The turbojet engine uses the turbine to extract sufficient power for
used to provide
compressor operation. The remaining exhaust gas energy is
the high exhaust gas velocity and jet thrust. The propulsive efficiency of
by creating
the turbojet engine is relatively low because thrust is produced
a
large velocity change with
a
relatively small mass flow. The gas turbine-
propeller combination. shown in Figure 4-10. has a higher propulsive
on a much
efficiency in subsonic flight because the propeller operates
greater mass flow. -----
a. The turboprop power plant requires
additional turbine stages to
continue expansion in the turbine section and to extract a large percent of
is
the exhaust gas energy as shaft power. In this sense. the a turboprop
is small proportion
primarily a power-producing machine and the jet thrust
of the turboprop
of the output propulsive power. Ordinarily. the jet thrust
Since the turbo-
accounts for 15 to 25 percent of the total thrust output.
prop is primarily a power-producing machine. the turboprop power plant
is
rated by an equivalent shaft horsepower.

b.The gas turbine engine must operate at very high rotative speed to
a

process large airflows and produce high power. However, high rotative speeds
of compressibility
are not conducive to high propeller efficiency because
be greatly reduced to match the power plant to the
effects. Shaft speed must
propeller. Propeller-reduction gearing provides
a shaft speed that the
such gearing cause
propeller can use effectively. However. the problems with
major difficulties in. developin.g turboprop power plants.

REDUCTION TURBINES
COMPRESSOR
I GEARING
c:::> c:::> c:::> c:::>
c:::> c:::> c:::>

c::::>

COMBUSTION EXHAUST
CHAMBER NOZZLE

Figure 4-10. Gas turbine-propeller combination

4-13
c. Turboprop power plants require a governing apparatus to account for
the variable propeller blade angle. If the propeller is governed separately
fro~ the turbine, an interaction can develop between the engine and propeller
governors; this May cause various "hunting," overspeed, and overtemperature
c ondi
tions. Fo r t his reason, the engi ne-propeller co~bination norma lly is
operated at a constant RPH throughout the major range of output power. The
principal variables of control are fuel flow and propeller blade angle. In
the major range of power output, the throttle commands a certain fuel flow
and the propeller blade angle adjusts to increase the propeller load to
remain at the governed speed.

d. Turboprop performance illustrates the typical advantages of the


propeller-engine combination. Higher propulsive efficiency and high thrust
at low speeds provide the range, endurance, and takeoff performance that are
characteristic of the turbojet. As is typical of all propeller-equipped
power plants, the power available is nearly constant with speed. The mass
flow and inlet temperature affect the equivalent shaft horsepower of the
turboprop. The ESHP varies with altitude because the higher altitude
produces a much lower density and engine mass flow. High air temperatures
cause a noticeable loss of output power because higher compressor inlet
temperatures reduce the fuel flow allowable within turbine temperature
limits. Typical values for specific fuel consumption range from 0.5 to 0.8
pounds per hour per ESHP. Hinimum specific fuel consumption is obtained at
relatively high power settings and high altitudes. High altitudes produce
lower inlet air temperatures, reducing specific fuel consumption. The lowest
values of specific fuel consumption are obtained near altitudes of 25,000 to
35,000 feet.

GAS TURBINE OPERATING LlMITATIONS


4-8.

The power plant must be operated within specified limitations to obtain


design service life with trouble-free operation. The following paragraphs
explain the critical areas encountered during the operation of the gas
turbine engine.

a. Fatigue and Creep Damage. The li~iting exhaust gas temperatures


provide the most important restrictions to the operation of the gas turbine
engine. Turbine components are subject to centrifugal loads of rotation and
impulse and reaction loads on the blades. Various vibratory loads may also
be inherent in design. When turbine components are subject to this variety
of stress along with high temperature, two types of structural phenomena may
occur. When a part is subject to a certain stress at some high temperature,
creep failure takes place after a period of time. An increase in temperature
or stress also increases the rate at which creep damage accumulates and
reduces the time required to cause failure. Another problem results when a
part is subjected to repeated, or cyclic, stress; fatigue damage causes
failure after a number of cycles. An increase in temperature or the
magnitude of cyclic stress also increases the rate of fatigue damage and
reduces the number of cycles necessary to produce that failure. Fatigue
damage and creep damage are both cumulative.

4-14
(1) A gross overstress or overtemperature of the turbine section
produces immediate damage. However, creep and fatigue damage accumulated
through periods of less extreme overstress or overtemperature is more subtle.
If the turbine is subject to repeated excessive temperatures, the greatly
increased rate of creep and fatigue dama~e causes failure early in the
anticipated service life of the turbine.

(2) Generally, the operations that produce the highest exhaust gas
temperatures are starting, acceleration, and maxiDum thrust at high altitude.
Time spent at these temperatures must be limited arbitrarily to prevent creep
and fatigue from accumulating excessively. Any time spent at temperatures
exceeding the operational lir.1Ï ts for turbine components increases the chance
of early failure.

(3) Turbine components are the most critically stressed elements at


high temperatures; however, the airframe structure and equipment adjacent to
the engine may also be subject to damage caused by excessive time at high
temperatures. In addition, the combustion-chamber components may be stressed
at low altitude where high cOI:lbustion-chamber pressures exist.
b. Compressor Stall or Surge. The compressor stall or surge may produce
damaging temperatures in the turbine and combustion chamber or unusual
transient loads in the compressor. While the stall-surge phenomenon may
occur with the centrifugal compressor, it
is more common with the axial-flow
compressor. To accelerate the engine to a greater speed, more fuel must be
added to increase turbine power above that required to operate the
compressor.

(1)
When the fuel flow is increased beyond the steady-state
require~nt without a change in rotative speed, the combustion-chamber
pressure increases and causes the compressor discharge pressure to increase.
The instant before a change in engine speed occurs, a decrease in compressor-
flow velocity accompanies an increase in compressor discharge pressure. The
component of velocity that is due to rotation remains unchanged for a given
rotative velocity of the single blade. The axial-flow velocity for the
steady-state operation combines with the rotational component to define a
resultant velocity and direction. If the axial-flow component is reduced,
the resultant velocity and direction increase the angle of attack for the
rotating blade, which then causes a subsequent increase in pressure rise. If
the change in angle of attack or pressure rise is beyond some critical value,
stall will occur. While the stall phenomenon of a series of rotating
compressor blades differs from that of a single airfoil section in a free
airstream, the cause and effect are essentially the same.

(2) If
an excessive pressure rise is required through the
compressor, stall may occur along with the breakdown of stable, steady flow
through the compressor. As stall occurs, the pressure rise drops and the
compressor does not furnish discharge at a pressure equal to the combustion-
chamber pressure. As a
result, a flow reversal or backfire takes place. An
intermittent "bang" will indicate the stall is transient. If the stall
develops and becomes steady, a strong vibration and a loud (possibly

4-15
explosive) roar develops from the continuous flow reversal. The increase in
the required compressor power tends to increase RPM. The reduced airflow and
increased fuel flow cause a rapid, immediate rise in the exhaust-gas
temperature. Because the steady stall may cause immediate damage, av~ators
must recover quickly by reducing the throttle setting, lO\<Jering the aircraft
angle of attack, and increasing the airspeed. Generally, the compresso
r

stall is caused by one or a combination of the follmJÍng conditions:


(a)malfunctioning fuel control or governing apparatus is a
A

common cause. Proper maintenance and adjustment are necessary for stall-free
operation. Usually, malfunctioning is mos apparen during engine
t t

acceleration.
Poor inlet conditions are typical at high angles of attack
(b)
and sideslip. They reduce inlet airflow and create nonuniform flow at the
compressor face. The aviator can control these conditions immediately by
reducing angle of attack or sideslip.

(c) Flight at very high altitudes produces high frictional


drag, which reduces the section maxÍl:lUm lif t coefficient. The effect is
similar to that of airfoil sections. These altitudes also reduce the maximum
pressure ratio of the compressor. The reduced stall margins increase the
likelihood of compressor stall. If high altitude is contributing to
a decrease in
compressor stall, recovery from a
compressor stall must entail
the power setting to reduce fuel flow and altitude.

c. Flameout. While flameout is


a
rare occurrence with modern engines,
various malfunctions and operating condi tions make ita possibili ty. A
uniform mixture of fuel and air sustains combustion within
a wide range of
fuel-air ratios--as rich as 1:5 or as lean as 1:25. Fuel-air ratiosof outside
these limits will not support combustion because of the deficiency air or
The characteristic of the fuel nozzle and the spray pattern and the
fuel.
of the governing apparatus mus t be such tha t combustion is
efficiency
maintained throughout t~e range of engine operation.

(1) the rich limit of fuel-air ratio is exceeded in the


If
combustion chamber, flameout will occur. While this condition is possible,
exceeding the lean limit is more often the cause of
a
flameout. Any
condition that produces a fuel-air ratio leaner than the lean limit of
the fuel
combustion will produce a flameout; for example, an interruption of
supply. Fuel system failure, fuel system icing, or prolonged unusual
attitudes could starve the flow of fuel to the engine. Most aviation fuels
are capable of holding a small amount of water in solution. If the aircraft
is refueled with relatively warm fuel and then flown to a high altitude, the
lower temperatures can precipitate this water out of solution in liquid or
ice crystal form.
(1) High-altitude flight produces a relatively small air-mass flow
through the engine and a relatively low fuel-flow rate. If
these conditions
apparatus to malfunction, flameout
cause the fuel control and the governing during
allows excessively low fuel flow
could occur. If the fuel control

4-16
controlled deceleration, the lean blowout limit may be exceeded. If the
governed idle condition allows any deceleration below the idle condition, the
engine usually continues to lose speed and flameout occurs.

(3) Restarting the engine in flight requires sufficient Rfl1 and


airflow to allow stabilized operation. Generally, the extremes of altitude
are critical for an attempted air start.

d. Temperature Increase. An increase in the compressor inlet


temperature can have a profound effect on the output thrust of a turbojet
engine. It produces an even greater increase in the compressor discharge
temperature. Since the turbine inlet temperature is limited by a maximum
value, an increase in compressor discharge temperature reduces the
temperature change that takes place in the combustion chaober. TI1erefore,
the fuel flow is limited and thrust is reduced.

e. Engine Overspeed. Critical vibration speed ranges and engine


overspeed affect the service life of an engine. One of the principal sources
of turbine loads is the centrifugal load resulting from rotation. Since
centrifugal loads vary as the square of the rotative s~eed, a 5 percent
overspeed would produce 10.25 percent overstress (1.05 1.1025). The large
=

increase in stress with rotative speed could cause creep and fatigue damage
to accumulate rapidly at a high temperature. Repeated overspeed and,
consequently, overstress can cause equipment to fail early in the anticipated
service life.
f. Vibratory Mode. The turbine engine has an elastic structure and is
composed of many different distributed masses. Therefore, certain vibratory
modes and frequencies for the shaft and blades exist. Resonant conditions
must be prevented within the normal operating range. However, certain
vibratory modes may be encountered in the 10017-power range common to ground
operation, low-altitude endurance, and acceleration or deceleration. If
operating RPM range restrictions are specified because of vibratory
conditions, operations must be conducted with a mininum of time in this area.
Greatly increased stress comn~n to vibratory conditions is likely to cause
fatigue failure of the offending components. The operating limitations of an
engine are usually specified by various combinations of RPM, exhaust gas
temperature, and allowable time.

SECTION IV

RECIPROCATING ENGINES

4-9. OPERATING CHARACTERISTICS

The reciprocating engine is one of the cost efficient power plants used for
aircraft power. Combined with the propeller, the reciprocating engine
efficiently converts fuel chemical energy into flying time or distance.

4-17
engine completes one
a. Operating Cycle. The typical reciprocating
operating cycle with four strokes of the piston. This cycle is shown in
Figure 4-11 by the variation of pressure and volume within a
cylinder.

(1) The first stroke of the cycle is the downstroke of the piston
with the intake valve open. This stroke draws in
a
fuel-air mixture charge
along line AB of the pressure-volume diagram. The second stroke compresses
the fuel-air mixture along line BC.
A spark ignition apparatus initiates
The
combustion, which takes place in an essentially constant volume.
combustion of the fuel-air mixture liberates heat, causing
a
rise of pressure
along line CD. The power stroke uses the increased pressure through
expansion along line DE. The exhaust then begins by the initial rejection
The area BCDE
along line EB and is completed by the upstroke along line BA.
on the pressure-volume diagram idealizes the network produced
by the cycle of
operation.
(2) During the actual cycle of operation, the intake pressure is
a pumping
lower than the exhaust pressure and the negative work represents
loss. The incomplete expansion during the power stroke represents a
basic
loss in the operating cycle because of the rejection of combustion products
along line EB. The area EFB represents
a basic loss in the operating cycle
because of the rejection of combustion products along line EB. The area EFB
represents a certain amount of exhaust gas energy. Exhaust turbines can
be coupled to the
extract part of the energy as additional shaft power to
crankshaft (turbo-compound engine) or to be used in operating the
In addition, exhaust gas energy may be
supercharger (turbo-supercharger).
used to augment engine cooling flow.

(3) The network produced during the operating cycle is represented


by the enclosed area of the pressure-volume diagram in Figure 4-11; engine
The weight of
output is affected by any factor that influences this area.
the energy released by combustion, and the
the fuel-air mixture determines
weight of charge can be altered by such factors as altitude and super-
charging. Mixture strength, preignition, and spark timing can affect the
the
energy release of a given airflow and alter the work produced during
operating cycle.

(4) The mechanical work accomplished during the power stroke is the
of the
result of the gas pressure sustained on the piston. The linkage
piston to a crankshaft by the connecting rod applies torque to the output
shaft. When pressure is converted to mechanical energy, certain losses are
inevitable because of friction and because the mechanical output is less than
available pressure energy. The magnitude and rate of power impulses
A brake or load device attached
determine the power output from the engine.
to the output shaft of the reciprocating engine measures the
power output.
The term "brake horsepower" denotes the output of the power plant.

4-18
INTAKE COMPRESSION COMBUSTION POWER EXHAUST

~
COMBUSTION

PRESSURE

COMPRESSION
E
"
....

INTAKE '.....
A ""....
.. ~
B. - - - - -

-- :. ..
- F

EXHAUST

VOLUME

Figure 4-11. Reciprocating engine operating cycle

(5) Output power is appreciated as a direct variable of torque and


revolutions per minute. Of course, output torque is a function of the
combustion gas pressure during the power stroke. It is helpful to consider
the mean effective pressure during the power stroke--the brake mean effective
pressure. The BMEP is not actual pressure within the cylinder; it is an
effective pressure representing the mean gas load acting on the piston during

4-19
the power stroke. As such, BrffiP is a convenient index for referring to
reciprocating engine output, efficiency, and operating limitations.

(6) The actual power output of any reciprocating engine is a


direct
function of engine torque and rotative speed. Thus, output brake horsepower
can be related by the combination of "3HEP and RPH or torque pressure and RPH.
No other engine instruments can immediately indicate this power output. If
all other factors are constant, the engine power output is directly related
to engine airflow. The reciprocating engine can be considered primarily as
an air pump with the pump capacity directly affecting the power output. Any
engine instruments which relate factors affecting airflow can indirectly
reflect engine power. The pressure and temperature of the fuel-air mixture
decide the density of the mixture entering the cylinder. The carburetor air
temperature provides the temperature of inlet air at the carburetor. While
carburetor inlet air is not the same temperature as the air in the cylinder
inlet manifold, the carburetor inlet temperature provides a stable indication
independent of fuel flow; this temperature can be used as a standard of
performance. Cylinder inlet manifold temperature is difficult to determine
with the same degree of accuracy because of normal variation of the fuel-air
mixture strength. Inlet manifold pressure also indicates the density of
airflow entering the combustion chamber. Manifold absolute pressure is.
affected by the carburetor inlet pressure, throttle position, and super-
charger or impeller pressure ratio. The throttle is the principal control of
manifold pressure, and the throttling action controls the pressure of the
fuel-air mixture delivered to the supercharger inlet. Pressure received by
the supercharger is magnified in some proportion, depending on impeller
speed, and the high-pressure mixture is then delivered to the manifold.

(7) Engine airflow is a function of RPM. A higher engine speed


increases the pumping rate and the volume flow through the engine. With the
engine-driven supercharger or impeller, an increase in engine speed increases
the supercharger pressure ratio. With the exception of a near-closed
throttle position, an increase in engine speed will produce an increase in
manifold pressure.

(8) An important aspect of reciprocating engine operation is the


many variables affecting the character of combustion. Uniform mixtures of
fuel and air support combustion between fuel-air ratios of approximately 0.04
and 0.20. The chemically correct proportions of air and hydrocarbon fuel are
15 pounds of air for each pound of fuel or a fuel-air ratio of 0.067. This
chemically correct, or stoichiometric, fuel-air ratio provides the
proportions of fuel and air to produce maximum release of heat during
combustion of a given weight of mixture. If the fuel-air ratio is leaner
than stoichiometric, the excess of air and the deficiency of fuel produce
lower combustion temperatures with a reduced heat release for a given weight
of charge. If the fuel-air ratio is richer than stoichiometric, the excess
of fuel and the deficiency of air produce lower combustion temperatures with
a reduced heat release for a given weight of charge.

(9) Stoichiometric conditions produce maximum heat release for ideal


conditions of combustion; they may apply quite closely for the individual

4-20
cylinders of the low-speed reciprocating engine. Because of the effects of
flame propagation speed, fuel distribution, and temperature variation, the
maximum power obtained with a fixed airflow occurs at
fuel-air ratios of
approximately 0.07 to 0.08. Figure 4-12 shows the variation of output power
with fuel-air ratio for a constant engine airflow; for example, constant RPM,
MAP, and carburetor air temperature. Combustion can be supported by fuel-air
ratios somewhat greater than 0.04, but the energy released is insufficient to
overcome pumping losses and engine mechanical friction. The same result is
obtained for the rich fuel-air ratios just below 0.20. Fuel-air ratios
between these limits produce varying amounts of output power; the maximum
power output generally occurs at fuel-air ratios of approximately 0.07 to
0.08. This range of fuel-air ratios, which produces maximum power for a
given airflow, is termed the best power range. At some lower range, a
maximum of power per fuel-air ratio is obtained; this is the best economy
range. The best economy range generally occurs between fuel-air ratios of
0.05 and 0.07. When maximum engine power is required for takeoff, fuel-air
ratios greater than 0.08 are necessary to suppress detonation. Fuel-air
ratios of 0.09 to 0.11 are typical during takeoff.

BEST CONSTANT
100
POWER AIRFLOW
BEST
ECONOMY
~-0
PERCENT
POWER

OVERLEAN OVER RICH

0
0 .04 .067 .08 .10 .20
FUEL-AIR RATIO

Figure 4-12. Power output in relation to fuel-air ratio


(10) The pattern of combustion in a cylinder is illustrated in Figure
4-13. The normal combustion process begins by spark ignition toward the end
of the, compression stroke. The electric spark provides the beginning of
combustion, and a flame front is propagated smoothly through the compressed
mixture. Such normal combustion is shown by the plot of cylinder pressure
versus piston travel. Spark ignition begins with a smooth rise of cylinder
pressure to some peak value with subsequent expansion through the power
stroke. The variation of pressure with piston travel must be controlled to
achieve the greatest network during the cycle of operation. Spark-ignition
timing is an important factor for controlling the initial rise of pressure in
the combustion chamber. The ignition of the fuel mixture must begin at the

4-21
and the release of heat to build
proper time to allow flame-front propagation
up peak pressure for the power stroke.

w DETONATION PREIGNITION
NORMAL COMBUSTION a:

z~
0111
-w
I-a:
~o..
SPARK ma:
PLUG :Ew
om
u:E

FLAME
Õ/J EXPLOSIVE
BURNING PREMATURE
PROPAGATION IGNITION
FROM HOT SPOT
PREIGNITION
NORMAL COMBUSTION
.--

COMPRESSION STROKE POWER STROKE

TOP CENTER

Figure 4-13. Pattern of cylinder combustion

(11) The speed of flame-front propagation is a major factor affecting


the rate of
the power output of the reciprocating engine, because it controls For this
chamber.
heat release and rate of pressure rise in the combustion high
plants with specific power
reason, dual ignition is necessary for power
be accomplished more rapidly with the propaga-
output. Normal combustion can
The two sources of ignition
tion of two flame fronts rather than one. shorter period
accomplish the combustion heat release and pressure rise in
a

the flame propagation speed in the


of time. The fuel-air ratio also affects
combustion chamber. Maximum f~ame-propagation speed occurs near a
fuel-air
given airflow tends to occur at
ratio of 0.08; maximum power output for
a

this value rather than at the stoichiometric value.


Preignition and Detonation. Two undesirable conditions of the
b.
combustion process are preignition and detonation. preignition is a
by hot spots in the
premature ignition and flame-front propagation caused
as feathered
combustion chamber. Various lead and carbon deposits, as well
glow ignition spot and start a flame
edges on metal surfaces, can supply
a

propagatiQn before normal spark ignition.

(1) in Figure 4-13, preignition causes


As shown
a premature rise of
a
result, preignition combustion pressures
pressure during piston travel. As
and temperatures exceed normal combustion values and are
cause likely to

4-22
engine damage. Because of a premature rise of pressure toward the end of the
compression stroke, the network of the operating cycle is reduced.
Preignition is evidenced by a rise in cylinder-head temperature and a drop in
BMEP or torque pressure.

(2)
Detonation can immediately destroy a power plant. The normal
combustion process is initiated by the spark and the beginning of flame-front
propagation. As the flame front is propagated, the combustion-chamber
pressure and temperature begin to rise. Under certain conditions of high
combustion pressure and temperature, the mixture ahead of the advancing flame
front may suddenly explode with considerable violence; this sends strong
detonation waves through the combustion chamber. The results are shown in
Figure 4-13; a sharp, explosive increase in pressure takes place with a
subsequent reduction of the mean pressure during the power stroke. Detona-
tion produces sharp, explosive pressure peaks many times higher than normal
combustion. The exploding gases also radiate considerable heat, causing
excessive temperatures on other parts of the engine. The effects of heavy
detonation are so severe t~at structural damage is the immediate result. A
rapid rise of cylinder-head temperature, a rapid drop in BMEP, and loud,
explosive noises are all evidence of detonation.

c. Power Output.

(1) Brake specific fuel consumption.


The primary factor relating
the efficient operation of the reciprocating engine is the brake specific
fuel consumption, or simply C. The BSFC can be computed by using the
following equation:
BSFC = engine fuel flow. C 1b/hr
BHP , or =

BHP
(Equation 4.1)

Typical minimum values for C range from 0.4 to 0.6 pounds per hour per BHP;
most aircraft power plants have a 0.5 average. Turbo-compound engines that
approach values of C 0.38 to 0.42 are generally the most efficient because
=

of the power-recovery turbines. The minimum values of specific fuel


consumption are obtained only within the range of cruise power operation--30
to 60 percent of the maximum power output. Generally, the conditions of
minimum specific fuel consumption are achieved with automatic lean or manual
lean scheduling of fuel-air ratios and high BMEP and low RPM. Low RPM is the
usual requirement to minimize friction horsepower and improve output
efficiency.
(2) Supercharging. Since altitude reduces engine airflow and power
output, supercharging is necessary to maintain high power output at high
altitude. Because the basic engine is able to process air only by basic
volume displacement, the function of the supercharger is to compress the
inlet air, providing a greater weight of air for the engine to process.
(a) The nonsupercharged or naturally aspirated engine cannot
provide manifold pressure greater than the induction system inlet pressure.
a

As altitude is increased with full throttle and governed RPM, the airflow
through the engine is reduced and BHP decreases. The first forms of

4-23
supercharging were of a relatively low pressure ratio, and the added airflow
and power could be handled at full throttle within detonation limits. Such a
ground-boosted engine achieves higher output power at all altitudes, but an
increase in altitude produces a
decrease in manifold pressure, airflow, and
power output.

(b) More advanced forms of supercharging with higher pressure


ratios produce very large engine airflow. The typical case of altitude
supercharging produces such high airflow at low-altitude operation that full-
throttle operation cannot be used within detonation limits. At sea level,
the limiting manifold pressure produces a certain amount of BHP. Full-
throttle operation produces a higher MAP and BHP if detonation is not a
problem. In this case, full-throttle operation is unavailable because of
detonation limits. As altitude is increased with the supercharger or blower
at low speed, the constant MAP is maintained by opening the throttle. The
BHP increases above the sea-level value because of reduced exhaust back

pressure. Opening the throttle allows the supercharger inlet to receive the
sa~e inlet pressure and to produce the same MAP. Finally, the increase of
altitude requires full throttle to produce the constant MAP with low blower.
This point is termed the critical altitude or full-throttle height. If
altitude is increased beyond the critical altitude, engine MAP, airflow, and
BHP decrease.

(c) The critical altitude with


a
particular supercharger
of MAP and RPM. A lower MAP
installation is specific to given combination
a

could be maintained to some higher altitude, a lower engine speed would


produce less supercharging, and a given MAP would require a greater throttle
opening. Generally, the most important critical altitudes are specified for
maximum rated and maximum cruise-power conditions.

(d) Changing the blower to a high speed provides greater super-


charging but requires more shaft power and incurs a greater temperature rise.
Thus the high blower speed produces an increase in altitude performance
within the detonation lilnitations. The variation of BHP with altitude for
the blower at high speed shows an increase in critical altitude and greater
BHP than that obtainable at a low blower speed. Operation below critical
altitude at high blower speed requires limiting the manifold pressure within
the detonation limits.

(e) Since exhaust gases have considerable energy, exhaust


turbines provide source of supercharger power. The turbo-supercharger
a

allows control of the supercharger speed and output to very high altitudes
with a variable discharge exhaust turbine. The turbo-supercharger provides
engine airflow with increasing altitude by increasing the turbine and super-
charger speed. Critical altitude for the turbo-supercharger is usually
defined by the altitude that produces the limiting exhaust turbine speed.

(f) Altitudes less than the critical altitude do not greatly


affect the minimum specific fuel consumption of the supercharged engine. At
maximum cruise power, specific fuel consumption decreases slightly with an
increase up to the critical altitude. Above critical altitude, maximum

4-24
cruise power cannot be maintained. However, the specific fuel consuaption is
not adversely affected as long as automatic lean or manual lean power can be
used at the cruise power setting.

4-10. OPERATING LIMITATIONS

The rec.iprocating engine has achieved a great degree of refinement and


development. It is one of the most reliable aircraft power plants when
specific operating limitations are adhered to strictly.

a. Prevention of Detonation and Preignition. The most important


operating limitations are those that ensure detonation and preignition do not
occur. The aviator must ensure that proper fuel grades are used and that
MAP, BMEP, RPM, and CAT limitations are not exceeded. Since heavy detonation
or preignition is common to high airflow at maxicum power, detonation or
preignition is most likely to occur at takeoff. To suppress detonation or
allow greater power for takeoff, water injection is often used in the
reciprocating engine. At high power settings, the injection of a water-
alcohol mixture replaces the excess fuel required to suppress detonation and
manual 0 r autorna tic leaning procedures reduce the fuel-air ratio t award
A
maximum heat release values. better fuel-air ratio results in an increase
in power. In some instances, higher manifold pressure is used to produce
additional power. This antidetonant injection fluid requires alcohol and
water proportions quite different from those of the injection fluid for jet-
engine thrust augmentation. Since leaning of the fuel-air ratio is desired,
the injection contains quantities of alcohol, which prevent residual fluids
from fouling the system.

b. Use of Proper Fuel. When fuel grades are altered during operation
and the engine is operated on the next lower fuel grade, proper accounting
must be made for changes in operating limitations. This accounting is made
for maximum power at takeoff and maximum cruise power, since both operating
conditions approach the detonation envelope. When higher grade fuel becomes
available, higher operating limits cannot be used until it is determined that
no contamination exists from lower grade fuel remaining in the tanks.

c. Fouling of Spark Plug. Spark plug fouling provides certain high, as


we 11as low, limi ts of operating temperatures. When an excessively low
operating temperature is encountered, rapid plug carbon fouling takes place.
On the other hand, an excessively high operating temperature produces plug
fouling from fuel additive lead bromide deposits.

d. Time Limits at High Power Settings.

(1) Generally, a time limit at various high power settings is


established to minimize the rate of wear and fatigue damage. Minimizing the
total time spent at a high power setting achieves a greater power plant
overhaul life. This should not imply that the takeoff rating of the engine
is not used. Full maximum power at takeoff actually accumulates less total
engine wear than a reduced power setting at the same RPM, because less time

4-25
is required for climbing to a
given altitude or accelerating to a
given
s pe e d .

(2) The most severe rate of wear and fatigue damage occurs at high
RPM and 1m" MAP. High RPH produces high centrifugal loads and reciprocating
inertia loads. When large reciprocating inertia loads are not cushioned by
high compression pressures) critical resultant loads are produced.
Therefore) operating time at maximuQRPM and maximuI:! MAP must be held to a
minimum. Operating time at maximum Rfl1 and low MAP must be avoided
altogether.

4-26
CHAPTER 5

HELICOPTER DESIGN AND MECHANICAL CHARACTERISTICS

This chapter describes the design and mechanical characteristics of the


typical helicopter. The aerodynamic forces acting on a helicopter, coupled
with the design characteristics, allow it to be maneuvered and controlled.

SECTION I

HELICOPTER CONTROL

5-1. BASIC SINGLE-ROTOR DESIGN

The basic design of the single-rotor helicopter consists of the fuselage,


main rotor, and tail rotor. Figure 5-1 shows a basic helicopter.

TAIL ROTOR
---....:..---
"
-...... ,

(
-

\
-- :::---

FUSELAGE

Figure 5-1. Basic helicopter

a. Fuselage. The fuselage is the primary structure of the helicopter.


It houses the engine, drive trains, crew and passenger compartments, flight
instruments, and avionics equipment.

b. Main Rotor. The main rotor produces the


needed to maneuver the helicopter.
lift
and thrust forces
The three primary types of main rotor
systems are the fully articulated, semirigid, and rigid.

c. Tail Rotor. The tail rotor provides the antitorque thrust required
to maintain the aircraft's heading and to turn the helicopter at low air-
speeds. Some tail rotors also provide a small amount of lift.

5-1
ROTOR SYSTEM TERMS
5-2.
The common terms used to describe helicopter rotor systems are shown in
Figure 5-2. Although the systems vary between aircraft, most manufact~rers
generally accept these terms.
A of Figure 5-2 is fully articulated. The
a. The system shown in part
semirigid type shown in part B of Figure 5-2 has no vertical or horizontal
hinge pin. Instead, the trunnion bearing that connects the yoke to the mast
allows the rotor to teeter or flap.

A FULLY ARTICULATED

SPAN

MAST

BLADE GRIP FEATHER BEARING

B SEMIRIGID

TRUNNION BLADE GRIP FEATHER BEARING


BEARING
Ð
.

SPAN CHORD
MAST
/

TRUNNION
BEARING

Figure 5-2. Rotor systems

b. The terms described below are keyed to the circled numbers in


Figure 5-2.

(1) longitudinal dimension of an airfoil section.


Chord. The It is
measured from the leading edge to the trailing edge.

5-2
(2) Span. The length of the rotor blade from the point of rotation
to the tip of the blade.

(3) Vertical hinge pin (drag hinge). The axis that permits fore and
aft blade movement (lead and lag) independent of the other blades in the
system.

(4) Horizontal hinge pin. The axis that permits up and down move-
ment (flapping) of the blade independent of the other blades in the system.

(5) Trunnion. Splined to the mast and secured to the yoke through
two bearings. The blades, mounted to the yoke, are free to teeter (flap)
about the trunnion bearings.

(6) Yoke. The structural to which the blades are mounted and
member
that fastens the rotor blades to the mast through the trunnion and trunnion
bearings.

(7) Blade grip feather bearing. The bearing that permits the blade
to rotate about its spanwise axis so the blade pitch can be changed (blade
feathering).
ROTOR SYSTEM CHARACTERISTICS
5-3.

a. Blade Twist. To distribute the lift more evenly along


the rotor
blade, manufacturers build a
twist into the blade. The blade twist increases
the angle of incidence near the root where the blade speed is slower and de-
creases the angle of incidence near the tip where the blade speed is faster.
Because the outboard portions have lower angles of incidence, less lift is
concentrated near the blade tip than at the root. Speed at any point on the
blade varies with the radius or distance the blade is from the center of the
main rotor shaft. The result is an extreme speed differential between the
blade tip and root. However, the lift differential between the blade tip and
root is even greater because lift varies with the square of the speed. When
speed doubles, lift increases four times. In Figure 5-3, for example, the
lift at the 50 percent blade-span point would be only one-fourth as great as
the lift at the blade tip, assuming the airfoil shape and angle of attack are
the same at both points.

b. Lift Differential. Because of the potential lift differential along


the blade resulting primarily from speed variations, blades are designed with
a
twist. Blade twist provides a higher pitch angle at the root where speed
is lower and a lower pitch angle nearer the tip where speed is higher. The
twist distributes the lift more evenly along the blade and increases both the
induced-air velocity and blade loading near the inboard section of the blade.
Figure 5-3 compares the distribution of lift on a twisted blade and on an un-
twisted blade. The twisted blade generates more lift near the root and less
lift at the tip than the untwisted blade.

5-3
UNTWISTED BLADE

IDEALLY TWISTED
BLADE

ROTOR
SHAFT"
50% 100%

Figure 5-3. Distribution of lift on twisted and untwisted blades

c. Rotor and Gyro Comparison. Helicopter rotors are often compared to


gyroscopes because of their obvious physical similarities. Although rotors
may look like gyros when operating, rotors do not always act like gyros.
Gyros exhibit three distinct characteristics worth comparing to helicopter
rotors: stability, phase lag, and precession.
( 1) stabili ty .

(a) Gyros are noted for their rigidity in space. A spinning


gyro is very stable and resists changes in the plane of rotation. This char-
acteristic is used as the foundation for some flight instruments such as the
attitude indicator.
(b) Rotors have plenty of rotational inertia, but little of the
gyros' stability. Helicopter rotors are capable of generating large aerody-
namic forces that can easily overwhelm the stability of the rotating mass in
the rotor. Changes in the cyclic pitch, turbulence, or center of gravity all
affect the aerodynamic forces on the rotor systems that ultimately result in
changes to fuselage attitude. Helicopter rotors add less stability to the
fuselage than do the wings of an airplane. stability can be enhanced, how-
ever, through a variety of methods; for example, the Bell stabilizer bar is
used on many older designs and the electromechanical stabilization system on
modern helicopters.

(2) Phase lag.

(a) Ifforce is applied to a gyro at a given point in the


a

plane of rotation, the gyro responds by tilting in the direction of the


applied force 90 degrees later in the direction of rotation. This phase lag,
between where the force is applied and where the maximum displacement occurs
seems to contradict normal physical laws. In fact, when a single point on a
gyro is analyzed, it is found to be reacting similar to an object in linear
motion, as shown in part A of Figure 5-4.

5-4
b
A LINEAR MOTION
0 DISPLACEMENT
A'
t
0 FROM ORIGINAL
~ LINE OF TRAVEL
0
O~
-.. ~å""
0 0 ORIGINAL
--.. 0 --Þo 0 ...
- - - - - - - - -

DIRECTION
.Â.
T FORCE

AIR JET I
1]

B CIRCULAR MOTION
AIR JET ~
1)
FORCE f ~ 0 HIGH
~
0
O~ C FlAT DISK b
POINT

I CREATEDBY~
~ ""
"'-
~ /~BAl~ (!

t?'~ ~I~~
0<>,,0+ DISPLACEMENT

~ON~<....o'..t; FROM ORIGINAL

o~
LINE OF TRAVEL

/ 0
6
.

a ,,~
,........--ORIGINAL DIRECTION

~o/
LOW d

()~
ORIGINAL
~ - ~
--..
- - -
POINT DIRECTION
0 ~ y
G::> a .
n
FORCE
THE TIME IT TAKES THE
' BAlL TO REACH MAXIMUM
\. \ AIR JET DEFLECTION FROM AN APPLIED
FORCE IS KNOWN AS PHASE LAG.

Figure 5-4. Force acting on a moving object

(b) of Figure 5-4 shows what happens to a ball rolling


Part A

ina
straight line if it is subjected to some lateral force such as that
caused by the air jet in the example. Where the force is applied (point a),
the ball starts to change direction. Even after the force is discontinued,
the ball continues to roll in the new direction. The further the ball rolls
in the new direction (point b), the further it is displaced from its original
line of travel (point a).
In part B of Figure 5-4, a similar ball attached to a
(c)
string is traveling in a circle. If an air jet causes the ball to start
rising (point a), the ball tends to continue in this new direction as the
ball did in the first example. The ball, continuing to climb, is higher at
point b than it is at point a. Since the path of the suspended object ap-
pears as a flat circle, or disk, the highest point of the disk must occur 90

5-5
degrees after the object passes over the air jet. Beyond the 90-degree
point, the object must start back down so that the disk remains flat. The
distance (time) it takes the ball to reach maximum displacement from where
the force was applied is called phase lag. This 90-degree phase lag is char-
acteristic of all rotating objects hinged at the axis of rotation. Semirigid
rotors have this same 90-degree phase lag between maximum applied force and
maximum displacement of a blade. Centrifugal force somewhat affects the
phase lag of blades that are not hinged exactly at the axis of rotation; for
example, those used on a fully articulated rotor. It causes the amount of
phase lag to be reduced to less than 90 degrees. This reduction is usually
only a few degrees and is compensated for by the flight control design.

(3) Precession.

(a) previously discussed, gyros react in a direction that is


As
90 degrees beyond an applied force. The movement of a gyro that results from
phase lag is called precession. As long as the force is applied to the same
point, the gyro will continue to precess at some specific rate. The rate of
precession is a function of the force applied, the rotational velocity of the
gyro, and the weight (mass) of the rotating body. Simply stated, a gyro
precesses faster with increased force and slower with increased RPM or mass.

(b) In this respect, helicopter rotors act like gyros. Moving


the cyclic creates a force differential across the rotor disk. As a result
of this force, the rotor begins to precess at some given rate, which is de-
termined by how far the cyclic is moved. Moving the cyclic further will
increase the rate of precession and, therefore, the roll or pitch rate of the
helicopter. The rotor can also be precessed by outside forces such as wind
or turbulence. To ensure satisfactory control for all authorized flight
maneuvers, helicopter manufacturers design rotors with adequate precession
rates.

5-4. ROTOR BLADE MOVEMENTS

Three rotor blade movements are involved in rotor systems: flapping,


feathering, and hunting. A fully articulated rotor system allows the rotor
blade to undergo all three movements, a semirigid system only flapping and
feathering, and a fully rigid system only blade feathering. In the rigid
rotor system, flapping and hunting are accomplished by flexing and bending
the blade.

a. Flapping. Flapping is the upward and downward movements of a rotor


blade about a flapping hinge during rotation. It is referenced to a plane
perpendicular to the drive shaft axis of the rotor. Blade flapping is caused
by the aerodynamic forces. When the tip-path plane is perpendicular to the
drive shaft axis, no flapping occurs. Flapping takes place around the trun-
nion bearing in the semirigid system and around the horizontal hinge pin in
the fully articulated system, as shown in Figures 5-5 and 5-6.

5-6
I MAIN ROTOR
~
I DRIVE SHAFT AXIS

TIP-PATH PLANE
~

-- -- -1-900-
\
-----
I
.....--
NO FLAPPING

--
The tip-path plane is perpen-
dicular to main rotor drive shaft
axis. No flapping is taking place.
""Þ'Þ-4'"
'11 ÞL.4 MAIN ROTOR
NE FLAPPING ANGLE DRIVE SHAFT
k' AXIS
--
JI

----LESS~ ~'
............

"" 900* I I
MORE
THAN 90" REFERENCE PLANE IS
PERPENDICULAR TO DRIVE
~ S~A~ ~I~O!.~A~ ROTOR

------~---
- - -

To get from point A to point a, the


blade must descend or flap down.
To get from point a to point A, the
""Þ.þ blade must climb or flap up.
'-4"''''
A
l.-4NE
MAIN ROTOR
I 50
----. ---- 1-" I~DRIVESHAFT
AXIS

~ 90"
REFERENCE PLANE IS
PERPENDICULAR TO DRIVE
~ I' ~ SHAFT AXIS OF MAIN ROTOR

The main rotor drive shaft is


mounted with a 50 forward till The
tip-path plane Is perpendicular to
the axis of the main rotor drive
shall No flapping is taking place.

Figure 5-5. Rotor blade flapping--semirigid rotor system

5-7
TIP-PATH PLANE

.------.-.,
-
- -
-
....

--
.....
"
-
- '\
,
"..
".. f
LAPPEOUP ,
".. BLAOE UPFLAP I
;I' ~ ANGLE /
,/
"
"
DOWNFLAp--.'Jo-
PLANE IS PERPENDiCuLAR
TO DRIVE SHAFT /
-/
/ ANGLE
AXIS OF MAIN ROTOR;I"
/
00~~
I \..p.~~EO .",
, OEf
,
,
\
ø\..Þ'

"., ./
" ..""
....... ---
-----
-

-- --
Figure 5-6. Blade flapping--fully articulated rotor system

b. Feathering. Feathering is the rotation of the blade about its span-


wise axis. It permits changes in the blade pitch angle and is accomplished
through the blade grip feather bearing. Feathering results from collective
or cyclic inputs, as shown in Figure 5-7.

BLADE ROTATES ABOUT ITS SPANWISE


AXIS. BLADE PITCH ANGLE CHANGES

Figure 5-7. Feathering

c. Hunting (Lead and Lag). Hunting is the fore and aft movement of the
blade in the plane of rotation in response to changes in the angular velocity
of the blade. The vertical hinge pin (drag hinge) is the axis that permits
this movemént independent of other blades in the system.

5-8
5-5. CONING

a. The rotor system depends primarily on blade rotation to produce rela-


tive wind, which develops the aerodynamic force required for flight. Because
of its rotation and weight, the rotor is subject to forces and moments pecu-
liar to all rotating masses. One of the forces produced is centrifugal
force--the force that tends to make rotating bodies move away from the center
of rotation. Centrifugal force is the dominant force affecting the rotor
system; all other forces act to modify this force. Another force produced in
the rotor system is centripetal force. It counteracts centrifugal force by
keeping an object a certain radius from the axis of rotation.

The rotating blades of a helicopter produce high centrifugal loads on


b.
the rotor head and the blade attachment assemblies. Centrifugal loads may be
from ó to 12 tons at the blade root of two- to four-passenger helicopters.
Larger helicopters may develop up to 40 tons of centrifugal load on each
blade root.

c. the rotor blades are not turning, they droop because of weight
When
and span. In fully articulated systems, the blade rests against a static or
droop stop, which prevents the blade from descending so low it strikes the
aircraft. When the rotor system begins to turn, the blade rises from
a

static position because of centrifugal force. At operating speed, the blades


extend straight out even though they are at flat pitch and are not producing
shown in Figure 5-8.
lift, as

24- TON LIFT FOR A SIX-BLADED SYSTEM

t CONED
STRAIGHT OUT
DROOPED

48,000 POUNDS GROSS WEIGHT

Figure 5-8. Flexible rotor blades supporting heavy load

d. the helicopter develops lift during takeoff and flight, the blades
As
rise above the straight-out position and assume a coned position (Figure 5-
8). The amount of coning depends on RPM, gross weight, and G-forces experi-
enced during flight. If RPM is held constant, coning increases as gross
weight and G-forces increase. If gross weight and G-forces are constant,
coning increases as RPM decreases. Excessive coning can occur if RPM is too

5-9
low, gross weight is too high, or G-forces are extreme. Excessive coning
will cause undesirable stress on the blade and a decrease in total lift
power.

e. Centrifugal force and lift effects on the blade can be illustrated


best by a vector. Figure 5-9 shows a rotor shaft and vector resulting from
the rotation of a blade. Figure 5-10 shows a vertical force vector acting at
the blade tip. This vertical force is lift produced when the blade assumes a
positive angle of attack. The blade is no longer being acted upon solely by
centrifugal force. It is also producing lift, which is manifested along the
blade by the lift vectors. Since one end of the blade is attached to the
rotor shaft, it cannot move. However, the rest of the blade moves and as-
sumes a position that is the resultant of the acting forces (vectors), as
shown in Figure 5-11. The blade position is coned and is the resultant of
two forces--lift force and centrifugal force.

CENTRIFUGAL FORCE

Cjf þ-

ROTOR
SHAFT

Figure 5-9. Centrifugal force

Cjf
ROTOR
CENTRIFUGAL FORCE

.~ LIFT FORCE
ON BLADE

SHAFT

Figure 5-10. Lift force and centrifugal force

DESU\,," AN")
B\..AOEtn

CENTRIFUGAL FORCE ~
LIFT FORCE
ON BLADE
ROTOR
SHAFT

Figure 5-11. Resultant lift force and centrifugal force

5-10
ROTOR HEAD CONTROL
5-6.

Cyclic and Collective Pitch. Aviator inputs to the collective and


a.
cyclic pitch controls are transmitted to the rotor blades through a complex
system. This system consists of levers, mixing units, input servos, station-
ary and rotating swashplates, and pitch-change arms (Figure 5-12). In its
simplest form, the movement of the collective pitch control causes the sta-
tionary and rotating swashplates mounted centrally on the rotor shaft to rise
and descend. The movement of the cyclic pitch control causes the swashplates
to tilt; the direction of tilt is controlled by the direction in which the
aviator moves the cyclic stick (Figure 5-13).

ROTOR HEAD

FEATHERING
BEARING

ROTOR
SHAFT.

ROTATING
SW ASHPLATE
BEARING CONNECTING
THE SWASHPLATES

STATIONARY
SWASHPLATE

CONTROL INPUT
SERVOS

Figure 5-12. Rotor head control systems

5-11
ROTATING PITCH-CHANGE ARM
SWASH PLATE
PULLS LEADING EDGE
OF BLADE DOWN.

STATIONARY
SWASHPLATE

-,
-

PIVOT
POINT f

..
Figure 5-13. Stationary and rotating swashplates tilted by cyclic control

b. Tilted Swashplate Assembly. Figure 5-14 illustrates a swashplate


that is tilted 2 degrees at two positions, points Band D. Points A and C
form the axis about which the tilt occurs. At that axis the swashplate re-
mains at 0 degrees. When the swashplate is moved, the resulting motion
change is transmitted to the rotor blade through pitch-change arms. As the
pitch-change arms move up and down with each rotation of the swashplate,
blade pitch constantly increases or decreases. If cyclic control is applied
to tilt the rotor, the addition of collective pitch does not change the tilt
of the swashplate and rotor. It simply moves the swashplate upward so that
pitch is increased an equal amount on all blades simultaneously, thereby
increasing the angle of attack and total lift.

c. Pitch-Change Arms. Figure 5-15 further illustrates how the pitch-


change arms move up and down on the tilted swashplate. The rate of vertical
change throughout the rotation is not uniform. Vertical movement is larger
during the 30 degrees of rotation at point A than at points Band C. This
variation is repeated during each 90 degrees of rotation. The rate of verti-
cal movement is the lowest at the low and high points of the swashplate and
the highest when the pitch-change arms pass by the tilt axis of the
swashplate.

5-12
PITCH-CHANGE
ARMS

t~
2"

TILT AXIS AC

Figure 5-14. stationary and rotating swashplates tilted in relation to mast

+2"

SWASHPLATE

-2"

Figure 5-15. Pitch-change arm rate of movement over 90 degrees of travel

5-13
d. Cyclic-Pitch Change.
Achange in cyclic pitch (cyclic feathering)
causes the rotor blades to climb from point A to point B and then dive or
descend from point B to point A, as shown in Figure 5-16. In this way, the
rotor is tilted in the direction of desired flight.
(1) To pass through points A and B, the blades must flap up and down
on a hinge or teeter on a trunnion. At the lowest flapping point (point A),
the blades would appear to be at their lowest pitch angle; at the highest
flapping point (point B), they would be at their highest pitch. only If
aerodynamic considerations were involved, this might be true. However, phase
lag causes these points to be separated by 90 degrees of rotation.

\
f==:::'- -
-
-
~B
-

Figure 5-16. Rotor flapping in response to cyclic input

(2) A cyclic stick movement decreases blade pitch at one point in


the rotor disk while simultaneöusly increasing blade pitch by the same amount
180 degrees of travel later. A decrease in
lift resulting from a decrease in
blade pitch angle causes the blade to flap down; the blade reaches its maxi-
mum downflapping displacement 90 degrees later in the direction of rotation.
An increase in lift resulting from an increase in blade pitch angle causes
the blade to flap up; the blade reaches its maximum upflapping displacement
90 degrees later in the direction of rotation. This results in a change to
the attitude of the rotor disk, as shown in Figure 5-17. The cyclic pitch
causing blade flap must be placed on the blades 90 degrees of rotation before
the lowest flap and highest flap are desired. Manufacturers consider phase
lag when they design rotors. They make sure that when the aviators push the
cyclic stick forward, the action tilts the swashplate assembly to place the
cyclic pitch appropriately. The lowest cyclic pitch on the blade needs to be
over the right side of the helicopter and the highest cyclic pitch over the
left side. The rotor always tilts in the direction in which the aviator
moves the cyclic stick.

5-14
C
~
-
-
EQUAL PITCH IN ALL
...
... -.... .....
BLADES. TIP-PATH PLANE
,- " PARALLEL TD HORIZON.
/ ,
I ,
I \
I \
I \
\
D .
B
I
J C
\
I
-~: --
\ -
- - -
- -

, / -
-
-

,
, /
I HORIZON - D ,'"
-........
-~
"", B
,
" '"
/ """""'--- --------
'"
...
......
... A
-
-
- ..,

A
WITHOUT CYCLE FEATHERING

C
A. POINT OF LOWEST BLADE PITCH
~ B. POINT OF GREATEST BLADE DISPLACEMENT DOWN
-
-
-
C. POINT OF GREATEST BLADE PITCH
.., ....
... .....
D. POINT OF GREATEST BLADE DISPLACEMENT UP
,- "
/ ,
I

'"
,
/ \
I DIRECTION OF \
I ROTATION \ TIP-PATH PLANE NO
\ LONGER PARALLEL TO
\
0 HORIZON
.
B
I
\
J
C
\ I
, /
,
, /
I
HORIZON
D~___-
-.~
, /

--
.
" '"

~-------
... '" B
...... ...
-
-
- .., A

.-----...
A

WITH CYCLIC FEATHERING

Figure 5-17. Phase lag (example)

5-15
e. Typical Design Features. Figure 5-18 illustrates a typical design
feature that offsets cyclic control input 90 degrees from where rotor tilt is
desired. Rotor control input locations are the left lateral servo (point A),
right lateral servo (point B), and fore and aft servo (point C). Each servo
is offset 45 degrees from the position corresponding to its name. The fore
and aft input servo, for example, is not located at the nose or tail position
3 o'clock
but at the right front and about halfway between the nose and the
position. Similarly, the left lateral servo is located halfway between the
nose and the 9 o'clock position. The right lateral servo is stationed half-
way between the tail and the
3 o'clock position. Locations of the input
servos account for part of the offset that is needed to correct for phase
lag. In addition, the rotor blade has
a
pitch-change horn that extends ahead
of the blade in the plane of rotation about 45 degrees. Aviator control
inputs are transmitted from the input servos to the pitch-change horn by
a

connecting rod called a pitch-change rod. The design of the pitch-change


horn, coupled with the placement of the servo, provides the total offset
necessary to compensate for phase lag.

HELICOPTER NOSE

t FORE AND AFT INPUT


(SERVO)
PITCH-CHANGE
HORN

LEFT
SIDE

MAIN
ROTOR
MAST

RIGHT LATERAL INPUT (SERVO)


HELICOPTER TAIL

Figure 5-18. Input servo and pitch horn offset

Cyclic Pitch Variation. Figure 5-19 illustrates the typical cyclic


f.
pitch variation for a blade through one revolution with the cyclic pitch
control full forward. The degrees shown are for a typical aircraft rotor
system; the figures would vary with the type of helicopter. As described in

5-16
the previous paragraph, the input servos and pitch-change horns are offset.
With the cyclic pitch control in the full-forward position, the blade pitch
angle is highest at the 9 o'clock position and lowest at the 3 o'clock posi-
tion. The pitch angle begins decreasing as it' passes the 9 o'clock position
and continues to decrease until it reaches the 3 o'clock position. As the
blade moves forward from the 3 o'clock position, the pitch begins to increase
and reaches the maximum pitch angle at the 9 o'clock position. Blade pitch
angles over the nose and tail are about equal.

(1) The blades reach a point of lowest flapping over the nose 90
degrees in the direction of rotation from the point of lowest pitch angle, as
shown in Figure 5-19. Highest flapping occurs over the tail 90 degrees in
the direction of rotation from the point of the highest pitch angle. Simply
stated, the force (pitch angle) that causes blade flap must be applied to the
blade 90 degrees of rotation before the point where maximum blade flap is
desired.

(2) A pattern similar to that in Figure 5-19 could be constructed


for other cyclic positions in the circle of cyclic travel. In each case, the
same principles apply. Points of highest and lowest flapping will be located
90 degrees in the direction of rotation from the points of highest and lowest
blade pitch.

LOW POINT OF ROTOR DISK


t ~o +9" +7"
TV

\ \ i I / i ïO+~
0
+10"

+1;12",\
+14~ \ n 1/+2"
+10
+150
" '\
I I
/ 0"
+160 ::
+17"
" I I /' -10

+18" '" I I LOW POINT


,/ -2"

+19"
-
"" ::
I I
r
SWASH PLATE
~ /'~ .--
-30

+20"-
~

.r FORE AND AFT _-40


HIGHEST
~ SERVO
LOWEST
.
PITCH
RIGHT LATERAL
-
--5'1,,0
.. PITCH

'~SERVO
,
-40

+19"
- ,
, ---~
.........

+18" ,;

+17"
+160
.//
/
/
PIVOT AXIS
',-2"
,," ,
.........

+10
0"
-10

+150
/ \ +2"
+140 \
+13:1t
/ \ \\
+~
+110
/ / I
+so
+40
+8"
+10" +9" +7" ..eo

..
HIGH POINT OF ROTOR DISK

Figure 5-19. Cyclic pitch variation--full forward, low pitch

5-17
5-7. BLADE LEAD AND LAG

Fully articulated rotors have hinged blades that are free to move
a.
fore and aft in the plane of rotation independent of the other blades in the
system. Movement around the hinge avoids bending stresses and is dampened by
a drag damper to avoid undesirable oscillations. When the helicopter moves
horizontally, the blade-pitch angle continually changes throughout each revo-
lution of the rotor to overcome dissYmmetry of lift. Pitch angle variation
creates changes in blade drag, causing the blade to lead or lag about the
drag hinge.

Another force, called Coriolis force, causes blades to lead and lag.
b.
Coriolis, a French mathematician, made a study of motion in a plane of rota-
tion caused by periodic mass forces. This type of motion is governed by the
law of conservation of angular momentum. The law states that a rotating body
will continue to rotate with the same rotational velocity until some external
force is applied to change the speed of rotation. Changes in the angular
velocity (angular acceleration or deceleration) will take place if the mass
of a rotating body is moved closer to or farther from the axis of rotation.
If the mass is moved closer to the axis of rotation, it accelerates. If the
mass is moved farther from the axis of rotation, it decelerates.

(1) The Coriolis force is illustrated by the ice skater performing


a
rotating movement. The skater begins rotating on one foot with both arms
and one leg extended outward from the body. Rotation is relatively slow
until the skater moves both arms and the leg closer to the body (axis of
rotation). Suddenly, rotational speed increases dramatically because the
skater's center of mass has moved closer to the axis of rotation. Mass did
not change, and an external force was not applied; however, velocity
increased.

(2) The principle of Coriolis force may be applied to a rotating


helicopter blade. A mass moving radially outward on a rotating disk will
A mass moving radially
exert a force opposite rotation (deceleration). in-
ward on a
rotating disk will exert a
force in the direction of rotation
(acceleration) .

Lead and lag can be explained by considering the following situation.


c.
A
helicopter is stationary on the ground in a no-wind condition with the
CG to the shaft
rotor turning. The distance from the blade axis is constant
throughout a complete blade revolution (points A and B, Figure 5-20). If the
cyclic stick is moved laterally, the rotor blade will climb on one side of
the disk and descend on the other side to produce a changed disk attitude.
Since the helicopter is stationary on the ground, the shaft axis about which
the blades are turning has not moved. The distance from the blades center of
gravity to the shaft axis changes continuously through each 360 degrees of
travel (points C and D, Figure 5-20). On the side where the blade climbs,
the radius (point D, Figure 5-20) decreases and the blade accelerates (lead).
The opposite side has a descending blade and an increasing radius (point C,
Figure 5-20), which causes the blade to decelerate (lag).

5-18
SHAFT
AXIS

~BC~
I~
ROTOR:
DISK CONE ~
I
I
I

\
SHAFT
AXIS
,
I
:
I

I~\
I
AXIS
I \
Dca'I
\ I I
\ I
c \ I

Figure 5-20. Radius change of blade CG relative to disk attitude

d. Figure 5-21, which shows a side view and a top view of the rotor
system, illustrates lead and lag on a four-bladed rotor. The blades are
evenly spaced 90 degrees apart on the level disk, and the tilted rotor disk
has uneven spacing between the blades because of lead and lag. At point A,
Figure 5-21, the blade has descended and has begun to decelerate. As the
blade decelerates, it lags enough to align with the rotor disk cone axis at
point B. At point C, the blade has climbed, decreasing the distance fr~_n CG
to shaft axis and resulting in an acceleration force. Point D shows the
blade is leading as a result of acceleration and has moved ahead to align
with the cone axis. This phenomenon occurs when the shaft axis and cone axis
are separated by a tilted rotor disk.

5-19
SHAFT
ROTOR DISK

~~IS
SHAFT CONE AXIS c
AXIS ,
\ I
I
.J--
I
---
."..., 1\
I'll

HOVERING~ A -

",
DISK LEVEL
ROTOR DISK
" ...- -
....
....
, ,,,,,-
/'" " TILTED .-
" , SHAFT
,
, I
I
, I AXIS
\ I
I
\ ,
I \ ,
, I
A
I I
, ,
\
I
\
, I
, ,
, ,
... "
"
.-
.. ",,'
"'- -- ROTOR DISK
CONE AXIS

Figure 5-21. Lead and lag

e. Because of the design (underslung) of the semirigid rotor system, no


change occurs in the travel radius of the CG of the blade associated with
blade flapping, as shown in Figure 5-22. Therefore, the angular velocity of
the blade does not change. Drag does impose significant stresses on the
blade roots; a drag brace is normally installed at the blade root to absorb
some of these bending forces.

5-20
I
I Because of the underslung rotor,
I A remains close to the same
distance from the mast after the
I MAST rolor is tilted: Compare A to A1

j..-
I
AXIS and B to B1-

I
I

A1

.. CG
I

This elbow moves away from the


This elbow moves toward the
mast as the rotor Is tilted.
mast as the rotor Is tilted.

Figure 5-22. Underslung design of semirigid rotor system

TORQUE
5-8.
According to Newton's law of action and reaction, the helicopter fuselage
tends to rotate in the direction opposite the rotor blades. This effect is
called torque (circle 2, Figure 5-23). Torque must be counteracted or con-
trolled before flight is possible. In tandem-rotor and coaxial helicopter
designs, the rotors turn in opposite directions to neutralize or eliminate
torque effect. In tip-jet helicopters, power originates at the blade tip
because of an equal and opposite reaction (Newton's third law). Since no
power is supplied to the rotor from the fuselage, no torque reaction occurs
between the rotor and fuselage. However, the torque effect is especially
important in the single main-rotor helicopters with a fuselage-mounted power
source. The torque effect on the fuselage (circle 2, Figure 5-23) is a di-
rect result of the work and resistance of the main rotor. Torque is at the
geometric center of the main rotor and results from the rotor being driven by
the engine power output through the transmission. Any change in engine power
output brings about a corresponding change in torque effect. Furthermore,
power varies with the flight maneuver and results in a variable torque effect
that must be continually corrected. Tandem-rotor helicopters require little
correction for torque changes because of their counterrotating rotors. See
paragraph 5-10b for more information on tandem-rotor helicopters.

5-21
Þ.Ð
0 ROTATION DIRECTION OF ENGINE.DRIVEN MAIN ROTOR.

A TORQUE EFFECT ROTATES FUSELAGE IN DIRECTION


V OPPOSITE MAIN ROTOR.
- - -

Ð TAIL ROTOR COUNTERACTS TORQUE EFFECT AND


PROVIDES POSITIVE FUSELAGE HEADING CONTROL.

Figure 5-23. Compensating torque reaction

5-9. ANTITORQUE ROTOR

Compensation for torque in the single main-rotor helicopter is accomplished


by means of a variable-pitch, antitorque rotor (tail rotor) located on the
end of the tail-boom extension at the rear of the fuselage. Except during
autorotation when it is driven by the main rotor, the tail rotor is driven by
the engine through the transmission at a constant ratio. The tail rotor pro-
duces thrust in a horizontal plane opposite the torque reaction developed by
the main rotor (circle 3, Figure 5-23). Since torque effect varies during
flight as power changes are made, the thrust of the tail rotor must be var-
ied. Antitorquepedals enable the aviator to compensate for torque variance.
From 5 to 15 percent of the available engine power may be required to drive
the tail rotor, especially during operations where maximum power is used.
The power needed depends on helicopter size and design. Normally, the larger
helicöpters use a higher percent of engine power to counteract torque than do
the smaller aircraft. A 9,500-horsepower helicopter might require 1,200
horsepower to drive the tail rotor; a 200-horsepower aircraft might require
only 10 horsepower to correct torque.

5-22
5-10. HEADING CONTROL

a. Single-Rotor Helicopters.

(1) In addition to counteracting torque, the tail rotor and its


control linkage allow the aviator to control the helicopter heading during
taxiing, hovering, and sideslip operations on takeoffs and approaches. Ap-
plying more pedal than needed to counteract torque causes the nose of the
helicopter to swing in the direction of pedal movement (left pedal to the
left). Applying ~ess pedal than needed causes the helicopter to turn in the
direction of torque (nose swings to the right). Aviators must use the anti-
torque pedals to maintain a constant heading at a hover or during a takeoff
or an approach. They apply just enough pitch on the tail rotor to neutralize
torque and to hold a slip.

(2) Heading control in forward trimmed flight is normally accom-


plished by cyclic control with a coordinated bank and turn to the desired
heading. The antitorque pedal must be applied when power changes are made.

NOTE: In autorotation, some degree of right pedal is required to maintain


correct pedal trim. When engine torque is not present, as in autorotation,
mast thrust bearing friction (frictional torque) tends to turn the fuselage
in the same direction as the main rotor turns. To counteract this frictional
torque, aviators sometimes reverse the tail rotor thrust and apply it in
a

direction opposite that required for torque correction in powered flight.

b. Tandem-Rotor Helicopters.

(1) is accomplished in tandem-rotor helicopters by


Heading control
differential lateral tilting of the rotor disks. When the directional pedal
(right or left) is applied, the forward rotor disk tilts in the same direc-
tion and the aft rotor disk tilts in the opposite direction. The result is
a
hovering turn around a vertical axis, midway between the rotors. Left
pedal application is illustrated in Figure 5-24.

Heading control in forward flight is accomplished by coordinated


(2)
use of lateral cyclic tilt
on both rotors for roll control and differential
cyclic tilt
on the rotors for yaw control. Only small changes in pedal trim
are required for changes in longitudinal speed trim or during descents,
climbs, and autorotations.

5-23
LEFT PEDAL

. FORWARD ROTOR DISK TILTS LEFT.


.
AFT ROTOR DISK TILTS RIGHT.
. AIRCRAFT ROTATES COUNTERCLOCKWISE ABOUT THE
CENTER.
. FUSELAGE PIVOTS.

Figure 5-24. Heading control (left pedal input)

5-11. TRANSLATING TENDENCY

a. During hovering tends to drift


flight, the single-rotor helicopter
laterally to the right. results from the thrust exerted to the
The tendency
right by the tail rotor for main rotor torque, as shown in
to compensate
Figure 5-25. The aviator may prevent this right lateral drift of the heli-
copter by tilting the main rotor disk to the left. This lateral tilt results
in a main rotor force to the left that compensates for tail rotor thrust to
.

the right.

5-24
0 THE TAPERED ARROWS INDICATE THE
ROTATION DIRECTION OFTHE ENGINE-
DRIVEN MAIN ROTOR.

THE TORQUE EFFECT ROTATES THE

t) FUSELAGE IN THE DIRECTION


OPPOSITE THE MAIN ROTOR.

e THE TAIL ROTOR COUNTERACTS THE


TORQUE EFFECT AND PROVIDES
POSITIVE FUSELAGE HEADING
CONTROL.

0.
THE TAIL ROTOR PULLS OR PUSHES
THE ENTIRE HELICOPTER INTO RIGHT
DRIFT (TRANSLATIONAL TENDENCY).

ø THE PilOT APPLIES LEFT ROTOR TILT,


AS NECESSARY, TO COUNTERACT TRANS-
LATING TENDENCY AND PREVENT
RIGHT DRIFT.

-+.
Figure 5-25. Compensating for translating tendency
b. Helicopter design usually includes at least one feature that helps
the aviator compensate for this translating tendency. One or more of the
.

following features may be included.

(1)Flight-control rigging may be designed so that the rotor disk is


tilted slightly to the left when the cyclic control is centered.
(2) The main transmission may be mounted so that the mast is tilted
slightly to the left when the helicopter fuselage is laterally level.
(3) The collective pitch control system may be designed so that the
rotor disk tilts slightly left when collective pitch is increased to hover
the aircraft.

5-12. FUSELAGE HOVERING ATTITUDE

a. Single-Rotor Helicopters.

(1) The design of most fully articulated rotor systems includes an


offset between the main rotor mast and the blade attachment point. Centrifu-
gal force acting on the offset tends to hold the mast perpendicular to the

5-25
tip-path plane (point A, Figure 5-26). When the rotor disk is tilted left to
counteract the translating tendency, the fuselage follows the main rotor mast
and hangs slightly low on the left side (point B,Figure 5-26).

OFFSET

OFFSET HINGING, COUPLED WITH HIGH


t

.~
CENTRIFUGAL LEADS AND LEFT TILT ...
OF ROTOR, CAUSES LEFT SIDE TO
HANG LOW.

Figure 5-26. Fully articulated rotor system

(2) fuselage suspended under a semirigid rotor system remains


A

level laterally unless the load is unbalanced or the tail rotor gearbox is
lower than the main rotor (point B, Figure 5-27). The fuselage remains level
because there is no offset between the rotor mast and the point where the
rotor system is attached to the mast (trunnion bearings). Because the trun-
nion bearings are centered on the mast, the mast does not tend to follow the
tilt of the rotor disk during hover. as Also,does the mast does not tend to remain
perpendicular to the tip-path plane it with the fully articulated
rotor system. Instead, the mast tends to hang vertically under the trunnion
bearings, even when the rotor disk is tilted left to compensate for the
translating tendency (point B, Figure 5-27). Because the mast remains verti-
cal, the fuselage hangs level laterally unless it is affected by other
forces.

5-26
A

TRUNNION
BEARINGS
~: I
:
I
..
MAST

I
I "II MAST
I

Figure 5-27. Semirigid rotor system

(3) the fuselage of the helicopter is tail low, the tail rotor
When
gearbox may be lower than the main rotor. Main rotor thrust acting to the
left above tail rotor thrust to the right causes the fuselage of the helicop-
ter to tilt
laterally to the left (point A, Figure 5-28). Although main
rotor thrust to the left is equal to tail rotor thrust to the right, it acts
at a greater distance from the CG, creating a greater turning moment on the
fuselage. This is more pronounced in helicopters with semirigid rotor sys-
tems than those with fully articulated rotor systems. Tail rotor thrust
acting at the plane of rotation of the main rotor would not change the at-
titude of the fuselage (point B, Figure 5-28).
(4) The main rotor mast in semirigid and fully articulated rotor
systems may be designed with a forward tilt
relative to the fuselage. During
forward flight, forward tilt
provides a level longitudinal fuselage attitude,
resulting in reduced parasite drag; during hover, it results in a tail-low
fuselage attitude.

5-27
MAIN ROTOR MAIN ROTOR
THRUST THRUST

MAIN ROTOR THRUST


TO LEFT TO COUNTER-
ACT TRANSLATING
....t .-.t
TAIL:OTOR .... TAIL ROTOR THRUST TO RIGHT TO
COMPENSATE FOR TORQUE EFFECT

TENDENCY AT A . THRUST
GREATER ARM FROM
THE CG CAUSES LEFT
TILT TO FUSELAGE

TAIL ROTOR THRUST


TO RIGHT TO
COMPENSATE FOR
TORQUE EFFECT

-
\Q~~OVE~
~~~ PLANE OF ROTATION

ANTITORQUE ROTOR PULLS


--~ -~-~ -0
TAIL ROTOR THRUST BELOW
MAIN ROTOR THRUST
LEFT SIDE LOW

Figure 5-28. Effect of tail-low attitude on lateral-hover attitude

b. Tandem-Rotor Helicopters. In tandem-rotor helicopters, the forward


and aft rotorsystems are tilted forward because of the transmission mounting
design. This tilt
causes decreasing excessive nose-low attitudes in forward
flight. Most tandem-rotor helicopters hover at a nose-high attitude of about
5
degrees. Some models will automatically compensate for this nose-high
attitude through automatic programming of the rotor systems.

SECTION
II
MECHANICAL CHARACTERISTICS AND REQUIREMENTS

5-13 . EFFECT OF ROTOR SYSTEM DESIGN ON WEIGHT AND BALANCE LIMITATIONS

Weight and balance limitations change dramatically with different configura-


tions of main rotor systems. The four types of systems are illustrated in
Figures 5-29, 5-30, and 5-31.

A
a. semirigid rotor system (Figures 5-29 and 5-31) supplies only sup-
port and mobility to the free-hanging pendulous mass of the fuselage. There-
fore, it has a limited allowance for CG travel.

5-28
BLADE CENTRIFUGAL FORCES HAVE NO INFLUENCE
ON MAST OR FUSELAGE ATTITUDE.

""---""'''''''''

\:~~S

SEMIRIGID ROTOR CG RANGE

Figure 5-29. Semirigid rotor system

b.In a fully articulated rotor system with offset hinges (Figures 5-30
and 5-31), blade centrifugal forces assist the fuselage to support wide-
range, center-of-gravity travel.

BLADE CENTRIFUGAL FORCES APPLIED TO OFFSET


HINGES HAVE STRONG INFLUENCE ON MAST AND
FUSELAGE ATTITUDE.

\
t .\
CF I~
/
FULLY ARTICULATED ROTOR CG RANGE

Figure 5-30. Fully articulated rotor system with offset hinges

5-29
A
c. rigid (hingeless) rotor system (Figure 5-31) provides rotor
rigidity in which centrifugal blade forces hold the fuselage level throughout
a wide CG tolerance laterally and fore and aft.

d. With differential collective pitch, a


multirotor system (Figure 5-31)
has the greatest allowance for CG travel.

SEMIRIGID ROTOR FULLY ARTICULATED ROTOR


-

ALLOWABLE CG
TRAVEL WITH
~~~~TM~~T
,.,
'.
----
~ ~
'
--- :/,/
~..
,. -~ -
-

--
---i----r--.. -"
RIGID ROTOR
MUL TIROTOR

---
(L~:--~
A
-fi) ---~
~
/
j/\ ð-
,~- --~
-

Figure 5- 31. Effect of different rotor systems on CG


travel

5-14. DANGER OF EXCEEDING CENTER-OF-GRAVITY LIMITS

a. Permissible CG travel is limited in many helicopters. The weight of


the crew, fuel, passengers, and cargo must be carefully distributed to pre-
vent the helicopter from flying with a dangerous nose-low, nose-high, or
lateral (side-low) attitude. If such CG attitudes exceed the limits of cy-
clic control, the rotor will be forced to follow the tilt of the fuselage
and control may be lost.
b. The helicopter will then move at a speed and in a direction that is
proportionate to the tilt of the rotor system. The amount of cyclic control
the aviator can apply to level the rotor system could be limited by the way
the helicopter is loaded. If the helicopter is loaded "out of CG limits,"
the aviator may find that when he applies corrective cyclic control as far as
it will go, the helicopter attitude will remain low in the direction CG lim-
its are exceeded. He will not be able to level the helicopter to decelerate
and land. Excessive loading forward of the CG, as shown in Figure 5-32,
creates an extremely dangerous situation.

5-30
CYCLIC CONTROL STICK
AGAINST REAR STOPS

~~:::.
~~;:::.-..-
~

--
.

:':::: II'
~

Figure 5-32. Excessive loading forward of the center of gravity

c. In newer helicopter designs, manufacturers have tried to minimize the


effects of CG travel by placing the loading compartment directly under the
main rotor mast. However, aviators must still arrange loads carefully to
assure that they are centered within the allowable CG-travel limits of the
helicopter as prescribed in the operator's manual.

d. The final CG is made


check for operationally at a hover just before
takeoff. In flight, CG
in the forward portion of the envelope results in
a

relatively nose-low attitudes. An aft CG condition results in relatively


nose-high flight attitudes. A lateral CG displacement results in a rela-
tively right-side or left-side, low-flight attitude. The correct procedure
is then to cruise with one side low. However, this procedure will cause an
uncentered slip indication and unlevel attitude indication.

e. trying to center the slip indicator and level the attitude


When
indicator, aviators make the common error of leveling the lateral CG attitude
by using the cyclic and opposite trim pedal. This causes a broadside rela-
tive wind and a lateral drag that levels the fuselage. This drag requires
more power at cruise, which then causes the range to be less. In autorota-
tion, fuselage drag results in a greater rate of descent, a shortened gliding
distance, and dangerous control problems.

5-15. PENDULAR ACTION

The fuselage of the helicopter has considerable mass and is suspended


a.
from single point. It is free to oscillate laterally or longitudinally
a

like a pendulum. Normally, the fuselage follows rules which govern pendu-
lums, balance, and inertia. Rotor systems, however, follow rules governing
aerodynamics, dynamics, and gyroscopics. These two unrelated systems form a
close and compatible partnership and normally avoid serious conflict.

5-31
and
b. other factors affect the relationship of the rotor system
and
fuselage. These factors are overcontrolling, cyclic-control response,
shift of attitude.
(1) Overcontrolling. Overcontrolling results when the aviator moves
the cyclic control stick causing rotor tip-path changes that are not re-
flected in corresponding fuselage-attitude changes. Correct cyclic control
movements (free of overcontrol) cause the rotor tip path and the fuselage to
move in unison. Erratic airspeed and altitude control may not be from
overcontrolling; they may result from a lack of knowledge of attitude-flying
techniques.

(2) Cyclic-control response. The rotor's response to cyclic control


input on the single-rotor helicopter has no lag. Rotor blades respond in-
stantly to the slightest touch of the cyclic control. The fuselage response
to lateral cyclic is noticeably different from the response to fore and aft
cyclic applications. Normally, considerably more fore and aft cyclic move-
ment is required to achieve the same fuselage response as that achieved from
an equal amount of lateral cyclic. This is not a lag in rotor response. It
is due to more fuselage inertia around the lateral axis than around the
longitudinal axis, as shown in Figure 5-33. For semirigid helicopters, the
normal corrective device is the addition of a synchronized elevator attached
to the tail boom and operated by the cyclic stick or a fixed horizontal sta-
bilizer.. The elevator or stabilizer produces lift forces that keep the fuse-
lage of the helicopter in proper alignment with the rotor at normal flight
airspeed. This alignment reduces blade flapping. The elevator or stabilizer
extends the allowable CG range of the helicopter; however, it is ineffective
at slow airspeeds.

LONGITUDINAL
LATERAL AXIS
AXIS

Figure 5-33. Cyclic-control response around the longitudinal axis


and the lateral axis

(3) Shift of attitude. Fuel cells normally have a slight aft CG.
As fuel is used, a slight shift to a more nose-low attitude occurs. Because
of the fuel expenditure and the lighter fuselage, cruise attitudes tend to
shift slightly lower. As fuel loads are reduced, the lighter fuselage is
affected more by drag, which results in a slight shift to
a more nose-down

attitude during flight.

5-32.
5-16. FUSELAGE ADD-ONS, FIXES, AND MODIFICATIONS

a. Fuselage Pitch-Attitude Changes.

(1) Fuselage nose-low attitude at cruise is typical of the single-


rotor helicopter. This condition is caused when the fuselage attitude aligns
itself to the tilted rotor disk at cruise airspeeds. It is also brought on
when the helicopter propulsion thrust is applied horizontally from the aero-
dynamic center of the main rotor. The total flat-plate drag of the fuselage,
centered many feet below the rotor, causes an additional nose-low influence.

(2) Tandem-rotor helicopters have both a preset transmission tilt


and a
programmed cyclic tilt.Cyclic tilting of the rotor disks automati-
cally increases with airspeed to minimize fuselage pitch-attitude changes
and blade flapping.

b.Fuselage Pitch-Attitude Corrections. Fuselage nose-low attitude is


usually corrected by mounting the transmission in the fuselage with some
degree of forward tilt. This presets the rotor system at the cruise airspeed
tilt angle, providinga level fuselage at cruise airspeeds. It can also be
a

corrected by adding horizontal stabilizer or synchronized elevator on the


tail boom. This counteracts the fuselage drag by holding the tail down and
the fuselage level in cruise flight.

c. Fuselage Add-On Devices. Fuselage add-on devices, external stores,


or sling loads are useful during certain modes of flight. However, these
add-on surfaces or devices often add flat-plate drag and develop side effects
at higher or lower airspeeds or during hover in crosswind or downwind condi-
tions. They also tend to reduce the payload and increase power requirements.
Fuselage add-on devices include--
0
Spoilers.
0
Sling loads.
0
External pods.
0
Dust or spray rigs.
0
Amphibious gear or floats.
0
Guns, cameras, or floodlights.
0
Fixed or controllable elevators.
0
Airfoil-shaped tail-rotor pylons.
0
Fixed-wing panels (experimental).
0
Ventral fins and vertical stabilizers.
0
External ordnance and related hardware.

5-33
d. Other Problems with the Pendulous Fuselage. Additional problems with
the pendulous fuselage include the following:

( 1) The weather-vane effect in crosswind hovering.

(2) Poor inherent pedal trim. The fuselage often drags slightly
sideward without an aviator-assist trim device.

(3) possibility of rotor blade strikes on the fuselage. Slope


The
operations or run-on landings with hard, jolting touchdowns and poor heading
control cause unacceptable force moments (or fuselage attitudes) that exceed
main rotor and fuselage compatibility. These impacts increase the possi-
bility of rotor blade strikes on the tail boom or the ground.

5-34
CHAPTER 6

ROTARY-WING PERFORMANCE

Aerodynamics of airplanes and helicopters are basically the same. Both use
airfoils to produce lift; both are subject to the same fundamental forces of
lift, drag, thrust, and gravity. Rotary-wing flight characteristics, how-
ever, differ greatly from those of fixed-wing aircraft. Helicopter perfor-
mance is discussed in this chapter. Airplane performance is covered in
Chapter 8.

SECTION I

BALANCE OF FORCES

6-1. NEWTON'S SECOND LAW OF MOTION

Newton's second law, covered in Chapter 1, is one of the most important


principles governing helicopter motion. It states that the rate of change of
motion of a body is directly proportional to the applied unbalanced force and
inversely proportional to the body's mass. This means that motion is
started, stopped, or changed when forces acting on the body become unbal-
anced. Rate of change (acceleration) depends on the magnitude of the unbal-
anced force' and on the mass of the body to which it is applied. This
principle provides the basis for all helicopter flight--vertical, forward,
backward, sideward, or 'hovering. In each case, the total force generated by
a
rotor system is always perpendicular to the tip-path plane (Figure 6-1).
For this discussion, this force is divided into two components--lift and
thrust. The vertical lift
component supports aircraft weight. The thrust
component acts horizontally to accelerate or decelerate the helicopter in the
desired direction. Aviators direct thrust in a desired direction by tilting
the tip-path plane.

TOTAL (RESULTANT) FORCE

------., I
,
\

TlP-PATH
!PLANE
ACCELERATION OR
FORWARD FLIGHT DECELERATION OR
REARWARD FLIGHT

MOVING (COASTING) HOVER


OR STATIONARY HOVER

Figure 6-1. Directing thrust by tilting tip-path plane


6-1
6-2. BALANCED AND UNBALANCED FORCES

a. At
a hover and in a no-wind conditiont all opposing forces are in
balance. The helicopter remains stationary, as illustrated in Figure 6-2.
The total force is acting opposite the aircraft weight. For a helicopter to
move in some direction, a
force must be applied to cause an unbalanced condi-
tion. Figure 6-3 shows an unbalanced condition. The aviator has changed the
attitude of the rotor disk; a lift vector and a thrust vector generate a
total force that is forward of the vertical. No parasite drag is shown be-
cause the aircraft has not started to move forward. As the aircraft begins
to move (accelerate) in the direction of the applied force (thrust), it de-
velops parasite drag. Parasite drag is covered in Chapter 2. When parasite
drag increases to equal the thrustt the helicopter no longer accelerates
because forces are again in balance, as shown in Figure 6-4. The aircraft
continues to move in the new direction and at the same speed until unbalanced
force is applied to change present motion.

TOTAL (RESULTANT) FORCE


t

~ WEIGHT

Figure 6-2. Balanced forces, helicopter hovering, no wind

WEIGHT

....
Unbalanced forces causing acceleration
Figure 6-3.

6-2
TOTAL (RESULTANT) FORCE
k""
---
LIFT
-.

~ ---
~

RESULTANT WEIGHT AND DRAG


+--
WEIGHT

Figure 6-4. Balanced forces, aircraft in motion

b. When the aircraft returns to a hover, the aircraft attitude changes


so the thrust vector is smaller than parasite drag, or directed to the rear,
as shown in Figure 6-5. Thrust and parasite drag now act opposite the for-
ward motion, so forces become unbalanced. The helicopter decelerates until
all motion stops. For the aircraft to remain stopped, the aviator must ad-
just the attitude to balance the forces, as shown in Figure 6-2.

I
---
~WEIGHT +--
RESULTANT WEIGHT AND DRAG

Figure 6-5. Unbalanced forces causing deceleration

6-3
SECTION
II
DEVELOPMENT OF AIRFLOW ON THE ROTOR SYSTEM

6-3. RELATIVE WIND

Relative wind is the movement of air molecules with respect to an object. In


other words, the relative wind acting on an airfoil may differ from the
rela-
tive wind acting on the fuselage of an aircraft.

a. The fuselage of a helicopter at a hover in a no-wind condition is


subject to a vertical flow of air (Figure 6-6) produced by the rotor system
(induced flow, or downwash). The rotor blades are subject to a
horizontal
flow of air (rotational relative wind), as well as a vertical flow of
air
(induced flow). The effects of rotational relative wind modified by induced
flow produce a flow of air (resultant relative wind) just ahead of the
air-
foil. This resultant wind direction and velocity act on the airfoil to pro-
duce the total aerodynamic force of the
airfoil. This resultant relative
wind also serves as the reference plane for the development of
and total aerodynamic force vectors on the lift, drag,
airfoil.
b. In forward flight, the situation changes somewhat. The airflow meet-
ing the fuselage (Figure 6-7) is opposite the flight path of the
aircraft;
however, the airflow meeting an airfoil is subject to a number of modifica-
tions. These modifications result from--
0
Coning.
0
Blade twist.
0
Blade flapping.
0
Forward airspeed.
0
Rotor system tilt:
0

Rotational velocity.
0
Pitch angle of the blade.
0

Vertical velocity of the aircraft.


0
Blade element position in the disk area at a
given instant.

6-4
AIRFLOW AT A HOVER

I t I

!j ~

RESULTANT

~
~ RW VECTOR

INDUCED FLOW
VECTOR

ROTATIONAL
RW

6 6 Airflow at hover
Figure -
.
a

6-5
.. I-
Z
<I:
I-
...J~
.. ~a:

2~~~
w

~--
RESULTANT RW a
-
a:
\f /l-Sf',~
.- /l-f',\)
fOt:'\~:'{\O~/I-\..
f',0"(Þ;~
.. I
I .

..

..

RESULTANT RW

-~ ----,~100
'101:
1'0
II'
7'10/v.
II.
FIW4F10:;
FIW-

Figure 6-7. Airflow in forward flight

6-4. ROTATIONAL RELATIVE WIND

The rotation of the rotor blades as they turn about the mast produces
rotational relative wind. Rotational relative wind flows opposite the physi-
cal flight path of the airfoil or airfoil segment, striking the blade at 90
degrees to the leading edge and parallel to the plane of rotation. Rota-
tional relative wind velocity is highest at the blade tips, decreasing
directly as the blade span decreases, as shown in Figure 6-8. No rotational
wind velocity exists at the axis of rotation (center of the mast). The equa-
tion for computing blade-tip speed (BTS) is as follows:
BTS (in ft/sec) = rotor radius (in ft) x RPM
(Equation 6.1)
9.55

6-6
TIP SPEED =
800 ftlsec

75% SPAN =
600 ftlsec

50% SPAN =
400 ftlsec

.
25% SPAN =
200 ftlsec

AXIS OF ROTATION = 0 ftlsec

Figure 6-8. Rotational relative wind vectors

6-5. INDUCED FLOW

Induced flow, or downwash, is the component of air flowing vertically through


the rotor system. It results from the production of lift.

a. To produce the required to hover, the helicopter rotor system


lift
accelerates large a of air vertically and down through the rotor system.
mass
The following analysis of the production of lift is known as the momentum
theory.
(1) the pitch angle of the blades must be increased
To produce lift,
from the flat-pitch position. This increases the pitch angle of the blade,
which increases the angle of attack and causes the air to be accelerated
downward.

Other factors involved in developing the downward velocity of


(2)
airflow, or induced flow, are the rotational velocity and the number of
blades in the rotor system. For example, for a four-bladed rotor system
operating at 240 RPM, 16 airfoils would pass any point in space each second.
The first airfoil to pass this point in space would meet air flowing opposite
the flight path of the airfoil. The air accelerating over the airfoil leaves
the with some slight downward velocity, as shown in Figure 6-9. The
airfoil
second airfoil meets a flow of air that already has some downward velocity.
The passage of this second airfoil causes further increases in the downward
velocity of the airflow. Each succeeding airfoil further increases this
downward velocity of the air until there is a vertical flow of air down
through the rotor system when the helicopter is at a hover. The blades are
adding energy (static pressure) to the airflow.

6-7
POINT IN SPACE POINT IN SPACE POINT IN SPACE POINT IN SPACE

AIRFOIL 1 AIRFOIL 2 AIRFOIL 3 AIRFOIL 4

Figure 6-9. Development of induced flow

b. This downward flow starts at an initial vertical velocity of zero,


about one rotor diameter above the rotor system (part A, Figure 6-10). This
downward velocity increases until the final downwash velocity is reached,
about one to two rotor diameters below the rotor system. The downward ve-
locity that exists at the rotor is the induced velocity, or average downward
velocity. The final downwash velocity is about twice the induced velocity at
the rotor. Part B, Figure 6-10, shows a velocity profile of this downwash.
These downwash velocities can reach 60 "knots beneath utility helicopters and
100 knots beneath cargo helicopters. The resulting velocity and pressure
changes occurring in the flow are shown in part C, Figure 6-10. Part C shows
the pressure changes that occur as air is accelerated down through the rotor
system. Pressure is less than ambient just above the rotor disk; energy is
added to the air by the blades as they rotate. This added energy manifests
itself as a
static pressure increase. Static pressure is higher than ambient
just below the rotor. This pressure returns to ambient about one to two
rotor diameters below the rotor disk, where the velocity is highest.

c. The lift produced by the rotor is equal to the change in pressure


across the rotor times the rotor-disk area. At a hover, this would be equal
to the weight of the helicopter.

6-8
INITIAL VELOCITY = 0

VELOCITY INITIAL

11
PLANE OF ROTOR

\\\llifl
VELOCITY FINAL =
2X VELOCITY INDUCED

VELOCITY FINAL =
A. AIRFLOW 2X VELOCITY INDUCED

B. VELOCITY

C. CHANGE IN PRESSURE

\\
PRESSURE = AMBIENT

\
VELOCITY INCREASES

11
PRESSURE LESS THAN AMBIENT

PRESSURE GREATER THAN AMBIENT

\ I I I
VELOCITY INCREASES

t ~ t! AMBIENT PRESSURE

MAXIMUM VELOCITY
PRESSURE = AMBIENT

Figure 6-10. Pressure change across rotor system (momentum theory)

6-9
6-6. RESULTANT RELATIVE WIND

The resultant relative wind athover is the rotational relative wind as


a

modified by the induced flow. This resultant relative wind (Figure 6-11) is
inclined downward at some angle and is opposite the effective flight path of
the airfoil, rather than the physical flight path (rotational relative wind).

-4NGI.
~O!: -4
7''1''-4
RESUl TANT CI( .....
RW
(EFFECTIVE FLIGHT
IF P ATH
OF AIRFOIL)

ROTATIONAL RW
(PHYSICAL FLIGHT PATH OF AIRFOil)

Figure "6-11. Resultant relative wind

6-7. AIRFLOW IN FORWARD FLIGHT

Airflow across the rotor system in forward flight varies somewhat


a.
from airflow at a hove~. In forward flight, air flows opposite the flight
path of the aircraft. The velocity of the flow of air equals the forward
speed of the helicopter. Because the blades of the helicopter turn in
a

blade depends on the


circular pattern, the velocity of the airflow across
a

position of the blade in the rotor disk at


a
given instant, its rotational
velocity, and the airspeed of the helicopter.

The airflow meeting each blade continually varies as the blade


(1)
rotates. Highest velocity of airflow occurs over the right side of the heli-
copter. Airflow decreases to rotational velocity over the nose; the decrease
continues until the lowest velocity of airflow is reached over the left side
of the helicopter. At that point, the velocity of the airflow starts to
increase, reaching rotational velocity over the tail and maximum velocity
over the right side (Figure 6-12).

(2) The blade over the right side of the helicopter moves in the
same direction as the helicopter. The velocity of the air meeting this ad-
vancing blade equals the rotational velocity of the blade plus the velocity

6-10
of wind resulting from forward airspeed (aircraft airspeed), as shown in
Figure 6-12.

(3) The blade over the left side of the helicopter moves in a flow
of air that is moving in the same direction as the blade. The velocity of
the airflow meeting this retreating blade equals the rotational velocity of
the blade minus forward airspeed (aircraft airspeed), as illustrated in
Figure 6-12.

(4) The blades over the nose and tail move essentially at right
angles to the airflow created by forward airspeed. Therefore, the velocity
of the airflow meeting these blades equals the rotational velocity, as shown
in Figure 6-12.

AIRSPEED = 120 KNOTS (200 ftIsec)

~
Q Q
800 ftlsec
RELATIVE WIND AS A RELATIVE WIND AS A
-....
RESULT OF AIRCRAFT -. ~- -- RESULT OF AIRCRAFT
.'
MOVEMENT AT 200 ftlsec

,
//'/" "''''''''''''''''
,
"

,
MOVEMENT AT 200 ftlsec

,
,
,

,
,
DIRECTION OF ROrATION
,
,
I
: 4 2

8
VELOCITY OF AIRFLOW
6 ,

,
,
IN ftlsec ALONG
I

,
, THE BLADE SPAN
,
I

"
,
, "
,

,/
,

, ,

ROTATIONAL VELOCITY

//
=
ROTATIONAL VELOCITY =
800 ftlsec 800 ftlsec
+
200 ftlsec = AIRCRAFT AIRSPEED
AIRCRAFT AIRSPEED =
200 ftlsec ,,'
""'''''''''''''-'''' -~,,,,,,
WIND VELOCITY 1,000 ftlsec = WIND VELOCITY
=
600 ftlsec 800 ftlsec

Figure 6-12. Differential velocities around rotor system


as a
result of forward airspeed

6-11
b. Change in airflow across the advancing and retreating blades causes a
number of changes in the rotor system. Forward airspeed is added to the
rotational velocity at all points along the span of the advancing blade to
determine the wind velocity at any given point when the blade is at the 3
o'clock position, as shown in Figure 6-12. Forward airspeed is subtracted
from all points along the span of the retreating blade to determine the wind
velocity at any given point along its span when the blade is at the 9 o'clock
position, as is also shown in Figure 6-12. The airflow across the hub of the
rotor system equals forward airspeed (Figure 6-12). This differential air-
flow velocity across the rotor system causes the aerodynamic forces to vary
along the blade span. Figure 6-13 depicts the force vectors acting on vari-
ous blade areas in forward flight.

(1) No-lift areas. The no-lift areas are reverse flow, negative
stall, and negative lift. These are depicted in parts A through C
of Figure
6-13.

(a) Reverse flow. At the root of the retreating blade is an


area where the air ~lows backward from the trailing edge to the leading edge
of the blade. This is because wind created by forward airspeed has a higher
velocity than the rotational velocity in this part of the blade span. The
reverse flow area extends from the mast to a point along the blade span where
the rotational velocity is equal to the airspeed of the helicopter. No posi-
tive lift is produced by this portion of the blade, as shown in part A,
Figure 6-13.

(b) Negative stall. In the negative stall area, rotational


velocity exceeds forward flight velocity, causing the resultant relative wind
to move toward the leading edge of the blade. However, rotational velocity,
induced flow, and flapping action do not increase the angle of attack above
the critical angle of attack. Therefore, this blade region is operating with
the resultant RW so far above the chord line that a negative stall angle of
attack results (part B, Figure 6-13).

(c)
Negative lift. In the negative lift area, rotational
velocity, induced flow, and blade flapping combine to reduce the angle of
attack from a negative stall to an angle of attack that causes the blade to
produce negative lift. In this area, the resultant relative wind is still
striking the airfoil above the chord line, but the angle of attack has been
reduced below the critical angle of attack. The air now has a greater dis-
tance to flow along the underside of the airfoil. This results in the
highest velocity and lowest pressure on the underside of the airfoil. The
resulting lift force vector in this blade region is pointing down, so this
blade region is producing negative lift, as shown in part C, Figure 6-13.

(2)Positive lift and positive stall. Positive lift is produced by


the blade outboard from the three no-lift areas, as shown in part D, Figure
6-13. In the positive lift region, the resultant relative wind is striking
the airfoil below the chord line, resulting in a positive angle of attack.
The air now has a greater distance to flow over the airfoil than it has to
flow beneath it. Because the highest velocity and lowest pressure are on the
top surface of the airfoil, the lift force is oriented up. This means the

6-12
airfoil is producing positive lift. Under certain conditions, it is possible
to have a positive stall area near the blade tip. This is shown in part E,
Figure 6-13. Retreating blade stall is covered in paragraph 6-20.

NEGATIVE
A
LIFT
"

REVERSE
FLOW
~

L
~f.:; ~\.'( þo1

IF
ROTATIONAL
D
RW (Forward airspeed
'.
.. .. L Is greater than
T AF rotational velocity.,)
IF

,."...."-,, ""''''''....

,,',,''-''- "''''... "


""''''''...
,
"

,
,

'. ,
.

\
,
,
.
\
I .
,
I

E -- ~.E t:= D
0
. I
I
\
POSITIVE .
. ,
,
,

\
STALL ,

I
I
,

B
,
, /
NEGATIVE
,
, "
,
,
,

,
,
,
STALL
,
,

/'
,

-'- "
" --
"'''''''''--... .....--.......,

ROTATIONAL.
RW
ANGLE OF ATTACK EXCEEDS THE CRITICAL ANGLE
TAF

ANGLE OF ATTACK EXCEEDS THE CRITICAL ANGLE

Figure 6-13. Blade areas in forward flight

6-13
SECTION
III
DIFFERENTIAL AIRFLOW IN THE ROTOR SYSTEM

6-8. CAUSES OF DISSYMMETRY OF LIFT IN THE SINGLE-ROTOR HELICOPTER

In forward flight, the combined effects of the differential airflow across


the advancing and retreating blades and the three no-lift areas on the
retreating blade result in a dissymmetry of lift potential (Figure 6-14)
between the advancing and retreating halves of the rotor disk.

120 knots (200 feet per second), as shown


a. At an aircraft airspeed of
in Figure 6-14, a 400-foot-per-second blade speed differential exists between
the advancing and retreating blades. Because lift increases with the square
of the velocity, a potential lift variation exists between the advancing and
retreating sides of the rotor disk. If the aviator fails to compensate for
this lift differential, the helicopter becomes uncontrollable.

b. the lift of the advancing half of the disk to the lift of


To compare
the retreating half, Equation 2.3 (L CL 1/2 P S V2) is used.
=
In forward
flight, two factors in the equation--air density and blade area--are the same
for the advancing and the retreating blades. The airfoil shape is fixed for
a
given blade. The only remaining variables are blade speed and blade angle
of attack. These two variables must compensate for each other during forward
RPM and aircraft
flight to overcome dissymmetry of lift. Two factors--rotor
airspeed--control blade speed during flight. Both can be varied to some
degree; however, they must remain within the operating limits specified for
the aircraft. Because blade speed is relatively constant, angle of attack
remains the one variable that can compensate for dissymmetry of lift. Through
cyclic feathering, the pitch angle of the rotor blades can be varied
throughout their range, from flat to stalling, to change the angle of attack
and to compensate for lift differential. Blade flapping also causes a change
in the angle of attack that helps to compensate for dissymetry of lift.

6-14
ROTATIONAL VELOCITY
MINUS FORWARD AIRSPEED

SMALL LIFT-PRODUCING AREA


"
/
RS

'
---
--
..- .- ---- -'--"
"---

\,.,~
~
-"'.
"-
DIRECTION OF
ROTATION

// ;;:;.,C
~:\;
,,~,,~
\\,
!

I
,
\
\
\
\
;;;~,<'i)

.
.
I
.
"

'
I
,
Q ~~R=:t~~~~R;=)D

\
\
I
\
,

\, I

I"

LARGE LIFT-PRODUCING AREA AS

ROTATIONAL VELOCITY
PLUS FORWARD AIRSPEED

Figure 6-14. Dissymmetry of lift


6-9. COMPENSATING FOR DISSYMMETRY OF LIFT IN THE SINGLE-ROTOR HELICOPTER

Blade Flapping. Blade flapping (Figure 6-15) is the up and down


a.
movement of the rotor blade about a flapping hinge. Flapping alone or in
conjunction with cyclic feathering can eliminate dissymmetry of lift and
allow the pilot to maneuver the helicopter. The following is a discussion of
blade flapping resulting only from the differential airflow across the ad-
vancing and retreating blades. The lift equation (L CL 1/2p S V2), Equa- =

tion 2.3, illustrates the cause of blade flapping; lift varies with the
square of the wind velocity (V2) across an airfoil.

(1) Wind velocity and upflap velocity.


(a) Maximum wind velocity/maximum upflap velocity. The
advancing blade meeting higher airflow velocities, caused by the addition of
forward flight velocity airspeed to rotational velocity, responds to this
increased airspeed by producing more lift. This additional lift causes the
blade to climb, or flap upward. The blade reaches its maximum upflapping
velocity at the point in the rotor disk where the maximum wind velocity
exists. This is normally at the 3 o'clock position, as shown in part A,
Figure 6-15.

6-15
MAXIMUM UPFLAP DISPLACEMENT OF BLADE
C
"
~.--- '" .. ~

~.. ..
~----'-'- ,

""""
- ,
, ,
,
,
,
,

MINIMUM WIND '. MAXIMUM WIND


VELOCITY \ VELOCITY
B. 'A
MAXIMUM DOWNFLAp: , : MAXIMUM UPFLAP
,
VELOCITY , ,: VELOCITY
, .
, .
I
, I
, ,
I
, -
I
-
,
,
,

~,~""
D
"''''''''''''''--''
-----""
-- --

AlJVG I MAXIMUM DOWNFLAP DISPLACEMENT OF BLADE


'l.~O
Þ
AI
1)-.
AlC/r
O....~~
JV08~
RESULTANT RW
IF

ROTATIONAL RW

F/
DOWN FLAP
VELOCITY
\ "'E
"
IF
/ 0
f,Of
p..~G'" p..\\.
/liE-v. "(
:f{
p..
p..C'"

RESULTANTRW
IF

Figure 6-15. Blade flapping

6-16
(b) Minimum wind velocity/maximum downflap velocity. The
retreating blade meeting progressively decreasing airflow velocities responds
by losing lift and descending, or flapping down. Because minimum airflow
velocity occurs at the 9 o'clock position, the maximum dåwnflapping velocity
of the blade is as shown in part B, Figure 6-15.

(c) Maximumupflap and downflap displacement of the blade.


Because of phase lag, the advancing blade responds to the maximum upflapping
velocity at the 3 o'clock position by continuing to flap up until maximum
upflap blade displacement is reached over the nose of the helicopter, as
shown in part C, Figure 6-15. The retreating blade responds to maximum
downflap velocity at the 9 o'clock position by continuing to flap down until
maximum downflap blade displacement is reached over the tail of the helicop-
ter, as shown in part D, Figure 6-15. This upflapping of the blade over the
nose and downflapping over the tail of the helicopter cause the tip-path
plane to tilt to the rear. This rearward tilt of the rotor disk is called
blowback.

(2) AnRle of attack. The upflapping and downflapping of the


rotor blades aerodynamically change the angle of attack and, therefore, the
amount of lift the blade produces.

(a) The upflapping blade causes air to move downward


through the rotor blades. This has the same effect as an increase in the
induced flow velocity, which decreases the angle of attack. The upflapping
of the blade reduces the angle of attack just enough to cancel any increase
in the lift expected because of the increased air velocity across the blade,
as illustrated in part E, Figure 6-15. Part E shows the angle of attack
decrease as the blade flaps up.

(b) The downflapping blade causes air to move upward


through the rotor blades. This has the same effect as reducing the induced
flow velocity and results in an increase in the angle of attack. The down-
flapping of the blade increases the angle of attack just enough to cancel
any loss of lift expected because of decreased airflow across the blade, as
illustrated in part F, Figure 6-15. Part F shows the angle of attack in-
crease as the blade flaps down.

(c) Upflapping and downf1apping do not change the amount


of lift produced by the rotor system. The blades flap to equilibrium. How-
ever, flapping changes the attitude of the rotor system (blowback) and,
therefore, the direction of the total lift vector (Figure 6-16). This re-
duces helicopter airspeed.

6-17
Blowback causes a change in attitude of rotor
system and therefore a change in the direction of
the total lilt vector.

Figure 6-16. Blowback

b. Cyclic Feathering. The aerodynamic flapping of the rotor blades as


they compensate for dissYmmetry of liftchanges the attitude of the rotor
disk; this directs the total thrust (lift)vector to the rear, which changes
airspeed. The aviator must then be able to control the attitude of the rotor
system and the direction of the total thrust vector.

CL 1/2p S
V2
(1) In the lift equation, L =
(Equation 2.3), an
increase in wind velocity causes the advancing blade to flap up; decrease
a

in wind velocity causes the retreating blade to flap down. Surface area or
air density does not change; however, the CL can change by changing the angle
of attack. The advancing blade flaps up, reaching maximum upflap displace-
ment over the nose of the helicopter.

(2) In forward flight, the lowest point of the rotor disk must be
over the nose. The advancing blade needs to flap down instead of up. Because
CL is based on angle of attack, the blade will lose lift and flap down if
angle of attack is reduced at the
3 o'clock position. Because of phase lag,
this downflap reaches maximum displacement over the nose of the helicopter.
For the retreating blade to flap up, the angle of attack at the 9 o'clock
position is increased. Phase lag causes maximum upflap blade displacement
over the tail of the helicopter. This is achieved through cyclic feathering.
Cyclic feathering changes the angle of incidence differentially around the
rotor system. This cyclic st~ck movement decreases the angle of incidence at
one point of the rotor disk and increases the angle of incidence by the same
amount 180 degrees of travel later.
(a) Forward movement of the cyclic reduces the pitch angle and
the angle of attack of the blade at the 3 o'clock position. At the same
time, it increases the pitch angle and the angle of attack of the blade at
the 9 o'clock position.

6-18
(b)decreased angle of attack at the 3 o'~lock position
A

decrEases CL and, therefore, the lift


produced by the bla~e
at this position.
The 1lade responds to this loss of lift
by flapping down. Maximum downflap-
ping velocity takes place where blade pitch and angle of -8attack are lowest,
in tbis instance at the 3 o'clock position. Because of p~ase lag, maximum
down1lapping displacement of the blade occurs over the no ~e of the
helioeopter.

(c) An increased blade pitch and angle of a- ttack at the 9


o'clock position increase the CL and, therefore, the lift- produced by the
blade at this position. The blade responds to this incr~ ase in lift by flap-
ping up. Maximum upflapping velocity takes place where L- :he blade pitch is
highest, in this instance at the 9 o'clock position. Bee:- ause of phase lag,
maxinum upflapping displacement of the blade occurs over the tail of the
helicopter.
(d) Maximum downflapping displacement of
nos~ of the helicopter and
t~e
blade over the
maximum upflapping displaceme~t of the blade over
the tail of the helicopter tilt the rotor disk and, ther~fore, the thrust
vector forward. As the velocity of the helicopter increB88ases, dissymmetry of
lift attempts to cause the advancing blade to flap up anClla the
blade to flap down. To prevent blowback from occurring,
the retreating
pilot must con-
the helicopter
tinually move the cyclic stick forward as the velocity
increases. Figure 6-17 shows the changes in pitch angle
o~
as the cyclic is
moved forward at increased airspeeds. At a hover, the c~clic is centered and
the pitch angles on the advancing and retreating blades e==are the same. At low
forward speeds, moving the cyclic forward reduces the pit===ch angle on the
advancing blade and increases the pitch angle on the ret~eating blade. This
causes slight rotor tilt. At higher forward speeds, the aviator continuas to
move the cyclic forward. This further reduces the pitch angle on the advanc-
ing blade and further increases the pitch angle on the re===treating blade. As
a
result, there is even more tilt to the rotor than at lc==Jwer forward speeds.

-J5'~
RB

~ 100
50

AB AB
HOVER
EQUAL PITCH ANGLES ON LOW FORWARD SPEED HIGH FORWARD SPEED
ADVANCING AND RETREATING
BLADES. CYCLIC CENTERED. REDUCED PITCH ANGLE ON R~DUCED PITCH ANGLE ON
ADVANCING BLADE; INCREASED A~VANCING BLADE; INCREASED
PITCH ANGLE ON RETREATING PI---CH ANGLE ON RETREATING
BLADE; SLIGHT TILT TO ROTOR. B~ADE; INCREASED TILT TO
CYCLIC SLIGHTLY FORWARD. RC::=:=>TOR. CYCLIC FURTHER
F~RWARD.

Figure 6-17. Blade pitch angles

6-19
6-10. TANDEM-ROTOR HELICOPTER DISSYMMETRY OF LIFT

Dissymmetry of lift
also applies to tandem-rotor helicopters. The most
important difference between single-rotor and tandem-rotor helicopters is
that the aviator cannot manually compensate for dissymmetry of lift by apply-
ing cyclic on tandem-rotor helicopters. Automatic cyclic-feathering systems
are installed on tandem-rotor helicopters. These systems are activated
through computer-generated commands at specified airspeeds, usually around 70
knots. This allows increased airspeeds and level fuselage attitudes at
higher airspeeds. At low airspeeds, blade flapping can compensate for dis-
symmetry of lift. Cyclic feathering is induced at different positions on the
forward and aft rotor systems because of the counterrotation of the rotor
systems. If the cyclic-feathering system fails to properly feather the
blades at higher airspeeds, greater blade-flapping angles and nose-low flight
attitudes induce bending forces on the rotor-driving mechanisms.

6'-11. TAIL ROTOR DISSYMMETRY OF LIFT

The tail rotor experiences 'dissymmetry of lift during forward flight because
it also has advancing and retreating blades. Dissymmetry is corrected by a
flapping action.

a. Hinged Tail Rotors. Two basic types of flapping hinges--the delta


hinge and the offset hinge--are used on most helicopters.

(1) Delta hinge. The delta hinge, as shown in Figure 6-18, is not
oriented parallel to the blade chord. It is designed so flapping automati-
cally causes the blades to feather, which corrects for dissymmetry of lift.

DELTA~HINGE

I : : {
r-
.'ø'o ---"

Figure 6-18. Delta hinge

6-20
(2)
Offset hinge. The offset hinge, or plain flapping
located outboard from the hub. This hinge, shown in hinge, is
Figure 6-19, uses cen-
trifugal force to produce substantial forces that act on ,the hub. One impor-
tant advantage of the offset hinge is that control exists
regardless of lift;
centrifugal force is independent of lift.

FLAPPING HINGE OFFSET FROM


PLAIN FLAPPING HINGE CENTER PRODUCES MOMENTS
WITHOUT DISK TILT

.(
:..$ ".: j
!

Figure 6-19. Offset hinge


b. Hingeless Tail Rotors. Some helicopters. such as the UH-60.
have
hingeless tail rotors. Twisting and bending of the blade
spar combine to
accomplish flapping and feathering of these
tail rotors.
6-12. TRANSVERSE FLOW EFFECT

As the forward velocity of the helicopter


increases. another phenomenon of
differential airflow in the rotor system occurs. While dissymmetry of
involves the advancing and retreating sides of the lift
rotor disk. transverse
flow effect involves the front and rear halves of the
rotor disk. Because of
coning and the forward tilt of the rotor system,
there is a differential
airflow across the front and rear halves of the rotor disk. This
is the
transverse flow effect. The pilot can recognize the transverse flow
(Figure 6-20) because of increased vibrations of the effect
helicopter at airspeeds
just below ETL on takeoff and after passing through ETL during
landing.
vibrations take place because the greatest lift differential between the These
and aft portions of the rotor system fore
occurs at those airspeeds. The vibra-
tions are caused by increased induced drag on the blades as they
pass over
the tail of the helicopter.

a. Causes of Transverse Flow Effect. Air moving across the rotor disk
in forward flight is deflected downward because of induced
the distance air must flow over the rotor disk. the flow. The greater
longer the disk has to
act on it and the greater the deflection. This
results in a more horizontal
flow of air over the forward half of the rotor disk
than over the rear half;
thus there is less induced flow over the
front half of the rotor disk than
over the rear half.

6-21
of the disk
(1) Greater induced flow velocity through the rear half
Because the
decreases the angle of attack, as shown in part A, Figure 6-20.
blades are operating at the same angle of incidence over the nose
and tail,
the nose causes an increase in the angle of
reducing the induced flow over
attack, as shown in part B, Figure 6-20.
(2) An increased angle of attack in the front half of the rotor disk
increases the lift of the blade at this location. Increased lift
on the
blade over the nose causes the blade to flap up. Because of phase lag, the
maximum upf1apping blade displacement occurs over
the left side of the heli-
by the blade over the tail, combined
copter. The decreased lift produced
in maximum downf1apping blade displacement over the
with phase lag, results
the disk to the
right side of the helicopter. The displacement tilts rotor
the
This change in at-
right, changing the direction of the thrust vector. is to maintain
titude of the rotor disk must be prevented if the helicopter
a straight flight path.
A cyclic input
b. Compensating for Transverse Flow Effect. left
and angle of attack of the blade over the nose
decreases the pitch angle
while increasing the pitch angle and angle of attack of the blade over the
changes to lift. As the
tail. These changes to blade angles of attack cause
pilot senses the right tilt of the rotor, he Asmust apply speedleft cyclic to pre-
vent a change in the attitude of the disk. forward increases, the
the fore and aft portions also increases.
potential lift differential between
prevent the right of the tilt
Additional left cyclic inputs are required to
rotor as a result of transverse flow effect.
At higher airspeeds, lift dif-
of the disk begins to decrease.
ferential between the fore and aft portions
The cyclic stick must be moved back to the right at higher cruise speeds.

ANGLE OF RESULTANT
ANGLE OF
RESULTANT
RELATIVE WIND ATTACK RELATIVE WIND ATTACK

Ð~;~~ ï-'N~~C::D

ROTATIONAL RELATIVE WIND


Q. ,
~-'\J;::::UCED
.. ~
ROTATIONAL RELATIVE WIND
FLOW

DOWNWARD VELOCITY OF AIR MOLECULES USED


BY AFT SECTION OF ROTOR

MORE HORIZONTAL
FLOW OF AIR

Figure 6-20. Transverse flow effect

6-22
SECTION IV

HOVERING

FORCES AT A HOVER
6-13.

a. Hovering is when helicopter maintains a constant position over a


a

selected point, usually fewa


feet above the ground. For a helicopter to
hover, the lift
produced by the rotor system must equal the total weight of
the helicopter. Forces of lift
and weight reach a state of balance during
the stationary hover. With the blades rotating at high velocity, an increase
of blade pitch (angle of attack) would induce the necessary lift
for a hover.

b. Hovering is actually an element of vertical flight. In a no-wind


condition, the tip-path plane of the blades remains horizontal. Increasing
the angle of attack of the blades while their velocity remains constant
generates additional vertical thrust. By the same principle, the reverse is
true; decreased pitch causes the helicopter to descend. Therefore, by up-
setting the vertical balance of forces, the helicopter will either climb or
descend vertically.
6-14. AIRFLOW DURING HOVERING

At a hover, the blade-tip vortex (air swirl at the tip of the rotor blades)
reduces the effectiveness of the outer blade portions (Figure 6-21). The
vortex of a preceding blade also severely affects the lift of the following
blades. If the vortex made by one passing blade remains swirling for a num-
ber of seconds, then two blades operating at 350 RPM create 700 long-lasting
vortex patterns per minute. The continuous creation of new vortexes and the
continuous ingestion of existing vortexes are the primary causes of the high
power required to hover.

BLADE-TIP
VORTEX

).

Figure 6-21. Blade-tip vortex at a hover

6-23
GROUND EFFECT
6-15.

a. Ground condition of improved performance encountered when


effect is a

the aircraft is operating near the ground. It is due to interference of the


surface with the airflow pattern of the rotor system and is more pronounced
closer to the ground. The high power requirement needed to hover aGE, shown
in Figure 6-22, is reduced when the aircraft is operating in-ground effect,
shown in Figure 6-23.

helicopter hovering out-of-ground effect (Figure 6-22) requires


A a
b.
great deal of power and high blade pitch angles to move
a tremendous quantity
of air at high velocities down through the rotor system (see paragraph 6-5)
to produce lift equal to weight. The velocity of this downwash continues to
increase until it reaches maximum velocity, about one rotor disk below the
plane of rotation. This results in a large induced-flow velocity. The re-
sultant relative wind produces a lift vector that is inclined well to the
rear.

INDUCED LARGE
FLOW VEL = BLADE-TIP
60 It! see VORTEXES

LIFT VECTOR INCLINED


WELL TO ,REAR
\..

MAX VEL =

120lUsee INDUCED
FLOW VEL =
60 fUsee

I
I LARGE BLADE PITCH ANGLE
I
AXIS
OF ROTATION

Figure 6-22. Aerodynamic forces, out-of-ground effect (example)

6-24
REDUCED BLADE-TIP
VORTEXES

MORE VERTICAL UFT VECTOR

\..
r ANGLE Of ATTACK IS THE
I
SAME IN- OR OUT-Of-GROUND EffECT
: ~
I
I
INDUCED fLOW
PITCH ANGLE 140 VEL = 45 ft/sec

I
I ÇUCED
BLADE PITCH ANGLE
AXIS
Of ROTATION

Figure 6-23. Aerodynamic forces, in-ground effect (example)


c. Increased blade efficiency during IGE is due to two phenomena.

(1) Reduction of the velocity of the induced


airflow. Because the
ground interrupts the airflow under the helicopter, the
entire flow is
altered. This reduces downward velocity of the induced flow. The result is
less induced drag and a more vertical lift
vector. The
tain a hover IGE can be produced with less power because the
lift
needed to sus-

shown in Figure 6-23, is more vertical.


as lift vector,

(2) Reduction of the rotor-tip


vortex. This is illustrated in
Figure 6-23. When the helicopter is IGE, the downward and outward
airflow
pattern tends to restrict vortex generation. Thus the outboard portion of
the rotor blade becomes more efficient, reducing overall system
turbulence
caused by ingestion and recirculation of the vortex
swirls.
d. To move from an OGE hover (Figure 6-24) to an IGE
hover
(Figure 6-25), the aviator must first reduce the total
system by reducing collective pitch. As the
of the lift rotor
helicopter moves nearer the
surface, the ground interrupts downward flow; the air then moves horizontally
outward from under the rotor disk. This prevents high induced-flow

6-25
velocities from developing below the rotor and is manifested as
a reduced
induced-flow velocity through the rotor disk. Because of the decreased blade
pitch angle and reduced induced-flow velocities, the lift vector is not
tilted as far to the rear. Less power is required to produce the same amount
of lift. Also, the horizontal flow of air nearer the plane of rotation tends
to restrict the formation of blade-tip vortexes.

. ;;~
\\q/JII;
\ t \I \ ~I ~ /
,?
~

LIFT VECTOR
INCLINED TO REAR
I~~'
\ \:7;
"-

Ij\\\ /J Ii \
~!) )
\
.......A
LARGE ROTOR-TIP VORTEX

AX'S
OF ROTATION

LARGE INDUCED
DRAG ANGLE HIGH BLADE PITCH ANGLE

LIFT IS
PERPENDICULAR
TO RESULTANT RWI

LARGE
INDUCED FLOW
VELOCITY

ROTATIONAL RW

ALTITUDE GREATER
THAN ONE ROTOR DIAMETER

Figure 6-24. Out-of-ground-effect hover

6-26
~~\\!/I//
~ \\\\VIII~~ REDUCED

~
ROTOR-TIP
VORTEX

~
AXIS
OF ROTATION
.

LIFT VECTOR MORE VERTICAL


LIFT
REDUCED INDUCED
DRAG ANGLE

DRAG ANGLE OF
ATTACK RESULTANTRW
-

~ REDUCED INDUCED
ROTATIONAL RW FLOW VELOCITY

ALTITUDE LESS THAN


ONE ROTOR DIAMETER

Figure 6-25. 1n-ground-effect hover

e. Rotor efficiency is increased by ground effect to a height of about


one rotor diameter for most helicopters. Figure 6-26 shows the increased
lift possible with 1GE. This figure illustrates the percent of increase in
rotor thrust at various rotor heights. At a
rotor height of one-half rotor
diameter, thrust increases about 7
percent. At rotor heights above one rotor
diameter, thrust increase is small, decreasing to zero at a height of about
1 1/4
rotor diameter.

f. effect occurs when aircraft hover over smooth, paved


Maximum ground
surfaces. Over tall grass, rough terrain, revetments, or water, ground ef-
fect may be seriously reduced. This phenomenon is due to the partial break-
down and cancellation of ground effect and the return of large vortex
patterns with increased downwash velocities.
OGE descends into a ground-effect hover,
g. If a helicopter hovering
blade efficiency increases because of the more favorable induced flow. As
efficiency of the rotor system increases, the aviator reduces the blade pitch
angle to remain in the ground-effect hover. Hovering in-ground effect re-
quires less power than hovering out-of-ground effect.

6-27
~ 20
W
/IJ
<I:
15
w~
a::/IJ
0:)
Za::
10

-J:~
~ 5
Za::
Wo
0 0
~
ffio 114 112 3/4 1 1 114
11. a::

ROTOR HEIGHT ABOVE GROUND IN ROTOR DIAMETERS

Figure 6-26. Increased lift capability, in-ground effect

SECTION V

AIRFLOW PATTERNS IN FORWARD FLIGHT

6-16. TRANSLATIONAL FLIGHT

Each knot of incoming wind gained by horizontal movement or surface wind


improves the efficiency of the rotor system. As the inco~ing wind enters the
rotor system, turbulence and vortexes are left behind and the flow of air
becomes more horizontal. Therefore, aircraft performance improves. Improved
rotor efficiency resulting from directional flight is called translational
lift. When a single-rotor helicopter makes the transition from hover to
forward flight, the tail rotor becomes more aerodynamically efficient. The
tail rotor works in progressively less turbulent air as speed increases. As
tail-rotor efficiency improves, more thrust is produced. This causes the
aircraft nose to yaw left if the main rotor turns counterclockwise. DUring a
takeoff, where power is constant, the aviator must apply right pedal as speed
increases to correct for, the left-yaw tendency.

a. Airflow Pattern, 1 to 5 Knots. An airflow pattern for a forward


speed of 1 to 5 knots is shown in Figure 6-27: In this figure, the downwind
vortex is beginning to dissipate; induced flow down through the rear of the
rotor disk is more horizontal than at the hover shown in Figure 6-21.

b. Airflow Pattern, 6 to 15 Knots. Figure 6-28 shows the airflow


pattern at a speed of 6 to 15 knots. Airflow is much more horizontal than at
the hover shown in Figure 6-21. The forward edge of the downwash pattern
moves closer to the helicopter as airspeed increases. At 10 knots, the for-
ward edge of the downwash pattern has moved to a
position under the blade
tips over the nose. At 15 knots, this downwash pattern moves to a position
toward the tail of the aircraft. This allows the flow of air through the
rotor system to become more horizontal and streamlined as airspeed increases.
NOTE: As the helicopter makes the transition from a hover to directional
flight, dissynnnetry of lift and the transverse flow effect begin to affect

6-28
the rotor system. These subjects are discussed in paragraphs 6-8 through
6-12.

DOWNWIND EDGE UPWIND EDGE OF DOWNWASH PATTERN

1 5 knots
Figure 6-27. Translational lift at to

6 15 knots
Figure 6-28. Translational lift at to

6-17. EFFECTIVE TRANSLATIONAL LIFT


RPM
16 to 24 knots (depending on the size, blade area, and
a. At about of old
of the rotor system), the rotor completely outruns the recirculation
as shown in
vortexes and begins to work in relatively undisturbed air,
Figure 6-29. The rotor no longer pumps the air in a
circular pattern but
of air passing through the
continually flies into undisturbed air. The flow flow is
Induced
rotor system is more horizontal, depending on forward speed. the angle of
reduced; therefore, induced drag is reduced. This increases
attack.

6-29
NO RECIRCULATION MORE AIR PER SECOND
OF AIR FOR ROTOR TO WORK ON

MORE HORIZONTAL
FLOW OF AIR
- -

------
TAIL ROTOR OPERATES IN REDUCED
RELATIVELY CLEAN AIR INDUCED FLOW
INCREASES
ANGLE OF ATTACK

Figure 6-29. Effective translational lift


b. As single-rotor aircraft speed increases, translational
lift becomes
more effective; the nose rises or pitches up and rolls to the right. This
tendency is caused by the combined effects of dissymmetry of lift and the
transverse flow effect. Dissymmetry of lift
causes blowback of the rotor
system; the transverse flow effect tilts the rotor to the right. Aviators
must correct for this tendency to maintain a constant rotor-disk attitude
that moves the helicopter through the speed range where ETL occurs.

SECTION VI

COMPRESSIBILITY EFFECTS

6-18. COMPRESSIBLE AND INCOMPRESSIBLE FLOW

At low airspeeds, air is incompressible. Incompressible airflow is similar


to the flow of water, hydraulic fluid, or any other incompressible fluid. At
low speeds, air experiences relatively small changes in pressure with little
change in density. However, at high speeds, greater pressure changes occur,
causing compression of the air, which results in significant changes to air
density. This compressible flow will occur when there is a transonic or
supersonic flow of air across the airfoil. Because helicopters are being
flown at increasingly higher speeds, aviators must learn more about coping
with the effects of compressible flow.

The major factor in high-speed airflow is the speed of sound. Speed


a.
of sound is the rate at which small pressure disturbances are propagated
through the air. This propagation speed is solely a function of air tempera-
ture. Figure 6-30 shows the variation of speed of sound with temperature at
various altitudes in the standard atmosphere.

6-30
r
ALTITUDE TEMPERATURE
SPEED of'
SOUND

FEET of oc KNOTS

SEA LEVEL. 59.0 15.0 661.7


..
. . . . . .

5,000 41.2 5.1 650.3


.........,... 23.3 -
4.8 638.6
10,000..... ........
15.000 5.5 -
14.7 626.7
..
. . . . . . . .
. . .

20,000 . . . . . . . . . . . . .
-12.3 -
24.6 614.6

25,000 -30.2 -
34.5 602.2
.,
. . . . . . .
. . . .

30,000 -48.0
-
44.4 589.6
..
. . . . . . . . .
. .

35,000 -65.8 -
54.3 576.6
..
. . . . . . . .
. . .

40,000 -697 -
56.5 573.8
..
. . . . . . . . .
. .

50,000 -69.7 -
56.5 573.8
..
. . . . . . .
. . . .

60,000 -
69.7 -
565 573.8
..
. . . . . . .
. . . .

~
\..

Figure 6-30. Variation of temperature and speed of


sound with altitude

b. the speed of sound, the airflow


If the airfoil is traveling. above
ahead of pressure field because pressure
it will not be influenced by the
disturbances cannot be propagated ahead of the airfoil. Therefore, as speed
nears the speed of sound, a compression wave forms at the leading edge; all
changes in velocity and pressure take place sharply and suddenly. The air-
flow ahead of the airfoil is not influenced until the air particles are
suddenly forced out of the way by the concentrated pressure wave set up by
the airfoil. A typical supersonic airflow is shown in Figure 6-31.

c. As an airfoil moves through the air, velocity and pressure changes


create pressure changes in the airflow surrounding the airfoil. These pres-
sure changes are propagated through the air at the speed of sound. If the
airfoil is traveling at low speed, the pressure disturbances are propagated
ahead of it. The airflow immediately ahead of the airfoil is influenced by
the pressure field on the airfoil. Figure 6-31 shows how this pressure field
influences the flow ahead of an airfoil traveling at low speed. This figure
compares subsonic and supersonic flow patterns. At high speeds, no change of
flow direction is apparent ahead of the leading edge.

6-31
SUBSONIC FLOW PATTERN SUPERSONIC FLOW PATTERN

D+
FLOW DIRECTION CHANGES V-
WELL AHEAD OF LEADING EDGE P+

)
V+ )
~

--- ~

::;;il1l!!!!fljfl///////;l~ ---
P-

-~ -
~

~~\~ It

>--
~ \\\\ ~
" - )

-= ~ '\.
-
\\\\
\\\\
" \\\\
PRESSURE WAVES PROPAGATED NO CHANGE OF FLOW DIRECTION
AHEAD OF AIRFOIL APPARENT AHEAD OF LEADING EDGE

Figure 6-31. Comparison of subsonic and supersonic flow patterns

d. All compressibility effects depend on the relationship of blade speed


to the speed of sound. The term used to describe this relationship is the
Mach number. This is the relationship of the true airspeed to the speed of
sound.

e. Compressibility effects are not limited to speeds at and above the


speed of sound. The aerodynamic shape of an airfoil causes local flow ve-
locities greater than the blade speed. Thus a blade can experience compress-
ibility effects at speeds well below the speed of sound. Because both
subsonic and supersonic flows can exist on a blade, the following regimes of
flight need to be defined:
.

0
Subsonic--Mach numbers below 0.75.
0

Transonic--Mach numbers from 0.75 to 1.20.


0
Supersonic--Mach numbers from 1.20 to 5.00.

The Mach numbers used to define these regimes of flight are approximate.
However, they are useful for comparing the types of flow existing in each
area. In the subsonic regime, pure subsonic airflow probably exists on all
parts of the blade. In the transonic regime, the transition range between
subsonic and supersonic, flow on the blades is probably partially subsonic
and partially supersonic.

f. Differences between subsonic and supersonic flow are due to the com-
pressibility of the supersonic flow. Figure 6-32 compares incompressible
and compressible flow through a closed tube. In this example, the mass flow
along the tube is constant.

6-32
INCOMPRESSIBLE
(SUBSONIC)
//1/111/ I1111
III 11/ III ~ r,............
--1'1 1 / / 1 1 1 1 1 1 / / / I 1 / /

~ 11/ I 1 1
III 1 / 1
~ I~
---
-
- ~

----+----...

"".
.
~
..
~

/11 1 1 1 1
.
Þ'- .

III 1 III 1 1 1 1
)>-
f;I../ ~-
..-~-
--.
--"'-
/ / 1 1 1 1 1 / I / I t-.- rr!:Q í I
~

1 1 1 1 1 1 1 1 1 1 1 1 1

CONVERGING
.. DIVERGING
. INCREASING VELOCITY . DECREASING VELOCITY
..
. DECREASING PRESSURE . INCREASING PRESSURE
. CONSTANT DENSITY . CONSTANT DENSITY

..
CONVERGING DIVERGING

. DECREASING VELOCITY . INCREASING VELOCITY


..
. INCREASING PRESSURE . DECREASING PRESSURE
. INCREASING DENSITY . DECREASING DENSITY

Figure 6-32. Comparison of compressible and incompressible flow


through a closed tube

(1)Subsonic incompressible flow. The example of subsonic incom-


pressible flow is simplified because density of the flow is constant
throughout the tube. As the flow approaches a constriction and the stream-
.lines converge, velocity increases as static pressure decreases. That is, a
convergence of the tube requires an increasing velocity to accommodate the
continuity of flow. Also, as the subsonic incompressible flow enters a di-
verging section of the tube, velocity decreases and static pressure in-
creases; density remains unchanged. The behavior of subsonic incompressible
flow is that a convergence causes expansion (decreasing pressure); a diver-
gence causes compression (increasing pressure).

6-33
(2)
Supersonic compressible flow. The example of supersonic
compressible flow is complicated because the variations of flow density are
related to the changes in velocity and static pressure. The behavior of
supersonic compressible flow is that a convergence causes compression; a
divergence causes expansion. Therefore, as the supersonic compressible flow
approaches a constriction and the streamlines converge, velocity decreases
and static pressure increases. Continuity of mass flow is maintained by the
increase in flow density that accompanies the decrease in velocity. As the
supersonic compressible flow enters a
diverging section of the tube, velocity
increases and static pressure decreases; density decreases to accommodate the
condition of continuity.

g. Three significant differences emerge from comparing supersonic


compressible and subsonic incompressible flow.

0) Compressible flow includes the additional variable of flow


density.

(2) Convergence of flow causes expansion of incompressible flow but


compression of compressible flow.

(3)
Divergence of flow causes compression of incompressible flow but
expansion of compressible flow.

6-19. TRANSONIC FLOW PATTERNS

a. In subsonic flight, an airfoil that is producing lift has local


velocities on the surface greater than the free-stream velocity. Compress-
ibility effects can then be expected to occur at flight speeds less than the
speed of sound. Mixed subsonic and supersonic flow may be encountered in the
transonic regime of flight. The first significant effects of compressibility
occur in this regime. Compressibility effects on the helicopter increase the
power required to maintain rotor RPM and cause rotor roughness, vibration,
cyclic shake, and an undesirable structural twisting of the blade.

b. the conventional airfoil shown in Figure 6-33 is at a flight Mach


If
number of 0.50 and a slight positive angle of attack, the maximum local
velocity on the surface will be greater than the blade speed but most likely
less than sonic speed. If an increase in blade Mach number to 0.72 is as-
sumed to produce the first evidence of local sonic flow, this condition of
flight is the highest blade speed possible without supersonic flow. This is
the critical Mach number. Therefore, shock waves, buffeting, and airflow
separation all take place above the critical Mach number.

c. the critical Mach number is exceeded, an area of supersonic


As
airflow is created. A normal shock wave then forms the boundary between the
supersonic and subsonic flow on the aft portion of the airfoil surface. The
acceleration of the airflow from subsonic to supersonic is smooth and without
shock waves if the surface is smooth and the transition gradual. However,

6-34
the transition of airflow from supersonic to subsonic is always accompanied
by a shock wave. When the airflow direction does not change, the wave formed
is a
normal shock wave.

M =
0.50""fUHj~ MAXIMUM LOCAL VELOCITY
IS LESS THAN SONIC

MAXIMUM LOCAL VELOCITY

r EQUAL TO SONIC

---- -----M;o~ii--- ---


-ØJ!}~~
-

(CRITICAL MACH NUMBER)


As air passes through shock wave, air density
Increases, heat Is created, velocity of the air
decreases, static pressure increases, and boundary
layer separation may occur. The aerodynamic
- NORMAL SHOCK WAVE center moves from the 25% chord toward the 50%
chord, and the nose of the airfoil pitches down.

M =
0.77 ~~-'
POSSIBLE SEPARATION

NORMAL
SHOCK WAVE

SEPARATION

M =
0.82 ~:"q. ~'~

M =
0.95 NORMAL SHOCK WAVE

M =
1.05

SUBSONIC
AIRFLOW
BOW WAVE

Figure 6-33. Transonic flow patterns

6-35
d. is detached from the leading edge of the
The normal shock wave
airfoil is
and perpendicular to the upstream flow. The flow immediately
behind the wave is subsonic. Figure 6-34 illustrates how an airfoil at high
subsonic speeds has local flow velocities that are supersonic. As the local
supersonic flow moves aft, a normal shock wave forms, slowing the flow to
subsonic. A supersonic airstream passing through a normal shock wave experi-
ences the following changes.

(1) The airstream is slowed to subsonic. The local Mach number


behind the wave is about equal to the reciprocal of the Mach number ahead of
the wave. For example, if the Mach number ahead of the wave is 1.25, the
.

Mach-number of the flow behindthê -wave isaòo-uCO~80~"

(2) The airflow direction immediately behind the wave is unchanged.


(3) The static pressure of the airstream behind the wave is in-
creased greatly.

(4) The density of the airstream behind the wave is increased


greatly.
(5) of the available energy of the airstream (indicated by the
Some
sum of dynamic and static pressure) is dissipated and turned into unavailable
heat energy. Total pressure behind the shock wave is greatly reduced. The
normal shock wave is very wasteful of energy.

e. The normal shock wave produces a large increase in the static


pressure of the airstream behind the wave. If the shock wave is strong, the
boundary layer may separate. At speeds only slightly beyond the critical
Mach number, the shock wave formed is not strong enough to cause separation
or any noticeable change in aerodynamic forces. However, an increase in
speed above the critical Mach number--enough to form a strong shock wave--can
separate the boundary layer and suddenly change aerodynamic forces. Such a
flow condition is shown in Figure 6-33 by the flow pattern for M 0.77. A =

further increase in the Mach number to 0.82 enlarges the supersonic area on
the upper surface and forms an additional area of supersonic flow and a nor-
mal shock wave on the lower surface.

f. As the blade speed approaches the speed of sound, the areas of


supersonic flow enlarge and the shock waves move nearer the trailing edge
(Figure 6-33; M 0.95). The boundary layer may remain separated or may
=

reattach, depending on airfoil shape and angle of attack. When the blade
speed exceeds the speed of sound, a bow wave forms at the leading edge. This
typical flow pattern is shown in Figure 6-33 by the drawing for M 1.05. If =

speed is increased to some higher supersonic value, all oblique portions of


the waves incline greatly and the detached normal shock portion of the bow
wave moves closer to the leading edge.

6-36
SUBSONIC

----.- .-_u,---..----.----... ----.- ""---------'--_.._._---'-~---_._-_._------------ ---- -

--------.----. --
-

--
---------~.- .--- ---
Figure 6-34. Normal shock wave formation

Airflow separation induced by shock-wave formation can create sig-


g.
nificant variations in aerodynamic force coefficients of the airfoil section.
(1) When the blade speed exceeds the critical Mach number, some
typical effects on an airfoil section are--

lift
"
An increase in the section drag coefficient for a
given section
coefficient.
A
lift
0
decrease in section coefficient for a
given section angle of
attack.
0
Vibrations becoming more severe as velocity increases and more of the
advancing blade is subjected to a supersonic flow of air.
"
A change in section pitching moment coefficient.
As the shock waves move toward the trailing edge of the airfoil, the
aerodynamic center begins to move away from its normal location at the 25
percent chord. By the time the shock wave has reached the trailing edge of
the airfoil, the aerodynamic center has retreated to the 50 percent chord.
This causes the leading edge of the airfoil to be deflected down, which may
result in structural failure of the blade (skin deformation or separation).
(2) A comparison of drag coefficient to the Mach number for a

constant lift coefficient is shown in Figure 6-35. The point of the sharp
increase in the drag coefficient is the force divergence Mach number; it
usually exceeds the critical Mach number by about 5 to 10 percent.

h. Because the speed of the helicopter is added to the speed of rotation


of the advancing blade, the highest relative velocities occur at the tip of
the advancing blade. When the Mach number of the tip section of the advanc-
ing blade exceeds the critical Mach number for the rotor blade section,
compressibility effects will result. The critical Mach number is the free-
stream Mach number that produces the first evidence of local sonic flow. The
principal effects of compressibility are the large increase in drag and the
rearward shift of the airfoil aerodynamic center.

6-37
t
CD FORCE DIVERGENCE
MACH NUMBER
CRITICAL

--.--... __,,___h_._,~ ." --_.-__~._-m'.~"


~~-C-~~COH3NU~~~~_, -
-

-^.,,~_..-,----"--"~-,-- u

'''"-----~----"._-~--'' -...-- -----'--"-'"-"'--


L -
. ,

L=

0.5 0.75 1.0


MACH NUMBER

Figure 6-35. Compressibility drag increase

i. Compressibility effects become more severe at higher lift coeffi-


cients (higher blade angles of attack) and higher Mach numbers. The more
adverse compressibility effects occur with the following operating
conditions:

(1) High airspeed.

(2) High rotor RPM.

(3) High gross weight.

(4) High-density altitude.


(5) Low temperature. The speed of sound is proportional to the
square root of the absolute temperature. Therefore, sonic velocity will be
more easily obtained at low temperatures when the sonic speed is lower.

(6) Turbulent air. Sharp gusts momentarily increase the blade angle
of attack, which lowers the critical Mach number to the point where
compressibility may be encountered on the blade.

Compressibility effects decrease in severity if blade pitch is


j.
decreased. The critical conditions for retreating blade stall, which is
discussed in the following section, and compressibility are similar. One
basic difference must be noted. Compressibility occurs at high RPM while
retreating blade stall occurs at low RPM. With the exception of RPM control,
the recovery technique is identical for both.

6-38
SECTION VII
EMERGENCY SITUATIONS

6-20. RETREATING BLADE STALL

--The-retTeat-ingblade -of-helicopters. tends-to-...stall.in forward_fJigþt, -.


.

- -

(Figure 6-36). Just as the stall of an airplane wing limits the low-speed
possibilities QL .the.ê.i:rp:l~ne t:l-1e_~t_~_~_l ()f a rotor blade limits the
L

high-speed potential of a helicopter. At a normal hover,fhelìrEpattern-is.


as shown in part A, Figure 6-36. The speed of the retreating blade--the
blade moving away from the direction of flight--decreases as forward speed
increases. As shown in part B, Figure 6-36, the smaller area of the retreat-
ing blade, with its high angles of attack, however, must still produce an
amount of lift equal to that of the larger area of the advancing blade, with
its low angles of attack. This normal cruise lift pattern is shown in part
B, Figure 6-36. As the speed of the retreating blade decreases with forward
speed, the blade angle of attack must be increased to equalize lift through-
out the rotor-disk area. If this angle increase is continued, the blade
stalls at some high forward speed. Part C, Figure 6-36, shows the lift pat-
tern at critical airspeed, where retreating blade stall occurs. Tip stall
causes vibration and buffeting at critical airspeeds. If the blade descends,
causing greater angles of attack, the stall spreads inboard; the helicopter
may pitch up and roll left.

As forward airspeed
a. Causes and Effects of Retreating Blade Stall.
increases, the no-lift areas (Figure 6-36) move left of center, covering more
of the retreating blade sectors. This requires more lift at the outer re-
treating blade portions to compensate for the loss of lift of the inboard
retreating sections. In the area of reversed flow, the rotational velocity
of this blade section is slower than the aircraft airspeed. Therefore, the
air flows from the trailing to leading edge of the airfoil. In the negative
stall areà, the rotational velocity of the airfoil is faster than aircraft
airspeed; air flows from the leading to trailing edge of the blade. However,
the resultant relative wind strikes the blade so far above the chord line
that the airfoil is operating with a negative angle of attack. The angle of
attack is above the critical angle; therefore, the blade stalls. In the
negative-lift area, rotational velocity and blade flapping have reduced the
angle of attack below the critical angle. However, the resultant relative
wind is still striking far enough above the airfoil to create negative lift.

(1) The loss of disk area on the retreating half of the rotor disk,
resulting from the increasing size of the no-lift areas, causes the blade to
flap down. This downflapping increases the angle of attack. Maximum flap-
ping velocity will occur at the blade tip. Because of this increase in angle
of attack caused by flapping, coupled with higher blade pitch angles, the
angle of attack at the tip of the retreating blade may exceed the critical
angle of attack.

6-39
BLADE ROOT AREA

" NORMAL HOVERING LIFT PATTERN

THE LIFT OF THIS


SMALL AREA WITH
HIGH ANGLES OF {
ATTACK
NEGATIVE LIFT

..
MUST EQUAL
NEGATIVE STALL

REVERSE FLOW
NO-LIFT
AREAS

..
THE LIFT OF THIS
LARGE AREA WITH
LOW ANGLES OF
ATTACK

Ð NORMAL CRUISE LIFT PATTERN

TIP STALL CAUSES VIBRATION AND


BUFFETING AT CRITICAL AIRSPEEDS.

CORRECTION FOR STALL:


.
. REDUCE POWER.
IF BLADE DESCENDS;
CAUSING GREATER
. REDUCE AIRSPEED.
ANGLES OF ATTACK, . REDUCE SEVERITY
STALL SPREADS
OF MANEUVERING.
INBOARD
. INCREASE RPM
TOWARD UPPER LIMIT.

..
HELICOPTER PITCHES UP
AND ROLLS LEFT
. CHECK PEDAL TRIM.

e LIFT PATTERN AT CRITICAL AIRSPEED


(Retreating blade stall)

Figure 6-36. Development of retreating blade stall

6-40
(2) Figure 6-37 shows a rotor disk that has stalled on the
is 14
retreating side. The stall angle of attack for this rotor system
degrees. Distribution of angle of attack along
the blade is shown at eight
and have less
positions in the rotor disk. Although the blades are twisted
pitch at the tip than at the root, angle of attack is higher at the tip be-
cause of blade flapping.

FORWARD

STALLREGIO~ -t
~
<:-,
120' ,
.....

',',
,
50
/ ,/
50

./
"
' 'J;>~
'
,

'.:' "

80','" 0 ,',~
"'"
"
"

"
80 5;"',
,
"

~
- .....
, / 20 40
,oJ

)
160 140 80 20
, ~, 20

\ ,','00
' ,

0
-'
o'~"
, 40
;
,
,
,
'""
,-
..........
"
"
"
"

#",'80 ',', ,

,','140 60 "',60 , ,
, ,
"

~~J;
"

Figure 6-37. Angle-of-attack distribution during


retreating blade stall

Blade stall for single-rotor and tandem-rotor helicopters is


(3)
discussed below. In operations at high forward speeds, the following
condi-
tions are most likely to produce blade stall in both types:
0
High blade loading (high gross weight).
0
Low rotor RPM.

High-density altitude.
0

0
Steep or abrupt turns.
0
Turbulent air.

6-41
(a)
Single-rotor helicopter. In a single-rotor helicopter upon
entry into blade stall, the first effect is generally a noticeable vibration.
This may be followed by a left roll and a tendency for the nose to pitch up.
The tendency to pitch up and roll left may be relatively insignificant for
helicopters with semirigid rotor systems because of pendular action. If the
cyclic stick is held forward and collective pitch is not reduced or is in-
creased, this condition worsens. The vibration greatly increases; control
can be lost. By being familiar with the conditions that lead to blade stall,
the aviator can recognize and correct them. In single-rotor helicopters,
warnings of approaching retreating blade stall are as follows (consult ap-
propriate operator's manual for specific indications):
0
Abnormal vibration.
0
Pitch up of the nose.
0
Tendency of the helicopter to roll in the direction of the stalled
side.
(b)
Tandem-rotor helicopter. In a tandem-rotor helicopter, the
pitch-up tendency is insignificant. Blade stall will often occur on the aft
rotor first because it operates in the turbulent wake of the forward rotor.
Aviators may experience aft-rotor blade stall when maneuvering at high
speed/high gross weight, especially with an aft CG. Blade stall is indicated
by increased vibration. Vibration increases as the aircraft is flown further
into blade stall. To recover, the aviator decreases thrust, increases rotor
RPM, decreases airspeed, reduces the severity of the maneuver, or uses a
combination of these actions. Applying forward cyclic stick before reducing
thrust may aggravate the stall. Roll-control response remains effective
throughout the stall. Most CH-47 aircraft have a cruise guide indicating
system that warns of blade-stall vibrations on the rotor systems. In
tandem-rotor helicopters, blade stall is indicated by a progressive increase
of one- and three-per-revolution vibrations.

b. Preventing Retreating Blade Stall.

(1)
The aviator should take corrective action immediately when blade
stall is likely and exercise extreme caution when maneuvering. An abrupt
maneuver, such as a steep turn or pull-up, may result in dangerously severe
blade stall. Aircraft control and structural limitations of the helicopter
would be threatened. Blade stall normally occurs at high airspeeds. To
prevent blade stall, the aviator must fly slower than normal when--
0
Density altitude is much higher than standard.

aircraft is flying in a high-drag configuration; for example, if


0
The
it is equipped with floats, external stores, weapons, speakers,
floodlights, or sling loads.
0
Air is turbulent.

6-42
blade stall is suspected, the aviator should follow the
(2) When
These steps will normally
procedures outlined in the operator's manual. .

include--
0
Reducing power.

0
Reducing airspeed.

Reducing the severity of the maneuvering.


0

0
Increasing RPM toward upper limi t.
0
Checking pedal trim.

The corrective action


In severe blade stall, aircraft control may be lost.
is to follow procedures outlined in the operator's manual to shorten the
duration of the stall and regain control.

6-21. SETTLING WITH POWER

which the helicopter


Settling with power is a condition of powered flight in
settles in its own downwash. The condition may also
be referred to as the
Before this
vortex ring state, which is one of the four flow state regions.
is
condition is further explained, a discussion of the flow state regions
to autorota-
necessary. The discussion of flow state regions also applies
tion, covered in Section VIII.
Flow States in Descending Flight. When
a helicopter is descending,
a.
the airflow through the rotor system is different from what it was in level
The more vertical the descent, the greater the
flight at the same speed. recognized as
changes in airflow pattern. Four flow states are generally
the thrusting state, the vortex ring
occurring in descending flight: normal
and the windmill brake state. Figure 6-38
state, the autorotative state, a describe the
represents these flow states on rotor blade. The flow states
with forward airspeed, similar
airflow in vertical flight only. In descents
which they
states may exist on the rotor system but the rates of descent at
occur and the symptoms of their existence observed by the pilot are
different.

6-43
A. NORMAL THRUSTING STATE

~ TIP VORTEX

VELOCITY PROFILE
OF AIR RELATIVE
TO ROTOR
B. VORTEX RING STATE

C. AUTO ROTATIVE STATE

~ ~
D. WINDMILL BRAKE STATE

~ ~
Figure 6-38. Flow states in vertical descents
(1) Normal thrusting state. For hovering and at low rates of
descent, the induced flow generated by the blades exceeds the rate of
descent. The airflow is down with respect to the rotor disk. There is a
difference in induced-flow velocities along the blade span. This results
from varying rotational velocities and angles of attack from the root to the
tip of the blade. The thrust generated by the rotor system is quite steady,
and engine power required to maintain rotor RPM remains constant. Part A,
Figure 6-38, depicts this flow state on a rotor blade.

(2) Vortex ring state. At greater rates of descent, the wind


developed by the vertical velocity of the helicopter is opposing the normal
induced flow developed by the rotor. This upward flow of air cancels the in-
duced flow and, in fact, allows an upward flow of air in that part of the
rotor disk where the upward velocity of flow exceeds the induced flow ve-
locity. This results in an unsteady turbulent flow of air through the rotor
6-44
it will enter the vortex ring state~ as illustrated in Figure 6-41. During
this vortex ring state~ roughness and loss of control occur because of the
turbulent rotational flow on the blades and the unsteady shifting of the flow
along the blade span.

Figure 6-39. Induced-flow velocity during hovering flight

~
Figure 6-40. Induced-flow velocity before vortex ring state

~1~~~
Figure 6-41. Vortex ring state

(3) Figure 6-42 shows the relationship of horizontal speed to


vertical speed for a typical helicopter in a descent. Straight lines
emanating from the upper left corner are lines of constant-descent angle.

6-46
disk and increasing rates of descentt even though additional power is sup-
plied by the engine. Part Bt Figure 6-38, depicts this flow state. This
flow state is discussed in greater detail in ~ below.

(3) Autorotative state. At rates of descent exceeding the vortex


ring state, the autorotative state is achieved. At these higher rates of
descent, the turbulent, unsteady flow begins to smooth out. There is some
rate of descent where no engine power is required to maintain rotor RPM. The
rotor system is extracting enough energy to provide the power requirement
from the air in this state of vertical flight. The autorotative state is the
boundary between conditions where engine power must be delivered to the rotor
to prevent rotor RPM decay and where power must be extracted from the rotor
system to prevent rotor overspeed. Part C, Figure 6-38, depicts the airflow
on a blade in the autorotative state. A more in-depth discussion of
autoro-
tation is found in Section VIII.
(4) Windmill brake state. At very high rates of descent, the air-
flow is almost entirely up through the rotor system. The rotor system is
acting similar to a windmill. It is extracting more energy from the air than
is required for flight. It is not a normal operating state for the rotor
system; some energy must be extracted to prevent a rotor overspeed. This can
usually be accomplished by increasing collective pitch, which adds more drag
to the system. Part D, Figure 6-38, depicts the airflow in the windmill
brake state.

b. Vortex Ring State. Conditions conducive to settling with power are a


vertical or near-vertical descent of at least 300 feet per minute and low
forward speed. The rotor system must also be using some of the available
engine power (from 20 to 100 percent) with insufficient power available to
retard the sink rate. These conditions can occur during downwind approaches,
formation approaches/takeoffs, steep approaches, NOE flight, mask/remask
operations, and hover OGE.
(1) Under the above conditions, the helicopter may descend at a
higher rate, exceeding the normal downward induced-flow rate of the inner
blade sections. result, the airflow of the inner blade sections is
As a

upward relative to the disk. This produces a secondary vortex ring in


addition to the normal tip-vortex system. The secondary vortex ring is
generated about the point on the blade where the airflow changes from up to
down. The result is an unsteady turbulent flow over a large area of the
disk; rotor efficiency is lost even though power is still supplied from the
eng ine .

(2) Figure 6-39 shows the induced flow along the blade span during
hovering flight. Downward velocity is greatest at the blade tip where blade
airspeed is highest. As blade airspeed decreases nearer the disk center,
downward velocity is less. Figure 6-40 shows the induced-airflow velocity
pattern along the blade span during a descent conducive to settling with
power. The descent is so rapid that induced flow at the inner portion of the
blades is upward rather than downward. The upflow caused by the descent has
overcome the downflow produced by blade rotation. If the helicopter descends
under these conditions, with insufficient power to slow or stop the descent,

6-45
Superimposed on this grid are the flow state regions for the typical
helicopter. From this illustration, several conclusions regarding the vortex
ring state can be drawn. The discussion on autorotation in Section VIII also
refers to this figure.
(a) The vortex ring state can be completely avoided by
descending on flight paths shallower than about 30 degrees (at any speed).

(b) For steeper approaches, the vortex ring state can be


avoided by using rates of descent versus horizontal velocity either faster or
slower than those passing through the area of severe turbulence and thrust
variation.

NORMAL THRUSTING STATE

LIGHT TURBULENCE AND THRUST VARIATION

}VORTEX RING STATE

SEVERE TURBULENCE AND THRUST V ARIATlDN


I
AUTOROT A TIVE
BOUNDARY
I-
Z
w
U
en
~ 1.0
LO.
0
w
l-
e[
a::

300

400

2.0
~
0 1.0 2.0 3.0
HORIZONTAL SPEED

Figure 6-42. Conditions encountered in settling with


power during a descending forward flight

At very shallòw angles of descent, the vortex ring wake is


(c)
dispersed behind the helicopter. Forward airspeed coupled with induced-flow
velocity prevents the upflow from materializing on the rotor system.
(d)
At steep angles, the vortex ring wake is below the
helicopter at slow rates of descent and above the helicopter at high rates of
descent. Low rates of descent prevent the upflow from exceeding the induced-
.
flow velocities. High rates of.descent result in autorotation or the wind-
mill brake state.

6-47
(4) Power-settling is an unstable condition. If allowed to
continue, the sink rate reaches sufficient proportions for the flow to be up
through a large portion of the rotor system. The rate of descent can reach
extremely high rates. If a large amount of excess power is applied, recovery
can begin during the early stages of power-settling. This excess power may
be enough to overcome the upflow near the center of the rotor. If the sink
rate reaches a
higher rate, power will not be available to break this upflow
and alter the vortex ring state of flow.

(5)
Aviators tend to recover from a descent by applying collective
pitch If not enough power is available for recovery, applying
and power.
collective pitch may aggravate power-settling. This results in more turbu-
lence and a higher rate of descent. The.aviator can recover by increasing
airspeed and lowering collective pitch. Normally, increasing airspeed is the
preferred method of recovery. Usually less altitude is lost by this method
than by the method of lowering collective pitch. The two methods may be
combined if altitude permits.

(6) In tandem-rotor helicopters, recovery should be attempted using


lateral cyclic/pedal inputs to make the transition to directional flight.
Fore and aft cyclic inputs may aggravate the situation.

6-22. RESONANCE

Certain helicopter designs are subject to sympathetic and ground resonance.

a. Sympathetic Resonance. Sympathetic resonance is the harmonic beat


between the main- and tail-rotor systems and other components or assemblies,
which might damage the helicopter. This type of resonance is not usually a
problem because most helicopters have been designed so that the main and tail
géarboxes are in odd decimal ratios. The beat of one component (assembly)
cannot--under normal conditions--harmonize with the beat of another compo-
nent; sympathetic resonance is thus not of immediate concern to the aviator.
However, when resonance ranges are not designed out, the helicopter tachome-
ter is appropriately marked and the resonance range must be avoided.
b. Ground Resonance. Ground resonance may develop in helicopters having
fully articulated rotor systems when a series of shocks causes the rotor
blades in the system to become positioned in unbalanced displacement. If
this oscillating condition progresses, it can be self-energizing and
extremely dangerous. Structural failure usually results. Ground resonance
is most common to three-bladed helicopters with landing wheels. The rotor
blades in a three-bladed helicopter are equally spaced (120 degrees) but are
constructed to allow some horizontal lead and lag action. Ground resonance
occurs when the helicopter contacts the ground during landing or takeoff
(Figure 6-43). If
one wheel of the helicopter strikes the ground ahead of
the others, a shock is transmitted through the fuselage to the rotor.
Another shock is transmitted when the next wheel hits. The first shock from
ground contact, shown in part A, Figure 6-43, causes the blades straddling
the contact point to jolt out of angular balance. If repeated by the next
contact, shown in part B, Figure 6-43, a resonance is established; this sets
up a self-energizing oscillation of the fuselage. The oscillation severity

6-48
increases rapidly and the helicopter may disintegrate unless one of the
following immediate corrections is made.

(1) Take off to a hover if the rotor RPM is in the normal range. A

change of rotor RPM may also aid in breaking the oscillation.

(2) Reduce power ifthe rotor RPM is below the normal range. Use of
a
rotor brake, if installed, may also aid in breaking the oscillation.

A 8

WHEEL HITS THEN OTHER


GROUND WHEEL HITS

Figure 6-43. Ground shock causing blade unbalance

6-49
6-23. DANGEROUS COMBINATIONS OF HEIGHT AND VELOCITY

Combinations of altitude and airspeed define the caution and avoid areas of
the height-velocity diagram (Figure 6-44). Some aviation missions include
portions with elements of risk. Aircraft should not be operated in the
caution or avoid areas unless training or a tactical mission requires it.
During terrain flight, aircraft may have to operate in the avoid areas of the
height-velocity diagram. Because of this, aviators need to be familiar with
the height-velocity caution and avoid areas and with emergency procedures.
Each helicopter has its own height-velocity diagram. The diagram depends on
gross weight, pressure altitude, ambient temperature, velocity, engine power
available, number of engines operating, and rotor speed of the particular
aircraft. Typically, data in the aircraft operator's manual is presented
graphically for standard temperature conditions at sea level and at design
gross weight. Scaling factors for other conditions may be noted in the op-
erator's manual and should be observed. The diagrams in Figure 6-44 are
plotted for a steady-state constant airspeed and altitude. Therefore, they
do not apply to climbing flight.
a. Engine failure occurring while climbing through the avoid or caution
area of the height-velocity combinations will usually damage the helicopter.
During a climb, the helicopter is operating at higher power settings and
blade angles of attack. An engine failure causes rapid rotor RPM decay be-
cause the helicopter stops going upward; then it begins its descent to drive
the rotor and stabilize and increase the RPM to its normal range. The rate
of descent must reach a value that is normal for the airspeed at that par-
ticular moment. Because altitude is not adequate for this sequence, the
helicopter will experience decaying RPM, increasing sink rate, no decelera-
tion lift, little translational lift, and little response to collective pitch
application to cushion touchdown. Airspeed-altitude combinations to be
avoided in case of engine failure are shown in part A, Figure 6-44. Figure
6-44 is based on sea-level operation over a hard landing surface. Higher
density altitudes or unfavorable landing surfaces create additional
autorotational landing difficulties and increased hazards, particularly in
the caution and avoid areas of the chart.

(1) Safe area. The safe area of the chart represents the region
from which safe autorotational landings can be performed with average pilot
alertness, skill, and reaction time. Emergency procedures provided in Chap-
ter 9 of the operator's manual apply.
(2)Caution area. The caution area of the chart represents the
region in which constant pilot alertness and rapid reactions are needed to
accomplish a safe autorotational landing. Aviators can operate in the cau-
tion area of part A, Figure 6-44, when they are over open, level terrain or
runways where obstacle evasion or direction change is not required or where a

short ground run is possible. Emergency procedures in Chapter 9 of the op-


erator's manual apply.
(3) Avoid areas. The avoid areas of the chart represent hazardous
airspeed-altitude combinations from which an autorotational landing would be
extremely difficult without incurring some degree of aircraft damage or

6-50
occupant injury. Operation within the avoid regions is not prohibited but
should be undertaken only if necessary and with a full understanding of the
9 of the operator's manual
risks involved. Emergency procedures in Chapter
apply.

(a)Avoid area (A). Operations in the avoid area (A) of part


A, Figure 6-44, are much less dangerous during descending flight
through any
combination if a landing site is available. The
included height-velocity
avoid area (A) of part A in Figure 6-44 is determined by flight tests and
from engineering data. Several factors are considered. One factor is the
rate of descent required to drive the rotor in autorotation for each 10-knot
increment of airspeed (from
0 through redline or top speed) for the specific
A second factor includes
helicopter configuration (see part B, Figure 6-44). RPM decay rate from the mo-
the rotor inertial characteristics or the rotor
ment of engine failure or until engine failure cues become available to the
added after the cues become available. Avoid
aviator. Reaction time must be
area (A) is also based on the rotor
RPM decay rate experienced while enough
vertical descent is achieved to drive the rotor. The third factor includes
the translational lift values and sink rates for each height-velocity condi-
RPM and pitch-pull energy then available for
tion, with the resulting rotor
cushioning ground impact. The fourth factor includes design stress limita-
tions of the landing gear and hard-landing damage risk to other components.
(b)Avoid area (B). Avoid area (B) of part A, Figure 6-44,
indicates restrictions against continuous operations in certain low-altitude,
based on the
airspeed, and terrain combinations. These restrictions are
following problems:

Aviator recognition time of engine-failure cues.


0

0
Time required to rotate from nose-low forward mode to a slight or
moderate nose-high attitude.
0
Altitude loss during the two problems above, with groundspeed
remaining as the tail-wheel, skid, guard, or cone hits the ground or
some other obstacle.

NOTE: These problems are very different from the usual low-level
autorotation, as practiced to a runway. The solution is to avoid operations
in avoid area (B) unless dictated by a tactical mission.
the aviator should
b. At slow airspeeds with an available landing site, 600
300 and 500 to feet for larger
normally allow feet for small helicopters
helicopters to set up a
steady-state autorotation and to complete a

reasonably safe landing.


10 knots, 200 feet, (point 1, parts A and B,
c. An engine failure at
Figure 6-44) requires a 2,700-FPM rate of descent to drive the rotorA at nor-
mal RPM. An engine failure at 20 knots, 150 feet, (point 2, parts and B,
Figure 6-44) requires a 2,100-FPM rate of descent to drive the rotor at nor-
mal RPM. Rates of descent in examples (1) and (2) of part B, Figure 6-44,
will not be attained. Therefore, rotor RPM decays. No deceleration lift is
6-51
possible to slow the rate of descent. Also, rotor inertia (RPM) will be low
for the collective-pitch application and touchdown. These combined effects
increase the possibility of a hard landing, which can cause structural damage
to the helicopter.
ALL GROSS WEIGHTS
A. 500

450

400

C SAFE-
A
350 U
T
j:'
IIJ
IIJ
I
!:!:.

~ 300
0
::I
0 N
a:
CJ

IIJ
>
0
~ 250 AVOID'
...
:E:
CJ
(A)
jjj
:E:

200
1

150 2

100

50

0
0 20 30 40 50 60 70 80
10

INDICATED AIRSPEED (KNOTS)

Figure 6-44. Height-velocity diagram

6-52
B.
3,400

3,100
:E
PL 2,800
\
BEST LIFT VERSUS DRAG MINIMUM
--0
I

I- RATE OF DESCENT IN AUTOROTATION


15 2,500
0
!II
~ 2,220
:\
I

~
w
1,900 ---- e
I
I I
I BEST CRUISE OR
I
I-
<( 1,600 I
I BEST DISTANCE IN
I
IX: I
I
AUTOROTATION
I
1,300 I
I

~~
I
1,000 I

:
I
I

:
\ '-"'-FULL FLARE AIRSPEED
PARTIAL FLARE AIRSPEED
EFFECTIVE DECELERATION AIRSPEED
500 I I MINIMUM SAFE AIRSPEED, LAST 100 FEET
: I
I
.

HEIGHT/VELOCITY CAÚTION AIRSPEEDS


I 10 to 400 FEET
I

0 10 20 30 40 50 60 70 III 90 100
AIRSPEED(J<NOTS)

Figure 6-44. Height-velocity diagram (continued)

SECTION
VIII
AUTOROTATION

6-24. AERODYNAMICS OF VERTICAL AUTOROTATION

a. During powered flight, rotor drag is overcome with engine power. When
the engine fails or is deliberately disengaged from the rotor system, some
other force must sustain rotor RPM so controlled flight can be continued to
the ground. This force is generated by adjusting the collective pitch to
allow a controlled descent. Airflow during helicopter descent provides
energy to overcome blade drag and to turn the rotor. When the helicopter
descends in this manner, it is in a state of autorotation. In effect, the
aviator gives up altitude at a controlled rate in return for energy to turn
the rotor at an RPM that provides aircraft control. Stated another way, the
helicopter has potential energy by virtue of its altitude. As this altitude
decreases, potential energy is converted to kinetic energy and stored in the
turning rotor. The aviator then uses this kinetic energy to cushion the
touchdown when nearing the ground.

b. Most autorotations are performed with forward speed. For simplicity,


the following aerodynamic explanation is based on a vertical autorotative

6-53
descent (no forward speed) in still air. Under these conditions, forces that
cause the blades to turn are similar for all blades regardless
of their posi-
tion in the plane of rotation. Therefore, dissymmetry of lift resulting from
helicopter airspeed is not a
factor. During vertical autorotation, the rotor
disk is divided into three regions, as illustrated in Figure 6-45: the
driven region, the driving region, and the stall region.

DRIVEN REGION 30"10

Figure 6-45. Blade regions in vertical autorotation descent

force vectors
c. Figure 6-46 shows four blade sections that illustrate
in the driven region (part A), a point of equilibrium (part B), the driving
region (part C), and the stall region (part E). Force vectors are different
wind is slower near the blade root
in each region because rotational relative
and increases continually toward the blade tip. Also, blade twist gives a
more positive angle of attack in the driving region than in the driven
region. The combination of the inflow up through the rotor with rotational
relative wind produces different combinations of aerodynamic force at every
point along the blade.
(1) Driven region. The driven region, also called the propeller
30 percent
region, is nearest the blade tips. Normally, it consists of about
of the radius (Figure 6-45). In the driven region (part A, Figure 6-46), the
total aerodynamic force acts behind the axis of rotation, resulting in an
overall dragging force. The driven region produces lift, but it also opposes
rotation and continually tends to decelerate the blade. This results in
a

drag force, which tends to slow the rotation of the blade. The size of this
region varies with the blade pitch setting, rate of descent, and rotor RPM.
When changing autorotative RPM, blade pitch, or rate of descent, the aviator
also changes the size of the driven region in relation to the other regions.

6-54
(2) Areas of equilibrium. There are two points of equilibrium on
the blade. Between the driven region and the driving region is a point of
equilibrium, shown in part B, Figure 6-46. Also, between the driving region
and the stall region is another point of equilibrium, shown in part D, Figure
6-46. At a point of equilibrium on the blade, total aerodynamic force is
aligned with the axis of rotation. Lift and drag are produced, but the total
effect produces neither acceleration nor deceleration.
(3) Driving Region. The driving region, or autorotative region,
normally lies between 25 and 70 percent of the blade radius (Figure 6-45).
Part C, Figure 6-46, shows the driving region of the blade, which produces
the forces needed to turn the blades during autorotation. Total aerodynamic
force in the driving region is inclined slightly forward of the axis of rota-
tion, producing a continual acceleration force. This inclination supplies
thrust, which tends to accelerate the rotation of the blade. Driving region
size varies with blade pitch setting, rate of descent, and rotor RPM. The
aviator controls the size of this region in relation to the driven and stall
regions to adjust autorotative RPM. For example, if the collective-pitch
stick is raised, the pitch angle increases in all regions. This causes the
point of equilibrium (part B, Figure 6-46) to move inboard along the blade
span, thus increasing the size of the driven region. The stall region also
becomes larger while the driving region becomes smaller. Reducing the size of
the driving region causes the acceleration force of the driving region and,
therefore, RPM to decrease.
(4) Stall region. The stall region includes the inboard 25 percent
of the blade radius (Figure 6-45). It operates above the stall angle of
attack, causing drag, which tends to slow the rotation of the blade. The
stall region is depicted in part E, Figure 6-46.

d. Aconstant rotor RPM is achieved by adjusting the collective-pitch


stick so blade acceleration forces from the driving region (part C, Figure
6-46) are balanced with the deceleration forces from the driven and stall
regions (parts A and E, Figure 6-46).

6-25. AERODYNAMICS OF AUTOROTATION IN FORWARD FLIGHT

a. Autorotative force in forward flight is produced in exactly the same


manner as when the helicopter is descending vertically in still
air. How-
ever, because forward speed changes the inflow of air up through the rotor
disk, all three regions move outboard along the blade span on the retreating
side of the disk where angle of attack is larger, as shown in Figure 6-47.
Because of lower angles of attack on the advancing-side blade, more of that
blade falls in the driven region. On the retreating side, more of the blade
is in the stall region. A small section near the root experiences a reversed
flow; the size of the driven region on the retreating side is reduced.

6-55
TAF A \
-1-- TOTAL \
\ AERODYNAMIC DRIVEN
I
REGION ,
I
~ FORCE AFT
I
I
t OF AXIS OF DRAG,
ROTATIONAL I
1 ROTATION
RELATIVE WIND I , I
I I
I
,
I
I , /
I \ I
I
I.; /
DRIVEN
í
INFLOW UP
~--1--
~ RESULTANT
RELATIVE WIND
REGION
/
/
/

THROUGH ROTOR TAF


\.. .
,B POINT OF

/
/' EQUILIBRIUM

/
/
I
I
EQUILIBRIUM. I
I W

I ~
, ~
I W c
('
INFLOW
\..
TAF
I
TOTAL
I
I
I
\
~
<(
b
a:
DRIVING
REGION
AERODYNAMIC
~
-

: FORCE FORWARD \
~ ----- OF AXIS OF \ ~
:\ ROTATION \
\
\
I \
\ ,
I \ ,
DRIVING
I \ "
ANGLE OF --', I
\ D REGION POINT OF
ATTACK 60 'D "- EQUILIBRIUM
"

"-
INFLOW '\
E \
\
STALL
\
REGION
DRAG \
I

ANGLE OF STALL REGION


ATTACK 240
(BLADE IS STALLED)

Figure 6-46. Force vectors in vertical autorotative descent

6-56
FORWARD

..
~ ADVANCING SIDE

Figure 6-47. Autorotative regions in forward flight

b. Autorotations may be divided into three distinct phases: entry,


steady-state descent, and deceleration and touchdown. Each phase is aerody-
namically different. The following describes forces pertinent to each phase.
(1) Entry. Entry into autorotation is performed after loss of en-
RPM decay and an
gine power. Immediate indications of power loss are rotor
out-of-trim condition as the nose of the helicopter yaws to the left because
of the loss of engine torque. Rate of RPM decay is most rapid when the heli-
copter is at high gross weight, high forward speed, or in high-density al-
titude conditions. All of these conditions demand increased collective pitch
rapid RPM decay
and torque to maintain powered flight; these conditions cause
when the engine stops. In most helicopters, it takes only seconds for the
RPM decay to
fall into minimum safe RPM
a
range. Aviators must quickly reduce
collective pitch to prevent excessive decay. A cyclic
flare will help
prevent excessive decay if the failure occurs at high speed. However, this
technique varies with the model of helicopter. Aviators should consult and
follow the appropriate aircraft operator's manual.

(a) Level powered fliRht at hiRh speed. Figure 6-48 shows the
airflow and force vectors for a blade in level powered flight at high speed.
The lift and drag vectors are large; the total aerodynamic force is inclined
well to the rear of the axis of rotation. If the engine stops when the heli-
RPM decay,
copter is in this condition, rotor RPM decay is rapid. To prevent
the aviator must immediately lower the collective pitch control to reduce
drag and incline the total aerodynamic force vector forward so it is near the
axis of rotation.

6-57
(b) Collective pitch reduction. Figure 6-49 shows the airflow
and force vectors for a blade just after power loss. Collective pitch has
been reduced, but the helicopter has not started to descend. Lift and drag
are reduced, and the total aerodynamic force vector is inclined further for-
ward than it in powered flight. As the helicopter begins to descend, the
was
airflow begins to flow up, under the rotor system. This causes the total
aerodynamic force to incline further forward until it reaches an equilibrium
that maintains a safe operating RPM. The aviator then establishes a glide at
the proper airspeed, depending on the type of helicopter and its gross
weight. Rotor RPM should be stabilized at autorotative RPM, which usually is
a few turns higher than normal operating RPM.

(2) Steady-state descent. Figure 6-50 shows the airflow and force
vectors for a blade in an autorotative, steady-state descent. Airflow is now
upward through the rotor disk because of the descent. Changed airflow
creates a
larger angle of attack, although blade pitch angle is the same as
it was in Figure 6-49 before the descent began. Total aerodynamic force is
increased and inclined further forward so equilibrium is established. Rate
of descent and RPM are stabilized; the helicopter descends at a constant
Angle of descent is normally 17 to 20 degrees, depending on variables
angle.
such as airspeed, density altitude, wind, and type of helicopter.
(3) Deceleration and touchdown. Figure 6-51 illustrates the aero-
dynamics of autorotative deceleration. To make an autorotative landing, the
aviator reduces airspeed and rate of descent just before touchdown. Both
actions can be partially accomplished 'by moving the cyclic control to the
rear, changing the attitude of the rotor disk with relation to the relative
wind. The attitude change inclines the total force of the rotor disk to the
rear and slows forward speed. It also increases angle of attack on all
blades by changing air inflow. As a result, total rotor lifting force is in-
creased and rate of descent is reduced. As shown in Figure 6-51, the RPM
also increases when the total aerodynamic force vector is lengthened; this
RPM
increases blade kinetic energy available to cushion the touchdown. The
increased because the rotor system entered the windmill brake state. After
forward speed is reduce~ to a safe landing speed, the helicopter is placed in
a landing attitude as collective pitch is applied to cushion the touchdown.
Specific values for RPM and airspeed and aviator technique are found in air-
craft operator's manuals.

6-58
AXIS OF
ROTATION
I
I
I TOTAL
I AERODYNAMIC
1- FORCE

ANGLE OF
ATTACK

RELATIVE WIND

Figure 6-48. Force vectors in level powered flight at high speed

TOTAL
AERODYNAMIC
FORCE

RELATIVE WIND

Figure 6-49. Force vectors after power loss with reduced collective

6-59
LIFT

CHORD LINE

ROTATIONAL
RELATIVE
WIND

DRAG

RESULTANT
RELATIVE WIND

Figure 6-50. Force vectors in autorotative steady-state descent

TOTAL
AERODYNAMIC

V
FORCE
".'"

\
\
ROTATIONAL \
RELATIVE \
WIND \
\
\

ANGLE
OF
ATTACK
RESULTANT
RELATIVE
INFLOW WIND

Figure 6-51. Autorotative deceleration

6-60
6-26. GLIDE AND RATE OF DESCENT IN AUTOROTATION

a. Helicopter airspeed is probably the most significant factor affecting


rate of descent in autorotation. Rate of descent is large at very low air-
speeds, decreases to a minimum at some intermediate speed, and then increases
again at faster speeds. The autorotative state is discussed in paragraph
6-21 on settling with power. The autorotative boundary in Figure 6-42 shows
the relationship between speed and rate of descent. Rate of descent is high-
est at zero horizontal speed. It decreases and reaches the minimum rate at
point A, an intermediate horizontal speed. Then, rate of descent increases
again and provides the shallowest possible flight-path angle at point B,
where the degree angle line is tangent to the autorotative boundary line.
This is the best horizontal speed for maximum glide distance.

b. specific airspeeds (given in the auto-


Each type of helicopter has
rotation chart of the operator's manual) at which a power-off glide will
cover maximum distance. This airspeed is usually at or slightly above normal
cruise airspeed. Airspeeds that result in the slowest rate of descent are
also shown in these charts. These airspeeds usually are at or near a slow-
cruise airspeed.

c. Specific airspeeds for maximum distance or slowest rate of descent


are based on the standard-density altitude, with average weather and wind
conditions and normal loading. When the helicopter is operated with exces-
sive loads in high-density altitudes or strong, gusty winds, a slightly in-
creased airspeed during the descent is best. For autorotation in light winds
and low-density altitudes, a slight decrease in normal airspeed is best.
Following this general procedure of fitting airspeed to existing conditions,
an aviator can achieve about the same glide angle in any circumstances and
estimate his touchdown point. For example, the best glide ratio (maximum
distance) for the average helicopter in a no-wind condition is about 4 feet
of forward glide to 1 foot of descent. Ideal airspeed for minimum rate of
descent is at slow-cruise values and with a glide ratio of 3 feet forward to
1
foot of descent. Above and below this airspeed, the rate of descent
rapidly increases.

d. The chart in Figure 6-52 shows typical rates of descent for the vari-
ous airspeeds for steady-state autorotation. This type of graph in an opera-
tor's manual gives basic information about precision autorotation. Normally,
the acceptable autorotation airspeed ranges for different model helicopters,
with aviators having average skills, vary from slightly less than slow-cruise
to slightly higher than cruise values (ranges 2 through 5 of Figures 6-52 and
6-53). A slight change of airspeed results in a large selection in rates of
descent in airspeeds of range 2 to midpoint range 3 of Figure 6-53. There-
fore, this is the best precision airspeed glide slope. Increasing or decreas-
ing the airspeed by as little as 5 knots in a steady-state autorotation in
this airspeed range may advance or retreat the point of ground contact
noticeably. Airspeeds of less than range 2 yield increasingly high rates of
descent.

e. Figure 6-53 shows examples of eight entry points for the entire
forced-landing and precision-autorotation envelope. These entry points show

6-61
positions on the front, back, and inside of the precision glide slope.
Important general considerations are listed below.

(1) The best precision airspeed range, shown in Figure 6-52, is


between range 2 and range 3. When plotted in profile, this airspeed spread
becomes the precision glide slope, or the cone of precision.

(2) The main effort in performing the precision autorotation at


entry points 1, 2, 4, 5, and
6 is to
intercept and stay inside the precision
glide slope. The precision glide slope must be intercepted as soon as pos-
sible. Then, a
steady-state airspeed is established and tested while a slow-
cruise attitude is held.
(3) The circle of action is a point on the ground that has no ap-
parent movement in the pilot's field of view during
a
steady-state autorota-
tion. The circle of action would be the point of impact if the pilot applied
no deceleration, initial pitch, or cushioning pitch during the last 100 feet
of autorotation. Depending on the amount of wind present and the rate and
amount of deceleration and collective application, the circle of action is
usually two or three helicopter lengths short of the touchdown point.
(4) For easier recognition, entry point 6 can be considered as the
entry position for the familiar standard autorotation.
A
(5) The precision-autorotation flight envelope ends at 100 feet.
basic termination can be made thereafter to a touchdown point, as shown in
of 3 and the
Figure 6-53, if the airspeed is within allowable tolerance range
rate of descent is normal.

6-62
RATE OF DESCENT (FPM)

~
~\
\

i\
\

1\
i' \
Q.
II.
II.
0
111
1\ "

0
W \ /
a:
0
z
;:)
~ ~ /
~
Z
/
w
0
111
W
'" " V
0 V
II.
0
'--
W
~
a:

AIRSPEED
0 VNE (KNOTS)

AIRSPEED
AIRSPEED RANGE 1 RANGE 4
AIRSPEED RANGE 5
HEIGHTNELOCITY CAUTION (BEST PARTIAL
AREAS USE ONLY AT ALTITUDE DECELERATION BEST DISTANCE RANGE
(SEE HEIGHT-VELOCITY DIAGRAM) SPEED) (BEST FULL-FLARE SPEED)

AIRSPEED RANGE 3
AIRSPEED RANGE 2 SLOWEST RATE/DESCENT
(BEST PRECISION GLIDE SLOPE) (BEST DECELERATION SPEED)
SAME AS SLOW CRUISE

Figure 6-52. Steady-state autorotative rate of descent


for various airspeeds

6-63
AIRSPEED RANGE 4 AIRSPEED RANGE 1
.
......

ï=:o(
400 UTILITY Õ ~
HELICOPTER g 0(
300
~ ~
OBSERVATION
j!.-
:I: ~
200 HELICOPTER ~~
:I:

TERRAIN
CA TD

CIRCLE OF ACTION (APPARENT POINT OF COLLISION OR ACTUAL POINT


~ /
OF COLLISION IF COLLECTIVE PITCH WERE NOT APPLIED)
TOUCHDOWN ./
(TWO HELICOPTER LENGTHS BEYOND CIRCLE OF ACTION)

Figure 6-53. Forced-landing autorotative flight envelope for various


types of helicopters

(6)For maneuver repeatability, exact attitudes must be used or


noted throughout autorotation. The center of attention is split between
attitude and the circle-of-action point. All other references, such as air-
speed and rotor RPM, are read in a running cross-check.

(7)Airspeed values and restrictions of the height-velocity diagram


must be scaled up to comply with the performance charts of larger
helicopters. Height-velocity diagrams are based on a standard day at sea
level; the envelopes are expanded in proportion to increasing density
altitude.

f. The following procedures described under entry points 1, 2, and 3 are


for discussion only and should not be performed in training. Considerable
risk is involved in performing these maneuvers. They are only included so
the aviator will know how to make these types of landings in case of an ac-
tual forced landing.
(1) Entry point 1 (not a training maneuver). At entry point 1 (Fig-
ure 6-53), the touchdown point appears to be almost vertical to the aviator.

At cruise airspeed into the wind and at 700 feet above


(a)
ground level when the throttle is reduced, lower collective pitch, hold head-
ing, and decelerate promptly for speed-reduction climb, stopping all apparent
ground speed at the intended landing spot.

6-64
(b) Hold a nose-high attitude until the airspeed goes through
15 knots; then slowly lower the attitude at a rate to establish
a O-knot
reading and a slow-cruise or hovering attitude. (Optional--make S-turns
holding range-2 airspeed.)

(c) Settle vertically; head wind will cause a


slight rearward
movement.

When the helicopter is about to intercept the precision


(d)
glide slope, lower attitude smoothly and progressively to a point slightly
below the normal takeoff acceleration attitude.

(e) When airspeed is between range 2 and range 3, rotate to a


slow-cruise attitude. When reading the precision line of descent, observe
that the circle-of-action point is reliable only when the attitude is at slow
cruise and when a steady-state autorotation is in progress (no deceleration,
no acceleration).

(f) Watch the circle-of-action point for evidence of overshoot-


ing or undershooting.
5 knots and then
(g) If undershooting, lower attitude to gain
return attitude to slow cruise.
5 knots and then
(h) overshooting, raise attitude to lose
If
return attitude to slow cruise. When reading the precision line of descent,
observe that the circle-of-action point is reliable only when the attitude is
at slow cruise and when a steady-state autorotation is in progress (no decel-
eration, no acceleration).
At 100 feet when airspeed is within allowable tolerance of
(i)
range 3, terminate as in a standard autorotation for
a landing at the touch-
down point.
At 100 feet when airspeed is within range 2, hold
(j)
slow-cruise attitude to about 50 feet and then rotate to the normal
deceleration attitude.
(k) Touch down at the desired point as in basic autorotation
touchdown.

(2) Entry point 2 (not a trainin~ maneuver). In entry point 2 (Fig-


ure 6-53), the aviator estimates that he is almost beyond the precision glide
slope.

(a) At cruise airspeed and 700 feet AGL when the throttle is
reduced, lower collective pitch, hold heading, and decelerate promptly for
a

speed-reduction climb, stopping all apparent ground speed at the intended


landing spot.

6-65
(b) As the apparent ground speed reaches zero knots, lower
attitude to the slow-cruise attitude. (The airspeed will now be equal to or
near the wind velocity.)

(c) Settle vertically and continue as indicated in the entry


1
point procedure to remain within the precision glide slope.

(3) Entry point 3 (not a training maneuver). In entry point 3 (Fig-


ure 6-53), the aviator estimates that he is well into the precision glide
slope.

(a)
At cruise airspeed and 700 feet AGL when the throttle is
reduced, lower collective pitch, hold heading, and make a speed-reduction
climb.

(b) As the airspeed approaches between range 2 and range 3


(depending on the head-wind effect on ground speed), lower to slow-cruise
attitude for a steady-state autorotation. Proceed as indicated in the entry
point 1 procedure to remain within the precision glide slope.

(4) Entry point 4. In entry point 4 (Figure 6-53), the aviator


estimates that he is just short of the precision glide slope.

(a) At cruise airspeed and 700 feet AGL when the throttle is
reduced, lower collective pitch, hold heading, and decelerate smoothly. This
will cause an ascent to the precision glide slope.
(b) As the airspeed approaches between range 2 and range 3
(depending on the head-wind effect on ground speed), lower attitude to the
slow-cruise attitude for a steady-state autorotation. Proceed as indicated
in the entry point 1 procedure to remain within the precision glide slope.
NOTE: Entry point 4 is used when an ideal precision autorotation is
demonstrated.

(5) Entry point 5. At entry point 5 (Figure 6-53), the aviator


estimates that he is well short of the precision glide slope.

(a) At cruise airspeed and 700 feet AGL when the throttle is
reduced, lower collective pitch and hold heading, cruise attitude, and rotor
RPM for best
distance. (Hold crab, rather than slip, for best distance.)
(b) When the precision glide slope is just ahead, do a
partial
deceleration. This causes an ascent to the precision glide slope.

(c)airspeed approaches between range 2 and range 3, rotate


As
attitude to slow cruise for a steady-state autorotation and proceed as indi-
cated in the entry point 1 procedure to remain within the precision glide
slope.

(6) Entry point 6. In entry point 6 (Figure 6-53), the aviator es-
timates that he is almost too far back to intercept the precision glide

6-66
slope. He proceeds as in entry point 5 to possibly intercept the precision
glide slope further down the line of descent. However, he may decide to
proceed as indicated in the procedure for entry point 7.

(7) Entry point 7. At entry point 7 (Figure 6-53), the aviator


estimates that he cannot intercept the precision glide slope. The line of
descent appears to be a spot well short of the touchdown point.

(a) At cruise airspeed and 700 feet AGL when the throttle is
reduced, lower collective pitch and hold heading and cruise attitude for best
distance.
(b)
At about 200 feet, begin a smooth partial deceleration,
converting speed to lift. This changes the line of descent toward the touch-
down point. By regulating the rate and amount of deceleration from 200 feet
on, the aviator can make
a basic termination at the touchdown point.

(8) Entry point 8. This exercise is identical to the procedure for


8 is set up farther away
entry point 7, except that the entry for entry point
from the precision glide slope. The line of descent appears to be to point
a

100 feet or more short of the normal circle-of-action point.

(a) Hold best distance attitude, rotor RPM, and pedal trim.
Upon reaching 40 to 60 feet altitude, execute full deceleration that is
a

regulated in rate and amount of attitude rotation so as to arrive at the


touchdown point at the end of deceleration.

(b) Allow the helicopter to settle to 15 to 20 feet, apply


initial collective pitch, rotate attitude to level landing attitude, and
apply a firm collective pitch in the amount and rate necessary to cushion the
landing.

6-27. THE LAST 100 FEET

ends at 100 feet and that


a. The Power-Off Landing. Assume autorotation
the power-off landing procedure begins there. To execute a power-off landing
for rotary-wing aircraft, the aviator obtains a smooth trade-off of airspeed
100 beginning at 100 feet, airspeed
for lift during the last feet. Ideally,
is converted to additional lift by deceleration. Deceleration is applied and
timed so rate of descent and forward speed are reduced just before touchdown
to the slowest rates possible for the existing conditions.
(1) Potential energy available for power-off landing. .At 100 feet,
the aviator must begin spending stored flight energies; that is, the forward
velocity of the helicopter and, just before touchdown, the rotational energy
of the main rotor. At 100 feet, he can predict with accuracy the amount of
potential energy (deceleration or cyclic lifting power) available for
a

power-off landing. He can also predict the effectiveness of applying collec-


tive pitch to cushion the touchdown.
Rate of descent and ground speed. All the heavy aerodynamic
(2)
work of reducing the rate of descent and slowing the ground speed should be
a

6-67
result of the aviator's decelerating, downto about 15 feet. Thereafter, his
use of collective pitch further slows andat times may delay the descent and
then cushions the touchdown. Figure 6-54 shows predictable conditions for
the power-off landing.

b. Terms and Definitions. So that the discussion of power-off landings


can be understood, the following terms are defined.

(1) Attitude rotation. A preplanned or scheduled change of aircraft


attitude at some specific point in a maneuver sequence.

(2)Deceleration. A trade-off of airspeed for lift while a


rela-
tively continuous line of descent is held or maintained.
(3) Partial deceleration. A trade-off of airspeed for
results in a moderate change to the line of flight.
lift that

(4) Full deceleration. A trade-off of airspeed for


lift that
results in a substantial change to the line of flight. In autorotation at
the appropriate altitude, the descending line of flight is changed by con-
verting airspeed to lift so as to parallel the ground for some distance.
NOTE: The terms deceleration, partial deceleration, and full deceleration do
not apply to the attitude rotation per se but to the change in lift/line of
flight that results from attitude rotation.

c. Conditions. Figure 6-53 shows airspeed conditions at 100 feet. The


following paragraphs describe landing sequences resulting from those airspeed
conditions, as illustrated in Figure 6-54.
(1) Condition 5. Condition 5 (airspeed range 5, Figure 6-54) exists
at 100 feet. (For best distance-gliding airspeed, see the appropriate air-
craft operatorls manual.) In condition 5, the helicopter descends on a very
narrow rotor profile to the line of descent. The helicopter then encounters
a
large volume of air per second. This can produce exceptional lifting
forces when the attitude is rotated smoothly and progressively and the full-
rotor-diameter profile is presented to (or against) the line of descending
flight. This attitude rotation should usually occur at 30 to 60 feet, depend-
ing on the aircraft. The added lift generated by the full deceleration is so
great that the descent will stop and the line of flight will parallel the
ground for some distance. As the deceleration ends--with the density
alti-
tude, wind, or gross weight favorable--the helicopter settles gently to a
point where the aviator applies initial pitch. This is followed by the avia-
torls final application of collective pitch for a soft touchdown and a near-
zero ground run. All of this is predictable at 100 feet.

6-68
" AIRSPEED RANGE 1
AIRSPEED RANGE 4 AIRSPEED RANGE 3 AIRSPEED RANGE 2
\
0\
I
\ \1
"
, " , \ \
,
"
, ,

\ \
\\
I I
'" "'- \

'~
'-....

\j~I\
........
,,"' \ \

~,",,'~
100 FEET \
\ \ ~
~...."':'''''',
.... "", " .... '\ ~ I
AIRSPEEÞ RANGE 5
'....
"

""""~:""'"....,
" ~
'-
'''''''-'
"
\~
"~\"''-
"'"
.......
"
.......
'.....
""'''......
"

~
\
~\ I
\ I
\

\
\
\
\
,\
\\1
,

\
r\
I
I

,,"'" """"
.......
'"

\\1\\ \ I
"",',
'..... """'.....
~ ',- " -"',
....~ \,
\

\~
I

............', ~ 'e.:.',> \\1 I


....

"'<:,'------
'--~~ ~
'U

~////////øø~ Figure 6-54.

Condition 4. Condition 4 (airspeed range 4, Figure 6-54) exists


(2)
The last 100 feet

at 100 feet. In condition 4, the helicopter descends on a narrow rotor pro-


file to the line of descent. A smooth and progressive attitude change (de-
celeration), which presents a full-rotor-diameter profile to (or against) the
line of descent, will alter the line of descent and add lift. When properly
timed, this added lift greatly reduces the rate of descent and forward speed
before collective pitch is applied initially. The deceleration may also be
used to increase rotor RPM before applying collective pitch. All of this is
predictable at 100 feet.
NOTE: The termination is necessary for a zero
full or partial deceleration
ground run and is mandatory for helicopters having low rotor inertia with
light, unweighted blades or poor collective-pitch effectiveness or both at
termination.
(3) Condition 3. Condition 3 (airspeed range 3, Figure 6-54) exists
at 100 feet. In condition 3, the helicopter descends on slightly less than a
full-rotor-diameter profile to the line of descent; translational lift is
near maximum effect; the rate of descent is minimum. A smooth and
progressive rotation of attitude, which presents a full-rotor-diameter
profile to (or against) the line of descent, will result in an effective
deceleration. This deceleration, while not noticeably changing the line of
descent, will reduce the rate of descent and the forward speed to a point
where collective-pitch energy is quite effective. If the deceleration is
timed correctly, descent is often stopped completely when the aviator applies
initial pitch. When wind, density, altitude, and weight are favorable, a this
leaves adequate pitch to delay and cushion the touchdown, resulting in
relatively short ground run. This is all predictable at 100 feet.
6-69
(4) Condition 2. Condition 2 (airspeed range 2, Figure 6-54) exists
at 100 feet. In condition 2, the helicopter descends on nearly a
full-rotor-diameter profile. The aviator should know that nothing will be
gained by attitude rotation; he should just hold a steady attitude to main-
tain the speed, at least down to 50 feet; he should know that, of the five
conditions, condition 2 will give the longest ground run. Therefore, at
about 50 feet, he will begin a progressive attitude change until a slight
rearward tilt of the rotor occurs just before applying collective pitch. The
attitude change will not supply additional lift; it will add a rearward com-
ponent of lift during pitch application. This reduces the ground run.

(a) With no effective deceleration lift in progress during the


last 100 feet, applying collective pitch alone does not provide enough
lift-
ing and braking action to appreciably delay touchdown and slow the ground
run. The ground run is about three to four lengths. This is all predictable
at 100 feet.
(b) Condition 2 falls on the border of the height-velocity
diagram (Figure 6-44). Often, a wind gradient/high-density altitude condi-
tion can then increase the rate of descent, which increases the lift demands
on the collective-pitch application. When accidents occur under condition 2,
the resulting accident summary usually states that damage was caused by a
late and insufficient application of collective pitch. Actually, the error
occurred earlier--at 100 feet--and was due to a lack of knowledge,
cross-check, projection, and prediction. When condition 2 is performed
knowledgeably with normal atmospheric and gross weights, it is a safe
operation.
(5) Condition 1. Condition 1 (airspeed'range 1, Figure 6-54) falls
in the avoid area of the height-velocity diagram (Figure 6-44). It exists at
100 feet when the helicopter is descending on a
full-rotor-diameter profile
to the line of descent. There is a high sink rate; no deceleration lift is
possible. Because of wind gusts, a wind gradient, or a wind shift, this
condition may suddenly occur in the last 100 feet of descent. The entire
rate of descent must then be stopped by applying collective pitch alone.
Usually, the lift produced is not enough for a safe landing. Condition 1 may
also cause obvious or hidden damage to the helicopter because of hard land-
ings. Such damage might be acceptable for an actual engine failure; it is
never acceptable for normal training practice. A termination with power is
necessary. This is all predictable at 100 feet.

6-70
SECTION IX

PERFORMANCE CURVES

6-28. PERFORMANCE FACTORS

a. To effectively use an aircraft, an aviator needs to know how well it


performs with varying combinations of weight, altitude, and temperature. This
section deals with the basic construction of performance charts, including
the performance factors listed in the following paragraph.

b. The ability of
helicopter to perform any maneuver depends on the
a

power required to perform the maneuver and the power available to the rotor
system during its execution. The performance section of each aircraft opera-
tor's manual has charts depicting certain performance factors with varying
combinations of weight, altitude, and temperature. Performance factors
available from these charts are--
0
Hover power.
0
Takeoff obstacle clearance distance.
0
Maximum velocity of the aircraft.
0

Velocity for best rate of climb.


0
Velocity for best angle of climb.
0
Velocity for maximum endurance.

0
Velocity for maximum range.

6-29. POWER REQUIREMENTS

a. In Chapter 2, paragraph 2-15, the total drag curve of a


helicopter
was developed. Figure 2-25 depicted total drag as being the sum of induced
drag, profile drag, and parasite drag. In that figure, total drag was higþ
during hovering flight, decreased to a minimum as airspeed increased, and
then increased with additional airspeed.

A
similar diagram can be used to depict the power requirements with
b.
varying airspeeds, as shown in Figure 6-55. The vertical scale on the chart
can be changed from drag (in pounds) to a variable directly related to power.
Some possible variables are as follows.

(1) Thrust (in pounds). Each pound of total drag requires 1 pound
of thrust to maintain steady-state flight. Some charts are drawn in this
manner. These are not meaningful to the aviator; he has no instrument to
indicate thrust.

6-71
(2) Power (in horsepower). The engine must perform a definite amount
of work (force times distance) on the rotor system per unit of time
(seconds). This relationship, which equates to power, is portrayed on the
vertical scale as horsepower. However, this is not meaningful to an aviator;
he has no instrument to indicate horsepower.

(3) Torque (in foot-pounds per square inch or percentage of power


engine can produce). Most Army helicopters have an instrument to indicate
the amount of turning force (torque) produced by engines. This variable is
used in the performance charts; torque is the main indication of the power
being developed.

(4) Fuel flow (in pounds per hour). With other variables held con-
stant, the rate of fuel consumption is directly proportional to the power
being developed. Some Army aircraft have fuel flowmeters; readings of these
can be useful.

!i
0
oJ
LL. POWER
oJ
w
::I
REQUIRED
LL.
W
::I
0
a:
0
l-
I

a: a:
w w

;: ;:
0 0
a. a.
w
VI
a:
0
:t
~
VI
::I
a:
:t
t-
O
AIRSPEED
0

Figure 6-55. Power required and airspeed relationship

6-72
6-30. POWER AVAILABLE

a. Power available from an engine can be shown on


performance chart, a

as indicated in Figure 6-56. The power available is fairly constant with


airspeed. Power available at higher airspeeds may slightly increase or
decrease, depending on the characteristics of the power plant installed in
the aircraft.

The power available is greatly influenced by density altitude. As


b.
altitude or temperature increases, power available to the rotor system will
decrease (Figure 6-56). For this reason, charts in the aircraft operator's
manual cover many different temperatures and altitudes.

NOTE: For turbine-powered aircraft, temperature is generally more power-


restrictive than pressure altitude is.

POWER AVAILABLE -
SEA LEVEL

5,000 FEET

10,000 FEET

W
::::I 15,000 FEET
a
II:
0
t.

II:
W
~
0
CL

0
0
AIRSPEED

Figure 6-56. Power available influenced by altitude

6-31. FORWARD FLIGHT PERFORMANCE

a. Maximum Speed Restrictions.


(1) level flight is often limited by structural
The maximum speed in
or controllability considerations. The most notable is the effect of
retreating blade stall (paragraph 6-20). Maximum speed is often labeled on
performance charts as a never exceed speed (Vne)' as shown in Figure 6-57.

6-73
(2) Another restriction on the maximum speed may be the availability
of power. The intersection of the power required and power available lines
defines the maximum speed in level flight with maximum continuous power (Vh)
of the helicopter. An increase in power required because of increased weight
or by G-producing maneuvers decreases Vh. A decrease in power available
caused by increased density altitude or faulty engines also decreases Vh'

(3)Higher speeds than Vh can be obtained by diving, provided Vne is


not exceeded. In Figure 6-57, the speed of the 5,OOO-pound aircraft is lim-
ited by Vne' while the speed of the 7,OOO-pound aircraft is limited by the
power available.

POWER AVAILABLE

POWER
REQUIRED
j
I:
0

t.
It
W
~
0
ø..

0
0 AIRSPEED -Vmax vne

Figure 6-57. Airspeed limitations

b. Best-Rate~of-Climb Airspeed (v~l. The best rate of climb (greatest


change in altitude per unit of time) occurs at an airspeed where the level
flight excess power (power available minus power required) is greatest. Be-
cause power available is fairly constant with airspeed, the best-rate-of-
climb airspeed is usually the point where minimum power is required, as shown
in Figure 6-58. Varying weights and density altitudes will shift the power-
required and power-available curves, thus changing the best-rate-of-climb
airspeed. Generally, an increase in weight increases the best-rate-of-climb
speed and an increase in density altitude decreases it. Because the horizon-
tal and vertical velocities are within the air mass (TAS) , wind does not
affect the rate of climb. The following factors do affect rate of climb.

6-74
(1)Altitude. Altitude affects engine performance. Increases in
altitude decrease the rate-of-climb performance. The rate of. climb at the
absolute ceiling of an aircraft is zero. At this altitude, there is no ex-
cess horsepower (HPA equals HPR)' At an altitude called the service ceil-
ing, an aircraft can maintain a 100-FPM rate of climb; when operating on a
single engine, the aircraft can maintain a 50-FPM rate of climb.
(2)Weight. As with angle of climb, weight affects climbing perfor-
mance. As weight increases, horsepower required increases; therefore, in-
creased weight and decreased excess horsepower will decrease the rate of
climb. As an aircraft burns fuel, its weight decreases. Because of this
weight decrease, more excess horsepower is available toward the end of a
flight for climb or acceleration.

POWER AVAILABLE

EXCESS POWER

CD
:I ~
D" 0
..
0
..J
t:.. u..
..J
a: W
w
~ ;:)
u..
0 MINIMUM REQUIRED MINIMUM FUEL
Q. FLOW
TORQUE

BEST-RATE-OF-
CLIMB AND MAXIMUM-
ENDURANCE AIRSPEED

0
0 AIRSPEED

Figure 6-58. Best-rate-of-climb airspeed and maximum-


endurance airspeed

c. Maximum-Endurance Airspeed. The helicopter achieves its maximum


flight endurance at an airspeed where fuel consumption is minimum. This
airspeed is when power required is minimum. Generally, this is the same as
the best-rate-of-climb airspeed in Figure 6-58.

d. Best-Angle-of-Climb Airspeed (VxL'

(1)The best angle of climb obtains the maximum altitude in a given


distance. excess OGE power exists, the climb can be made vertically at
If
zero airspeed. This is the case at sea level, as shown in Figure 6-59. When
the power required for an OGE hover exceeds the power available, a vertical

6-75
(zero airspeed) climb is impossible; forward airspeed is needed to achieve
the maximum angle of climb. This airspeed is less than the best-rate-of-
climb airspeed and can be determined mathematically. A simplified method is
A
shown in Figure 6-59. straight line is drawn from the intersection of zero
airspeed and power available tangent to the power-required curve. The best-
angle-of-climb airspeed is at the point of tangency.

POWER AVAILABLE. SEA LEVEL

10,000 FEET

-;-
~
CT
...
0
!:.
a:
w
==
0
Do
BEST-ANGLE-OF-: BEST-RATE-OF-
CLIMB AIRSPEED CLIMB AIRSPEED

0
0 AIRSPEED

Figure 6-59. Best-angle-of-climb airspeed

(2) The angle of climb that can be maintained depends on the amount
of excess thrust available. Flight at Vx may put the helicopter in a criti-
cal flight condition where a loss of power may place the helicopter in the
caution or avoid areas of the height-velocity diagram (Figure 6-44). The
recommended airspeed for maximum-climb angle, such as the obstacle-clearance
airspeed listed in some operator's manuals, is not a true Vx but is what
could be termed a safe best-angle-of-climb speed. This airspeed is greater
than the true maximum-angle-of-climb speed; it places the aircraft in a safer
flight envelope while only slightly sacrificing climb performance. During
takeoff, when obstacle clearance is the primary concern, Vx should be used.
This gives the most altitude for the horizontal distance covered.

NOTE: Because helicopter takes off vertically, angle of climb is academic


a

as long as the helicopter is below its hovering ceiling. However, an obsta-


cle clear?nce takeoff needs enough airspeed to reduce exposure time in the
avoid and caution areas of the height-velocity diagram (Figure 6-44).

6-76
(3) Altitude, weight, and wind affect angle of climb. These
effects are discussed below.
(a) Altitude. As an aircraft gains altitude, thrust developed
by the engine normally decreases. This is true for both turbine and recipro-
cating engines. The angle of climb must also decrease because as thrust
available (TA) decreases, excess thrust also decreases. Thrust required (TR)
remains essentially constant at all altitudes. Because of this, the aircraft
angle of climb decreases to zero degrees when it reaches its absolute ceil-
ing, where TA equals TR'

(b) Weight. Increased weight adversely affects angle-of-climb


performance in two ways; an increase in weight (W) increases both Wand TR'
Simplified, this means there is more weight to raise with less excess thrust.
The angle of climb is, therefore, shallower.

(c) Wind. When obstacle clearance is primary, the aircraft's


best-angle-of-climb speed (Vx) should be used. This provides the most al-
titude gained for distance covered. Wind is a factor because it affects the
horizontal distance covered to clear an obstacle. If an aircraft climbs at
Vx' the horizontal distance covered across the ground in a head wind is less
than the horizontal distance covered with no wind or with a tail wind (Figure
6-60). This affects the angle the aircraft climbs over the ground.

(NOTE CHANGES IN CLIMB ANGLE.)

Figure 6-60. Wind effect on aircraft at maximum-climb angle

e. Maximum-Range Airspeed.

(1) For any combination of weight and altitude, there is an airspeed


that results in the greatest distance traveled (best range). This occurs
when the specific range (nautical miles per pound of fuel consumed)
is maxi-
mum. Specific range is obtained by dividing the fuel flow (pounds per hour)
by true airspeed (nautical miles per hour).

6-77
(2) The airspeed maximum specific range occurs is de-
at which the
termined mathematically, but a much easier method is shown in Figure 6-61.
The speed for the best range is found at the point where the power-required
line is tangent to a line drawn from the point of origin (point of zero fuel
flow at zero airspeed). In some instances, maximum-range airspeed is higher
than Vne' In such cases, the Vne will be the maximum airspeed to be used for
maximum range. The airspeed for maximum range decreases as weight decreases,
as shown in Figure 6-62. To obtain the maximum range on an extended flight,
the aviator will have to periodically reduce airspeed as fuel is consumed.

POWER AVAILABLE

-;-
::I
C"
...
0
t:-
o::
w

;:
0
Q.

AIRSPEED FOR
MAXIMUM RANGE

0
0
AIRSPEED

Figure 6-61. Airspeed for maximum range

6-78
POWER AVAILABLE

POWER
REQUIRED

'i'
::::I

E"
0

t..
a::
w
~
0
D.

AIRSPEED FOR
MAXIMUM RANGE

0
0
AIRSPEED

Figure 6-62. Maximum-range airspeeds at various weights

SECTION X

MANEUVERING FLIGHT

6-32. ACCELERATION

In maneuvering flight, the aircraft changes airspeed or direction or both.


For a change to occur, an unbalanced force (accelerating force) must exist in
the direction of the change. The magnitude of this accelerating force de-
pends on the rate of change (acceleration) and mass of the helicopter. This
accelerating force can be obtained by tilting the rotor disk in the appropri-
ate direction. For example, for the helicopter to accelerate forward, the
disk must be tilted forward; to turn right, the disk is tilted right. These
changes are shown in Figure 6-63.

6-79
ACCELERATION VECTORS

ACCELERATION IN
ACCELERATION IN TURNING FLIGHT
FORWARD FLIGHT

Figure 6-63. Acceleration vectors in forward and turning flight

a. Straight and Level Acceleration. To accelerate forward, the rotor


disk must be tilted forward to produce a horizontal accelerating force (Fig-
ure 6-63). To maintain enough vertical thrust (force) to balance weight
during acceleration, the main rotor system must produce additional thrust.
The additional power required is directly related to the degree of tilt of
the rotor disk. For example, a disk tilt of 17 degrees requires a 5 percent
increase in power; a 30-degree tilt requires a 15 percent increase. The
maximum translational acceleration of a helicopter flying at constant alti-
tude is limited by the available power and any disk-tilting restriction.
During NOE flight, the altitude of the helicopter further restricts accelera-
tion. At extremely low altitudes, tilting the rotor disk excessively--unless
additional power is added at the same time--can cause the helicopter to
descend into the ground or obstacles.

b. Level Turning Flight. In this discussion, level turn is


a
turn in
a

which altitude is kept constant. To turn the helicopter, a


turning force
(centripetal force) must be applied perpendicular to the flight path. This
force is created by banking the rotor system in the direction of turn, as
shown in Figure 6-64. The horizontal component of the thrust is the turning
A
force; the vertical component is equal to and opposite weight. steeper
bank angle produces a
greater turning force but also requires additional
power (thrust). The percentage of increase in power required during turning
flight, as opposed to straight flight, is directly related to the bank angle,
as shown in Figure 6-64. A 15-degree bank, as used for standard-rate turns

at 90 knots, only requires a power increase of 3.6 percent; a 45-degree bank


requires a power increase of more than 41 percent. Figure 6-65 shows how
turning flight is restricted by airspeed and angle of bank. This chart
represents the performance of a helicopter of a certain gross weight operat-
ing at a given density altitude. Enough power is available for sustained
30-degree bank turns at all airspeeds. Power is not sufficient to sustain
45-degree banks at very high or low airspeeds; a 60-degree bank can only be
sustained at speeds near the best-rate-of-climb airspeed or at minimum gross
weight. Turning maneuvers requiring more power than is available can be

6-80
completed only by losing altitude. During terrain-flight modes, such an
altitude loss can be disastrous.
TURNING FORCE
f
-

t BANK
'I ANGLE

r-
I BANK ANGLE
(0)
INCREASE IN
TR(%)
I
0 -

15 3.6
30 15.4
45 41.4
60 100.0

WEIGHT

Figure 6-64. Effect of turning in forward flight

J
~
0

t:.
a:
w
==
0
A. BANK ANGLE

0
0
AIRSPEED
Vne

Figure 6-65. Power required at various angles of bank


during level turning flight

6-81
6-33. TURNING PERFORMANCE

Unlike ground-supported vehicle, an aircraft can rotate


an automobile or a

about three axes; this gives it six degrees of motion. It can pitch up or
down, yaw left or right, and roll left or right. Because of this freedom of
motion, an aircraft can perform many useful maneuvers. All of these maneu-
vers use turns--vertical, horizontal, or both. This section considers verti-
cal and horizontal turns separately, as well as the limits imposed on these
turns. In this discussion, certain assumptions are made about the terms
defined below.

a. When an it is not in static equilibrium. Forces must


aircraft turns
be unbalanced to produce the acceleration for the turn. When properly per-
formed, the turn does not produce any sideward force pulling the aviator
inward or outward from the turn; the resultant lift force acts toward the
A
center of the turn. This turn is called a coordinated turn. term that
sometimes causes confusion is level turn. This refers to a
constant-altitude
turn, not to a wings-level turn. Normally, an aircraft is never turned a with
the wings level. Forces acting on the aircraft must be unbalanced for turn
to occur; this does not happen if wings are level. Incorrect banking during
uncoordinated turns causes slips or skids and is uncomfortable for crew and
be
passengers. During uncoordinated turns at slow airspeeds, control can
inadvertently lost.
The force that actually turns the aircraft is the
b. force. Thelift
horizontal component of is the force that accelerates the aircraft
lift
toward the center of the turn. Increased collective pitch increases the
angle of attack; this produces the added required because of the reduc-
lift
tion of the vertical lift
component and the apparent increase in weight pro-
duced by centrifugal force.
'

6-34. TURNING FLIGHT

An aircraft, like any moving object, requires a sideward force to


a.
make turn. In
a normal turn, this force is supplied by banking the air-
a

craft so that lift is exerted inward as well as upward. The force The of lift is
then separated into two components at right angles to each other. liftThe
acting upward and opposing weight is called the vertical-lift component.
and opposing inertia or centrifugal force is the
lift acting horizontally
horizontal-lift component (centripetal force), as shown in Figure 6-66. The
sideward force that forces the aircraft from straight flight, causing it to
turn, is the horizontal-lift component. If an aircraft is not banked, there
is no force to make it turn unless the aircraft is skidded in the turn by
pedal application. Likewise, if an aircraft is banked, it turns unless held
on a constant heading with the opposite pedal. Proper control technique
assumes that an aircraft is turned by banking and that in a banking attitude
it should be turning.

6-82
Banking an aircraft in a level turn does not by itself change the
b.
amount of lift.
However, the division of lift
into horizontal and vertical
components reduces the amount of lift
supporting the weight of the aircraft.
Thus with a reduced vertical component, altitude is lost unless the
total
lift is increased by increasing collective pitch.

CENTRIPETAL FORCE
VERTICAL ,
COMPONENT
OF LIFT t
~
~
I
BANK ANGLE ~
I ~

/
f..tþ
I
'f( \
CENTRIFUGAL FORCE

Figure 6-66. Turning flight

6-35. RADIUS OF TURN

The radius of turn varies directly with the square of its velocity
(true
airspeed) and inversely with the angle of bank. Thus any two aircraft that
can fly at the same velocity and angle of bank can fly in formation, regard-
less of their weights. Only these two variables directly affect the radius
of turn. However, certain other factors affect the velocity and indirectly
affect the radius of turn. These factors are the weight, altitude, load
factor, angle of attack, and wing area. All of these aerodynamic factors
affect the lift
force that must be produced. To turn an aircraft in the
smallest possible radius, an aviator flies at the slowest possible speed and
the highest possible angle of bank, as shown in Figure 6-67. Also imposed on
the minimum radius of turn are the aerodynamic, structural, and power limits.
The aviator must be constantly aware of these limits while maneuvering the
aircraft at or near its design limitations.

Aerodynamic Limit of Performance. Fixed-wing aircraft reach their


a.
aerodynamic radius-of-turn limit at their stall airspeed. However, the
helicopter has no aerodynamic limit to its radius of turn as long as its
power and structural limitations are not exceeded. Structural limits can be
exceeded by using too much tail-rotor thrust to speed up a turn. Of course,
an increase in airspeed increases the radius of turn at a given bank angle.

6-83
10 ,ooo~
9,000~ '"
8 ,OOO~ ~
00
= c
7,000:; z
: 0
6,00~ -
u
UJ
-

!J)
5,000: a:
UJ
Q.
4.000= !J)
UJ
UJ
a:
3.000~ æ
Z
a: e.
:J z
~
a:
LL
2.000: ~ 2 N
0
!J) LL
0
:J
Õ UJ
~
~
~
a:
a:

'"
00

200

N ~
'" .... '" N
'"

Figure 6-67. General turning performance (constant altitude, steady turn)

6-84
b. Structural Limit of Performance.
(1) The load function of the angle of bank; the
factor is purely a

weight of the aircraft itself has no effect on the G-load imposed on the
2-G acceleration in a
aircraft. Both the OH-58 and the UH-60 experience
a

60-degree bank. The table in part A, Figure 6-68. shows the load factor at
various angles of bank. The graph (part B, Figure 6-68) shows rapid in-
a

bank increases. In the first 60 degrees


crease in the load factor as angle
of angle of bank, the load factor increases only by one. However, in the
next 10 degrees, the load factor increases almost one. At high angles of
A steep turn immediately after
bank, the load factor increases rapidly.
takeoff is extremely dangerous because of the load factor imposed and the low
velocity of the aircraft.

ANGLE LOAD 3.0


OF BANK FACTOR
A ~ n
B L
2.8

0 2.6
00 1.0 A
2.4
D
100 1.015 2.2
F
300 1.154 A 2.0
C 1.8
450 1.414 T
0 1.6
600 2.000
R 1.4
700 2.923 (Gs) 1.2
800 5.747 1.0
10 20 30 40 50 60 70 80
850 11.473
ANGLE OF BANK (DEGREES)
900 e><:::>

Figure 6-68. Load factors at various angles of bank

(2) The design strength of the aircraft must be considered. An


aircraft pulling 3 Gs is developing a lift force three times its weight.
Aircraft are designed to take certain load factors. The aircrafta is over-
stressed if these load limits are exceeded. If an aircraft has load limit
of 3 Gs, the minúnuID radius of turn would be at its 3-G bank angle, about 73
degrees (part B, Figure 6-68). This is the maneuvering speed of the
aircraft--the velocity at which the minimum radius of turn can be performed
for a given altitude without exceeding the load limits.
The preceding paragraph has added a structural limit to turning
(3)
performance. The aircraft mentioned above cannot make
a coordinated level
bank angle greater than 73 degrees because of the structural limits
turn at a

of the airframe (part B, Figure 6-68). A 90-degree bank has an infinite load
factor. However, the structural limitations of the aircraft make
a 90-degree
banked level turn impossible; also, at a 90-degree bank, no force develops to
90
oppose weight (Figure 6-69). In this case, the force would lift
be acting
degrees to the weight. Therefore, could not support the weight, and the
it
aircraft would not maintain altitude.
6-85
.
,,''''.''''''.::
f
:i~ji!f6jfi:.:~
i=U/;jjl!}J}Ji

LIFT

....;.",e,"'i',,:','
""'"

I'..,.. ....,....,)0- CENTRIFUGAL FORCE

,::U'!.'
Um., WEIGHT
y!;. +
(NO FORCE TO OPPOSE WEIGHT FORCE)

Figure 6-69. Aircraft unable to maintain altitude in a 90-degree banked turn

c. Power Limit of Performance. The third limit imposed on the turning


performance is the power, or thrust, limit, The induced drag developed at
high-load factors can become quite large. Induced drag is directly propor-
tional to the lift squared. Therefore, in a 73-degree bank--where three
times more lift
is produced than in level flight--induced drag is nine times
that of a level flight at the same velocity. This is a tremendous amount of
drag to overcome. The power available for the power plant is a limiting
factor.
6-36. VERTICAL TURNS

a. An aircraft can perform another type of turn--a vertical turn. This


is used during a pullout from a dive or during any vertical maneuvers such as
split-Ss, returns to target, or evasive maneuvers. Again, the aviator is
interested in the radius of turn and whether enough altitude is available for
a
maneuver.

b. Figure 6-70 shows an aircraft just passing level flight when the lift
is opposite the weight. Because the aircraft is turning about some point in
space, there is also centrifugal force. The radius of turn is the greatest
at the bottom of the pullout where forces are directly opposite the lift
force.
L

T 0(

,
,

f
CF

Figure 6-70. Vertical balance

6-86
6-37. RATE OF TURN

The rate of turn is used primarily during instrument flight and is simply the
aircraft heading rate of change. The angle of bank and the velocity are
important because they affect the rate of turn as well as the radius of turn.
Higher velocities and shallower bank angles result in slower changes in air-
craft headings. Rate of turn is at its maximum at the maneuvering speed and
angle of bank for the G-limits.

6-87
CHAPTER 7

ROTARY-WING FLIGHT TECHNIQUES

This chapter implements STANAG 3554


CEditio--- Jl Two).

Basic flying techniques described in this chapter genera ly apply


rotary-wing aircraft. The attitude-flying concept intro- ..luced and
to all
enlarged
during primary flight training promotes learning and est .blishes sound habit
patterns. This concept prepares aviators for transition into larger, more
complex aircraft. Aviators are then able to progress th ough instrument
flight training and are prepared for operational status ~ n
an aviation unit.
New aviators should be encouraged to study and use the b
.sic concepts of
attjtude contact flying and, later, those of attitude in ;trument flying.
They need to develop a working knowledge of how the airc aft components and
vital systems function. Aviators should also develop a ~orking knowledge of
the functions of the aircraft systems and components so hey can meet flight
performðnce requirements.

SECTION I

PLANNING CONSIDERATIONS

7-1. TRAINING

Aviator performance is built upon adequate knowledge, th .rough planning, and


the ability to project or predict what an aircraft will ~o. Coordination,
feel, and control touch are also important factors, but 4lhey are secondary to
the first three. Subject matter for aviation training, ~ccording to these
priDciples, is listed below. Emphasis should be on the ~ubject areas in the
order listed.

a. Academic Subjects. Aviators should possess--

(1) Knowledge of aerodynamics, physics, and mee -2banics of flight.


(2)
Specific knowledge of the systems, componeD- controls, and
str~ctures of the helicopter being used.
-ts,
(3) Knowledge of the methods and rules of
attit- ---ude flying, which
are similar to the rules of attitude instrument flying i --=n FM 1-240.

(4)
Specific knowledge of the breakdown of atti ~udes and cross-
checks-for each maneuver; development in dividing
atten~
checking outward from a specific center of attention
ion and cross-
each segment of
maneuver.
for--- a

7-1
b. Physical Application of the Controls. Physical application
of the
the initial stages of training than
controls is probably less important in
the other four subject areas in the preceding list. Physical skill is devel-
oped more rapidly after the aviator has mastered the first four subject
Aviators must learn what to do before they develop the physical
areas.
skills to perform a maneuver. The aviator needs to develop smooth and
coordinated physical application of control and the ability to hold attitudes
and power settings or to change attitude and power to perform maneuvers.
Descriptions of specific flight maneuvers for each Army aircraft are found in
the appropriate aircrew training manual and the operator's manual.

7-2. ATTITUDE FLYING

Aircraft attitude is the position of the aircraft in relation to the horizon.


Attitude is controlled, as shown in Figure 7-1, about three imaginary axes--
the longitudinal, the lateral, and the vertical. When an aircraft banks
(rolls), it changes attitude about the longitudinal axis. Attitude change
about the lateral axis is called pitch; this refers to raising or lowering
the aircraft nose in relation to the horizon. Yaw, or turning right or
During flight, an aircraft
left, is attitude change about the vertical axis.
can change attitude about one of these axes at
a
time. However, attitude
change frequently includes movement about more 'than one of the three prin-
cipal axes at the same time.

VERTICAL
AXIS
I
I "
r;;;
......'"
I ,I'
"
~t.j. .......
I ,
"".
I
I

00 ....
.....
.....
" I .....
.....
" I .....
.....
" I
~" .....
I ~.q)':
~~/
~..:> " ~-9.q~
o~(j~'ö
cp .q-t/.s-
" ~

Figure 7-1. Axes about which aircraft attitude is controlled

a. attitude of the aircraft in relation to the horizon and the power


The
applied are the only two elements of control in all aircraft. With these two
elements of control, the aviator can produce any desired maneuver, within the
flight envelope of the Therefore, all maneuvers must be based
aircraft.
solidly on attitude- and power-control references.

7-2
b. Aircraft attitude and power are modified by the aviator in two ways--
the time of application of an attitude or power change and the rate of change
of an attitude or power adjustment.

c. Keeping the basic control elements and modifiers in mind, the aviator
tracks what the aircraft is doing at any given time. With knowledge gained
from experience, the aviator can project what the aircraft is going to do
based on the power setting and the attitude that is being maintained. The
aviator then applies attitude and power changes smoothly so the aircraft
performs the desired maneuver. The result is attitude flying.

7-3. ATTITUDE CONTROL AND COORDINATED TURNS

During coordinated flight, turns are a result of bank attitude control about
the longitudinal axis of the aircraft. To hold a desired heading, the avia-
tor must keep the rotor disk laterally level in relation to the horizon.

a. Starting and Stopping a


Turn.
(1) To make a turn, the aviator banks (rolls) the aircraft about the
longitudinal axis until the rotor disk is tilted laterally. Rate of turn is
controlled by the degree the rotor disk is tilted. The aviator must learn
to bank the aircraft smoothly to a degree of lateral tilt that produces the
desired rate of turn.

(2) To stop a turn, the aviator smoothly rolls the aircraft to a


level position. Rollout is started before the desired heading is reached,
so the turn is stopped on the desired heading.

Vertical Lift. The aircraft shown in part A of Figure 7-2 is in


b.
straight and level flight; weight and lift are equal. For turning flight,
the aviator decreases the vertical component of lift by changing part of the
vertical lifting force toward the horizontal. The turn produces centrifugal
force, which tends to move the aircraft toward the outside of the turn, as
shown in part B of Figure 7-2. The resultant weight and centrifugal force
is outward and downward and is greater than the weight of the aircraft shown
in part A, Figure 7-2. This resultant weight and centrifugal force must be
overcome by adding to total lift. Otherwise, the aircraft will lose altitude
during a turn. Part C, Figure 7-2, shows an increase of the vertical com-
ponent of lift because the aviator has increased collective pitch and power.
Total lift equals centrifugal force and weight; the aircraft will now turn
without losing altitude.

7-3
VERTICALCOMPO
t CEN:R~:~~~CREASED.
VERTICAL COMPONENT
i.'.
"
I
OF LIFT INCREASED. I
I
I
,
I

CEN;~:C~GA~~
.
.

i FORCE ~
.:
.....
ff
j

J
O
I/- 0
Vt G

Figure 7-2. Loss of vertical lift during turns

c. Lateral Control. For lateral control of tandem-rotor helicopters,


the aviator increases sideward tilt of both rotors, which causes the fuselage
to roll about its longitudinal axis. Figure 7-3 depicts a left cyclic input,
which tilts the rotors equally to the left, causing the helicopter to roll
left or increase left lateral speed.

LEFT CYCLIC

FORWARD AND AFT ROTOR DISKS TILT LEFT; AIRCRAFT


MOVES TO THE LEFT.

Figure 7-3. Lateral control

7-4
d. Load Factors. Resultant weight and centrifugal force during turns
produce an increased load factor on the aircraft. The load factor is the
total load imposed on an aircraft, divided by the weight of the aircraft. It
is expressed in G units. The loa~ factor during a turn varies with the angle
of bank, as shown in Figure 7-4. Airspeed during a turn does not affect the
load factor, because for a given bank angle, the rate of turn decreases with
increased airspeed and does not change centrifugal force. For
a 60-degree
2 Gs, regardless of airspeed, as
bank, the load factor for any aircraft is
shown in Figure 7-4. This means that a 10,000-pound aircraft in a 60-degree
bank will, in effect, exert 20,000 pounds of force on the aircraft structure.
Bank angles up to 30 degrees produce only moderate increases in the load
factor, which are acceptable under most flight conditions. The load factor
rises at an increasing rate for banks over 30 degrees and may produce unac-
ceptable disk loading, depending on aircraft gross weight and varying flight
conditions.

7
'"
c,q,

o~
,~ 6

ú~
~~ 5
~
,,0
4

20 30 40 50 60 70 80 90

BANK ANGLE IN DEGREES-

Figure 7-4. Load factors in various angles of bank during level turns

7-4. ATTITUDE CONTROL AND AIRSPEED

Airspeed is controlled by adjusting the pitch attitude about the


a.
lateral axis of the aircraft. Therefore, the aviator must learn the aircraft
pitch attitudes that result in acceleration, deceleration, hover, and the
desired cruising airspeeds.

b. For a given power setting, there is only one pitch attitude and
airspeed that will maintain altitude. If power is constant and pitch atti-
tude is changed, increasing airspeed causes a loss of altitude. Conversely,

7-5
altitude is usually gained when airspeed is reduced with power remaining
constant.

c. If power is increased while pitch attitude is constant~ a constant


airspeed and climb result. If the power setting is decreased while pitch
attitude is constant~ airspeed remains constant and a descent results.

d. Differential thrust of forward and aft rotors is used for longitudi-


nal control of tandem-rotor helicopters. This creates pitch-attitude changes
of the fuselage about its lateral axis. Figure 7-5 shows a forward cyclic
(longitudinal control) input that decreases thrust (collective blade angles)
on the forward rotor and increases thrust on the aft rotor; the helicopter
pitches nosedown~ increasing forward speed. Equal but opposite input to the
forward and aft rotors is known as differential collective pitch.

INCREASED COLLECTIVE

lJ
DECREASED COLLECTIVE

~
Figure 7-5. Forward cyclic input

7-5. HEhDING CONTROL AND ANTITORQUE PEDALS

a. Antitorque pedals primarily counteract torque. Usually~ however~ the


antitorque system also has surplus thrust~ designed into the tail-rotor sys-
tem~ far beyond that required to counteract torque. This additional thrust
provides positive ànd negative thrust for taxi directional control and coun-
teracts crosswind effects on the fuselage during hovering operations. In
certain helicopter configurations~ the thrust power of the antitorque system
must be used carefully. Damage to the tail pylon area can result from over-
stress during fast-rate hovering pedal turns and during taxi conditions over
rough ground.

b. Proper use of antitorque pedals is required in three separate modes


of flight. Each of the following modes should be analyzed and treated
separately.

7-6
(1) During normal helicopter operations below 50 feet, the fuselage
needs to be aligned with a distant point. Maneuvers performed in this area
include taking off to and landing from a hover, the stationary hover, the
moving hover, the takeoff and climb slip control, and the approach slip con-
trol.
(2) Coordinated flight and all operations above 50 feet require
pedal use to align and rlold the fuselage into the relative wind.

(3) Coordinated turns at altitude require the proper use of pedals


to keep the fuselage into the relativè wind as the bank is initiated, estab-
lished, and maintained.

c. Aviators can counteract torque in tandem-rotor helicopters by tilting


the thrust vector of each rotor an equal amount in opposite directions
(cross-controlling).
(1) Aviators use differential lateral tilting of the rotor disk to
achieve heading control in tandem-rotor helicopters. When directional pedal
is applied, the forward rotor disk is tilted in the same direction; the aft
rotor disk is tilted in the opposite direction. The result is a hovering
turn around a vertical axis midway between the rotors.
(2) Figures 7-6 through 7-9 show combinations of cross-controlling
and the resultants. Trley are all examples of tandem-rotor helicopter control
characteristics.

7-7
-
RIGHT PEDAL
-

RESULTANT

RIGHT CYCLIC

Figure 7-6. Right pedal and right cyclic inputs

RIGHT PEDAL

-
-

RESULTANT

LEFT CYCLIC

Figure 7-7. Right pedal and left cyclic inputs

7-8
LEFT PEDAL
-
-

RESULTANT

LEFT CYCLIC

Figure 7-8. Left pedal and left cyclic inputs

LEFT PEDAL
-
-

RESULTANT

RIGHT CYCLIC

Figure 7-9. Left pedal and right cyclic inputs

7-9
(3) For heading control in forward flight, the aviator coordinates
lateral cyclic tilt on both rotors for roll control and differential cyclic
tilt on when
the rotors for yaw control. Only small pedal-trim changes are re-
quired longitudinal speed is changed or during climbs, descents, and
autorotations.
(4) If a tandem-rotor helicopter is rotated rapidly about one of the
rotors rather than about the center of gravity, the helicopter tends to pitch
because of the translational lift that one of the rotors achieves. The di-
rection of pitch depends on which rotor is being rotated about.

d. Below 50 feet, these guidelines for heading and track control should
be followed.
(1) For taking off to and landing from a hover, the aviator can use
references for heading control. Pedals should be repositioned to hold and
maintain the nose alignment with a distant fixed reference point. The avia-
tor uses an imaginary line to a distant object and applies pedal to position
and maintain a line of sight from his seat through the cyclic and gap between
his pedals, as shown in Figure 7-10. Aviators in either seat can use the
same distant fixed reference point. Fuselage alignment for hovering or take-
off direction is shown in Figure 7-11.

~~.
."-
"..
,
~
FIXED
REFERENCE
POINT

MAST

Figure 7-10. Heading alignment during hovering

(2) During a moving hover and initial climb to 50 feet, pedals


control the heading, as shown in Figure 7-11. Cyclic control is used for
direction and lateral positioning over the intended track, as shown in
Figure 7-12. Using peripheral vision and cross-check, the aviator positions
the helicopter with lateral cyclic so that the imaginary line is seen running
through position 1 during taxi or run-on landings (touchdown, takeoff, and
approach) and through position 2 for hover and climb through 20 feet; the
line should be seen between pedals, as shown at position 3, for all altitudes
over 20 feet. All track reference points should line up and pass between
pedals as each point is crossed.

7-10
(3) In crosswind operations~ the combined use of pedals and cyclic~ II
as described above~ results in a sideslip; cormnonly referred tö as "slip.
a

The aviator does not consciously think slip; he is automatically in true


a

slip if the fuselage is aligned on a


distant object with pedals~ as shown in
Figure 7-10. Positioning is also maintained over the line with cyclic
(Figure 7-12).

FUSELAGE ALIGNMENT TO HOVERING OR TAKEOFF DIRECTION

1/
FIXED
BUSH
REFERENCE
í' CORRECT
e POINT

LINE OF HOVER OR

DIRECTION OF TAKEOFF
)Þ .E3
@('..~)....
. -
\
.

TREE
. '.
.

Figure 7-11. Use of references for heading control below 50 feet

MAST CYCLIC LEFT PEDAL

RIGHT PEDAL

TRUE TRACK/POSITION
APPEARS TO BE-
FOR TOUCHDOWN,
TAKEOFF, AND APPROACH
BELOW 50 FEET AGL

FOR HOVERING AND


CLIMB THROUGH 20 FEET

FOR CLiMBOUT AND


APPROACH ABOVE 20 FEET AGL

Figure 7-12. Lateral positioning required to maintain desired ground track

and track control should


e. Above 50 feet~ these guidelines for heading
be followed.

7-11
(1) For coordinated flight above 50 feet, the pedals function in a
purely antitorque role; the aviator promptly repositions them to a climb
pedal setting when he reaches 50 feet. This pedal action aligns the fuselage
with the relative wind, rather than with a distant object.

(a) The helicopter is now in coordinated flight, during which


the cyclic controls the fuselage heading, the rotor disk is level laterally,
and the ball in the turn-and-slip indicator is centered.

(b) The track is now controlled by a


coordinated cyclic bank
and turn to heading that results in the desired track. Tracking toward and
a

over selected ground reference points causes these reference points to pass
directly under the aviator's seat cushion.
(2) Power changes require enough coordinated pedal to prevent the
fuselage from yawing left or right. When a power change is completed, the
aviator cross-checks the new pedal setting and lateral trim of the fuselage.
Because of counterrotating rotors, tandem-rotor helicopters require very
small, if any, pedal-trim changes with power changes.

(3) Generally, most single-rotor helicopters will have pedal


settings that are normal for various power-speed combinations. These
settings should be coordinated with power changes and the helicopter held
in cross-check for all operations and coordinated flight above 50 feet.

(4)Pedal-control linkage rigging varies in helicopters of the same


type. Therefore, in a steady climb, a cruise, a descent, or an
autorotation--with pedals set--the aviator should cross-check the following:
0

Turn-and-slip indicator for a centered ball. Pedal into the low ball;
note the exact pedal setting required when the ball is centered.
0
Door frames or windshield frames for lateral level trim. Pedal into
the low side; note the exact pedal setting required.
0
Main rotor in tip-path plane. It should be the same distance above
the horizon on each side. For level rotor, pedal into the low side.

(5) In semirigid main-rotor configurations, note the lateral hang of


the fuselage at a hover (into the wind). If the fuselage is not level, be-
cause of a lateral CG displacement, then the one-side-low condition must be
accepted as level; thereafter, in flight (air work higher than 50 feet), the
aviator adjusts pedals for a lateral trim of one-side low as existed at a
hover. Even though the fuselage is one-side low, the rotor is laterally
level to the horizon and the helicopter is in trimmed flight.

f. Use of pedal to enter and maintain a


turn requires study of and
experiment with the particular helicopter being flown.

7-12
(1) To determine if pedal is required for a coordinated entry to
bank and turn--

Start at cruise airspeed with the correct pedal setting for lateral
0

trim in straight and level flight.


0
Begin a bank with cyclic only. Use no pedal.

Note whether the nose turns in proportion to the bank.


0

the nose begins to turn as the bank is initiated, the aviator


(2) If
does not have to use pedal for the entry to
a
turn in this helicopter.

the nose does not begin to turn as the bank is initiated, the
(3) If
aviator uses only that pedal required to make the nose turn in proportion to
the bank and entry.

After the bank is established, the aviator anticipates the


(4)
normal requirement in all helicopters for
a slight pedal pressure in the
direction of the turn for coordinated flight and a centered ball.

7-6. POWER CONTROL AND RESULTING ALTITUDE, CLIMB, OR DESCENT

Altitude is the result of power control. To hold a desired altitude or make


changes of altitude, the aviator must apply power settings that produce the
desired climb or descent when coupled with the possible combinations of at-
titude and airspeed. Power settings that are normal for hover, climb,
The
cruise, slow cruise, and descent are used for precise altitude control.
aviator must also be able to adjust power to compensate for the variation in
atmospheric conditions as well as aircraft gross weight.
power setting that
a. For a given attitude and airspeed, there a is
a

maintains altitude. If a climb is desired with constant attitude and air-


speed, power must be increased above that required to maintain altitude. If
with constant attitude and airspeed, power must be re-
a descent is desired
duced below the power required for maintaining altitude.

constant altitude is maintained by minor pitch-attitude adjustments


b. A
and by power adjustments. After the altitude is stabilized and the desired
When
airspeed is established, whenever altitude changes, airspeed changes.
altitude again stabilizes, the airspeed returns to its previous indication,
if power is maintained at the previous setting. If airspeed is high because
of a
loss of altitude, excess airspeed, with an upward pitch-attitude adjust-
ment, can return the aircraft to the desired altitude and airspeed. Con-
downward
versely, if airspeed is lost because altitude is gained,
a

pitch-attitude adjustment can return the aircraft to the desired airspeed


and altitude.

aviator obtains collective control (thrust) of tandem-rotor


a
c. The
helicopter by increasing the collective blade angle of all blades the same
amount at the same time. Figure 7-13 depicts increases and decreases of
collective control inputs that cause vertical thrust and speed changes.
7-13
Because each rotor turns in an opposite direction, the torque effect is
canceled; the helicopter does not tend to yaw.

INCREASED COLLECTIVE

íS íS

DECREASED COLLECTIVE

ö ~

Figure 7-13. Collective inputs

7-7. TRAFFIC PATTERN

A
traffic pattern controls the flow of traffic, particularly at airports or
landing areas that are not radio-controlled. Patterns afford a measure of
safety, separation, and protection, as well as administrative control over
arriving, departing, and circling aircraft. During nontactical training, a
precise traffic pattern is flown to promote knowledge, planning, prediction,
and flight discipline. All pattern procedures must be strictly followed so
every aviator working in the traffic pattern and transient aviators arriving
and departing can determine at a glance the intentions of the other aviators.

a. When approaching a
radio-controlled airport, expedite traffic by
stating--
0
Call sign or aircraft serial number; for example, Army helicopter
16123.
0

Position; for example, 10 miles east.


0
Request; for example, for landing and hover to refuel point.

The tower will often clear an aircraft direct to an approach point or


b.
to particular runway intersection nearest the destination point. At uncon-
a

trolled airports, the aviator should adhere strictly to standard practices


and patterns.

c. Figure 7-14 depicts a typical nontactical traffic pattern with


general procedures outlined. If there is no identifiable helicopter traffic

7-14
pattern, the aviator sets up one inside the normal airplane pattern. He uses
touchdown and takeoff points to one side of the active runway. If he intends
to land on the runway, he approaches to the near end and then immediately
hovers clear of the runway.

d. To fly a
correct traffic pattern, the aviator should visualize a

rectangular ground track and--


0
Follow necessary outbound tracking on takeoff and climbout, with
steady climb airspeed.
0
Normally, turn less than 90 degrees for drift correction in the turn
to the crosswind leg, so as to track 90 degrees to the takeoff leg.
0

Select a
point on the horizon for the turn to the downwind leg, so as
to fly a
track parallel to the takeoff and landing direction. The
aviator should maintain constant airspeed and altitude.
0

Normally, turn more than 90 degrees for drift correction in the turn
to the base leg. The aviator should change attitude to slow cruise to
establish an approach entry airspeed. He changes power and pedals to
descend at about 500 feet per minute or to lose 5 miles per hour for
each 100 feet of descent. He watches for the reference point for a
turn to the final approach leg, as shown in Figure 7-15.
0
Turn short or beyond 90 degrees on the turn to final, depending on
crosswind conditions. Before entering the approach or not later than
the last 50 feet of the approach, he establishes a slip with the
fuselage of the helicopter on line with and over the line of
approach.

7-15

j;4
ø
"

~
$
~~

QO!lÇJ$'
$
II:
<C
~~
w
..J QO
U
C
Z
<C
Q.
0
I-
en

x-
{-
- - - -

...--- --"
-----x
PARK WHERE NO
AIRPLANE PARKING
POSSIBLE OR LIKELY.

't$o
OO"'~~-9"
í'IvO,.,0
I..~G

Figure 7-14. Typical nontactical traffic pattern

7-16
AVIATORS SHOULD WATCH THIS POINT TO
DECIDE WHEN, HOW, AND AT WHAT RATE TO
REFERENCE.

--
FAR
TURN FINAL. COMPLETE TURN AND ROll
POINT
lEVEL, WATCHING THIS POINT.

w
0
2
w
a:
æ
w I-
Z
a: Õ
tL

~ ~\V~v~
I ~~
-,\~~
~""' ~~~ ~.
~to~
I
~ ~~ ~(()~\O
/ to~ O"Q~'\ to
~ ~
~
/ ,\0 O\~
~
/
I
BASE LEG

8---------
~
SELECT AND USE THIS POINT FOR TIME-TO-TURN CUES.

Figure 7-15. Turn to final approach

7-17
SECTION
II
FIELD OPERATIONS

7-8. BASIC CONSIDERATIONS

In this discussion. a confined area is where the


a.
helicopter is limited in some direction by flight of the
terrain or the presence of
obstructions, either natural or man-made. For example, a
woods, the top of a clearing in the
mountain, the slope of a
hill, or the deck of a
ship
could each be regarded as a confined
area.
b. Generally, takeoffs and landings should be made
obtain maximum airspeed with minimum ground into the wind to
speed. However, exceptions to
the rule may occur.

c. Turbulence is defined as smaller masses of


direction contrary to that of the larger air mass. air in any moving
and the ground itself may Barriers on the ground
interfere with the smooth flow of air. This
interference is transmitted to upper air levels as
larger but less intense
disturbances. Therefore, the greatest turbulence is
usually found at low
altitudes. Gusts are sudden variations in wind velocity. These are normally
dangerous only during flight at very low
altitudes. The aviator may be un-
aware of a gust. Its cessation may reduce
lift, causing the aircraft to sink
or descend abruptly. Gusts cannot be anticipated.
However, turbulence can
normally be predicted. It exists during moderate
to strong wind conditions
in the following places:
0

Near the ground on the downwind side of


The turbulent area is always
trees, buildings, or hills.
relative in size to that of the obstacle
and relative in intensity to the
velocity of the wind. Figure 7-16
shows some turbulence
patterns.
0
On the ground and on the immediate upwind side
of any solid barrier
such as leafy trees and buildings. This
condition generally is not
dangerous unless the wind velocity is about 17
knots or higher.
0

In the air, over and slightly downwind of any


sizable
hill. The size of the barrier and the wind velocity barrier
a such as
height to which the turbulence extends. determine the

0
At low altitudes on bright,
sunny days near the border of two
dissimilar types of ground, such as the edge of a ramp
bordered by sod, as shown in Figure 7-17. or runway
This type of turbulence is
caused by the upward and downward passage
of heated or cooled air.

7-18
.... ...... ......

Figure 7-16. Air turbulence (building and trees)

,
::, r.::: ..'
'

oj ::".

.' .',';
. . .

"

.f",,",
00 "

.:.,'
00 ,
" 0

0,:~o~oo<
000

0
0

.
.
.
"
00
00

6. .
. .
. "

," 0
.
,

,.

Figure 7-17. Air turbulence (dissimilar ground)

7-19
7-9. RECONNAISSANCE

An aviator should conduct a high reconnaissance and a low reconnaissance


before landing in an unfamiliar area.

a. High Reconnaissance. A high reconnaissance determines--


0
A
point for touchdown.
0
Approach and departure axes.
0

Barriers and their wind effect.


0
Suitability of the landing area.
0
A
flight path for both approach ~nd takeoff.
Altitude, airspeed, and flight pattern for a high reconnaissance are governed
by wind and terrain features, including availability of forced-landing
areas.
The aviator must strike a balance between a reconnaissance conducted too high
and one conducted too low. The aviator should not fly so low that he has
to
divide his attention between studying the area and avoiding obstructions to
flight. On the other hand, he should not fly so high that he cannot study
the proposed area adequately. The ideal high reconnaissance is one low
enough to permit study of the general area and high enough to make a success-
ful forced landing in case of an emergency. A high reconnaissance is imprac-
tical during conditions that require terrain flight because the aviator would
have to climb; this would expose the aircraft to Threat forces for an unac-
ceptable period of time. .

b. Low Reconnaissance.

(1) Except when a running landing is necessary, a low reconnaissance


and an approach can often be combined. The aviator studies his approach path
and the immediate vicinity of his selected touchdown point when making an
approach. However, before losing ETL or before descending below the barrier,
the aviator must decide whether the landing and takeoff can be completed
successfully. The aviator should never land in an area from which a success-
ful takeoff cannot be made. A low reconnaissance should confirm what was
learned from the high reconnaissance.

(2) When
running landing is contemplated because of load or high-
a

density altitude conditions, a fly-by low reconnaissance is made. Airspeed


should be adequate to maintain ETL at an altitude that will clear
all obsta-
cles and allow the aviator to concentrate on terrain features. The intended
landing area should be checked for obstacles or obstructions in the approach
path or on the landing site. The point of intended touchdown must also be
selected.
(3) A
low reconnaissance can be conducted during an approach from
terrain flight. However, the landing area will normally be visible for only
a
short time before touchdown. More time for low reconnaissance is available

7-20
if a
circling approach from terrain flight can be made. Aviators must
determine whether a
circling approach will expose them to Threat forces.

7-10. CONFINED-AREA OPERATIONS

a. Approach.

(1) The confined-area approach begins with high reconnaissance. The


aviator should plan the approach by considering several different and some-
times conflicting factors. to determine wind conditions; he should
He needs
obtain the best possible advantage from them. He should consider the height
of barriers, identifying the lowest obstruction providing the best entry into
the area under favorable wind conditions. When possible, the aviator should
plan the flight path that places the helicopter within reach of the best
areas for a forced landing. When it is not possible to keep the area in
sight, the aviator should select specific reference points along the approach
path that will prevent him from losing sight of the area completely.

(2) The point of touchdown should be as far beyond barriers as


possible so the approach will not become too steep. The final stages of the
approach should be conducted short of downdrafts and turbulence, which may be
encountered at the far end of the area.

(3) The angle of descent should be just steep enough to clear


barriers, as shown in Figure 7-18.
(4)
aviator should terminate the approach to the ground
The when
surface conditions permit.

CONFINED AREA APPROACH


FROM HIGHER ALTITUDE

CLEAR BARRIER
BY A SAFE MARGIN.
/
, /" OR

l ~... TERRAIN FLIGHT


-

ALTITUDE
/

/'"

Figure 7-18. Confined-area approach and landing

7-21
Ground Operations.
b. Before the helicopter is operated within the
area, a
ground reconnaissance should be conducted to determine the if
is suitable. This reconnaissance can either be made from the cockpit areaon
or
foot.

c. Takeoff.
(1) The aviator positions the helicopter
for takeoff, taking
advantage of wind, barriers, and anticipated forced-landing areas on
takeoff.
(2) The aviator performs power and before-takeoff checks.
(3) As shown in Figure 7-19, the
aviator forms an imaginary line
from a
point on the leading edge of the helicopter, such as the
gear, to the
highest barrier that must be cleared. This line of ascent is flown using
only enough power to clear the obstacle by a safe distance.

CLIMB TO
USE POWER TO MAINTAIN
CONSTANT ANGLE OF ASCENT. ~IGHER ALTITUDE

BY A SAFE~MARGIN'
AND CLEAR OBSTACLE

,;II"
/ OR

MAKE TRANSITION

/-
~
Tõ TERRAIN FLIGHT
;""
""., ALTITUDE.
" ..
.--;..;.
N,"',J, .

;"" .
~~:"~..1f.
" .1""
..'-,(
.

Figure 7-19. Confined-area takeoff


(4) As the barrier is cleared, the attitude of the
helicopter should
be'adjusted to achieve normal climb airspeed and rate of climb.

7-11. PINNACLE AND RIDGELINE OPERATIONS

A
pinnacle is an area from which the ground drops away steeply on
as shown in Figure 7-20. A
all sides,
ridgeline is a long area from which the ground
drops away steeply on one or two sides such as a bluff or
precipice. Just
because pinnacle barriers may be absent does not mean pinnacle operations
are
easy. Updrafts, downdrafts, and turbulence may still be extreme hazards.
Landing areas may have barely enough room for a safe touchdown.

A climb
a. to a pinnacle or ridgeline should be executed on the upwind
side to take advantage of the updrafts, as illustrated in Figure
7-20. The
approach flight path should be parallel to a ridgeline and into the wind as
much as possible. The approach angle should be commensurate with the winds.
As a general rule, the greater the
winds, the steeper the approach needs to

7-22
be to avoid turbulent air and downdrafts. Ground speed during the approach
is more difficult to judge because visual references are farther away than
during approaches over trees or flat terrain. The aviator should avoid down-
wind turbulence; he should keep the helicopter within reach of forced-
a

landing area as long as possible. Load, altitude, wind conditions, and


terrain features determine the angle a to use in the final part of the
approach. If wind velocity creates hazardous crosswind landing condition,
he should make a low coordinated turn into the wind just before landing.

CAUTION: clear of downdrafts on the downwind side.


Remain Figure 7-20
illustrates the pinnacle approach.
b. landing on a pinnacle should take advantage of the long axis of the
A

area when wind conditions permit. Touchdown should be made in the forward
portion of the area. A stability check should also be made to ensure the
gear is on firm terrain that can safely support the weight of the helicopter.

c. pinnacle is higher than the immediate surrounding terrain,


Because a

The
gaining airspeed on takeoff is more important than gaining altitude.
higher the airspeed, the more rapid the departure from the slopes of the pin-
A higher airspeed enables aviators to cover unsafe ground quickly.
nacle.
Also, the higher airspeed affords a more favorable glide angle and improves
the chances of reaching a safe area in case of a forced landing. If no suit-
able area is available, with a higher airspeed the aviator can decelerate to
decrease forward speed before making an autorotative landing. The aviator
should not dive the helicopter d?wn the slope after clearing the pinnacle.
Diving the helicopter would result in a high rate of descent, which could
prevent a successful autorotative landing.

.
..-----
~
~
.

Figure 7-20. Pinnacle approach

7-12. TERRAIN-FLIGHT OPERATIONS

a. Takeoffs from confined or other remote areas can be made


Takeoff.
into terrain flight. This transition differs little from
a takeoff made into
flight at a higher altitude. Direction of takeoff is determined by current
Threat situations, wind, density altitude, and the long axis of the area. If

7-23
the Threat is a
factor, a takeoff should use available terrain and vegetation
for masking. If not, a takeoff is made into the wind over the lowest bar-
riers, taking advantage of the long axis of the area. As bårriers are
cleared, the aviator adjusts attitude and power as necessary to make the
transition into the desired terrain-flight altitude and airspeed. Also, the
aviator must consider updrafts and downdrafts that may be present, particu-
larly at terrain-flight altitudes.
b. Approach.Landings in confined areas can be made from terrain
flight; this transition differs little from landings made from a higher al-
titude. Because a high reconnaissance is not possible during terrain flight,
aviators rely more on flight planning to understand landing zone conditions.
Ground units in the LZ should be contacted for landing
instructions while the
aircraft is still several miles away. For a smooth transition into the ap-
proach, airspeed should be decreased from cruise to approach speed when the
landing zone comes into view. In NOE or contour flight, the approach angle
may be intercepted at a low altitude, leaving
little time to make the transi-
tion into the approach. Entry speed may have to be decreased before inter-
cepting the approach angle to prevent overshooting the landing zone.

c. Landing. Approaches to pinnacles or ridgelines will probably be


started at a lower altitude than the altitude of the landing zone. A climb
to the landing zone is made while maintaining a minimum of 40 knots airspeed.
As the landing zone is approached, the
aviator may need to reduce airspeed to
attain a safe rate of closure to the touchdown area.
7-13. SLOPE OPERATIONS

a. Limits. When a helicopter rests on a slope, the mast is perpendicu-


lar to the inclined surface; the plane of the main rotor must parallel the
true horizon or tilt slightly upslope (the rotor tilts with respect to the
mast). Normally, the cyclic control available for this rotor
by cyclic control stops, static stops, mast bumping, tilt is limited
or other mechanical
limits of control travel. These control limits are reached much sooner in
downslope wind conditions. When the hovering helicopter hangs with one side
low and is landing with the low side upslope, there is also
less control
travel. A slope landing site used previously may not be acceptable with a
different wind or CG helicopter loading. Conditions that permitted a slope
landing may have changed enough to cause hazardous conditions for takeoff,
for example, wind or a CG loading change.

b. Types of Operations.

(1) Approach. The approach to a slope may not differ


greatly from
the approach to another landing area. However, the slope may obstruct wind
passage, causing turbulence and downdrafts. Wind, barriers, and forced-
landing sites must be considered.

(2) Landing upslope or cross-slope. Brakes must be set before a


wheeled helicopter is landed. The landing is then usually made heading
up-
slope. This type of landing requires cautious and positive control touch.
The helicopter must be lowered from the true
vertical by placing the uphill
7-24
gear or skid on the ground first. The downhill gear or skid is then lowered
gently to the ground. Corrective cyclic control is applied at the same time
to keep the helicopter on the landing point. The aviator must maintain posi-
tive heading control on a forward reference point and normal operating RPM
until the landing is completed. To avoid mast bumping, sliding downslope, or
rolling over, the aviator should abort the landing attempt if cyclic control
travel limits are reached before the downhill gear or skid is firmly on the
ground.

(3) Landin~ downhill. Landing downhill is not recommended with some


single-rotor helicopters because the tail rotor may strike the ground.
(4) Landin~ uphill. If an uphill landing is necessary, landing too
near the bottom of the slope may cause the tail rotor to strike the ground in
some single-rotor helicopters. In this case and when landing downhill, the
mission may sometimes be completed at a low hover.

(5) Takeoff from a slope. To lift off from a slope, the aviator
moves cyclic control toward the slope and slowly adds collective pitch. The
downhill gear or skid must first be raised to place the helicopter in a level
attitude before lifting it vertically to a hover.
A
c. Dynamic Rollover. helicopter is susceptible to a
lateral rolling
tendency called dynamic rollover. This dynamic rollover can occur on level
ground; however, it is more likely to occur and more hazardous during slope
or crosswind landing and takeoff maneuvers. Each helicopter has a
critical
rollover angle beyond which recovery is impossible. If the critical rollover
angle is exceeded, the helicopter will roll on its side regardless of the
cyclic corrections made. The rate of rolling motion is also critical. As
the roll rate increases, the critical rollover angle at which recovery is
still possible is reduced. Depending on the type of helicopter, the critical
rollover angle may change based on which skid or wheel is touching the ground
(acting as a pivot point), crosswind component, lateral offsets in CG, and
left pedal inputs for torque correction (single-rotor systems).
(1) Characteristics.

(a) Dynamic rollover starts when the helicopter has only one
skid or wheel on the ground. That gear may become a pivot point for lateral
roll, as shown in Figures 7-21 and 7-22. When this happens, lateral cyclic
control response is more sluggish and less effective than. for a
free-hovering
helicopter. The gear may become a pivot point for a variety of reasons.
Most are aviator-induced. The gear or skid can become caught on objects
projecting from the landing surface such as a bent piece of steel planking;
it can possibly become stuck in soft asphalt or mud. Another way the gear
becomes a
pivot point is if the helicopter is forced into a slope by an im-
proper landing or takeoff technique. Whatever the cause, if the gear or skid
becomes a pivot point, dynamic rollover is possible when later aviator ac-
tions are incorrect.

(b) The tail rotor may add to this rolling tendency ifA cyclic
is not correctly applied to counteract lateral tail-rotor thrust.

7-25
cross-wind can also contribute to rollover by causing sideward drift or by
further accentuating the aircraft bank angle needed to land on a
slope.
(c) A smooth, moderate collective-pitch change may be the most
effectiveway to stop rolling motion. Collective must not be changed so fast
as to cause fuselage and rotor-blade contact. If a helicopter is on a slope
and the roll starts to the upslope side, reducing collective too fast can
create a high roll rate in the opposite direction. If collective reduction
causes the downslope gear or skid to hit the ground abruptly, the rate of
motion may cause a roll or pivot about the downslope gear.

(d) Suddenly increasing collective pitch to become airborne may


not stop dynamic rollover. If the gear or skid that acts as a pivot point
does not break free of the ground as collective is increased, the rollover
tendency will increase and worsen. If the skid or gear does break free of
the ground as collective is increased, it can cause an abrupt rolling move-
ment in the opposite direction because of the pendulum effect. This movement
can become uncontrollable.
(e) When performing maneuvers with one skid or gear on the
ground, the aviator should keep the helicopter trimmed, especially laterally.
Control is maintained if the aviator maintains trim, does not allow lateral
roll rates to become rapid, and keeps the bank angle from exceeding the
critical rollover angle for the helicopter. The aviator must take off
smoothly with only small changes in pitch, roll, and yaw. Untrimmed moments
must be avoided.

(2) Types of motion.

(a) A downslope rolling motion is caused when the aviator


applies too much cyclic into the slope. During landings or takeoffs when the
downslope skid is on the slope, the upslope skid may rise enough to exceed
lateral cyclic control limits. Thus a downslope rolling motion occurs.
(b) An upslope rolling motion is caused when the aviator
applies cyclic into the slope in coordination with collective-pitch applica-
tion. During landings or takeoffs when the upslope skid is on the slope,
the downslope skid may rise enough to exceed lateral cyclic control limits.
Thus an upslope rolling motion occurs.

7-26
DOWNSLOPE ROLLING MOTION
Downslope rolling motion Is caused by excessive
application of collective pitch In coordination with
cyclic application into the slope. When the
downslope skid is on the slope, applying too
much collective may result In the upslope skid þ.~EÞ.Of
rising enough to exceed lateral cyclic limits. A C~\"\~E~
\..

downslope rolling motion occurs. ~O\.\.

-- ----
.
Figure 7-21. Downslope rolling motion

FULL OPPO
TO PREVEN~'TECYCLIC
ROLLING
MOTION
UPSLOPE ROLLING MOTION

Upslope rolling motion Is caused by excessive


application of cyclic Into the slope, in
coordination with collective pitch application.
During landings or takeoffs, this condition results
In the downslope skid rising enough to exceed
laleral cyclic control limits. An upslope rolling
motion occurs.

Figure 7-22. Upslope rolling motion


(3) Prevention.

(a) Preventing upslope rollover during lift-off. Upslope


rollover characteristics are possible during lift-off. Upslope rollover
results from applying too much cyclic to hold the upslope gear against the
slope. If collective pitch is improperly applied, the aircraft then rapidly
pivots around the longitudinal axis of the upslope landing gear to the point
of rollover. To prevent upslope rollover, the aviator needs to cautiously

7-27
lift the downslope side of the helicopter to a level position, simultaneously
working the cyclic control to neutral. Once the cyclic is neutral and the
upslope landing gear has no side pressure applied, the aviator is cleared for
a
vertical lift-off to a hover and then to a normal takeoff.
(b) Preventing downslope rollover during landing. Downslope
rollover is caused when the helicopter becomes tilted beyond the cyclic-
control limits by the steepness of a slope. If
the slope (wind or CG condi-
tions) exceeds lateral cyclic-control limits, the mast forces the rotor to
tilt downslope. The resultant ~otor lift
has a downslope component, even
.

with full upslope cyclic applied. To prevent downslope rollover during land-
ing, the aviator slowly descends vertically to a light ground contact with
the upslope gear. While observing lateral, level reference frames, the avia-
tor pauses and maintains a positive-heading control. Then using careful
collective-pitch control, he slowly and cautiously lowers the downslope gear.
As the cyclic stick nears the lateral stop, he pauses to compare the distance
to go with the lateral control travel remaining (limits are given in the
appropriate operator's manual). If it appears the cyclic will contact the
upslope control, the aviator stops before the downslope gear is firmly on the
ground, returns the helicopter to a level attitude, and aborts the slope
landing. The aviator lifts off and moves a few feet for another attempt on a
lesser slope.

(c) Preventing downslope rollover during lift-off. After


landing inadvertently on an excessive slope, the aviator will attempt to lift
off. If the upslope gear tends to rise, the aviator should smoothly lower
the collective pitch. With full cyclic applied, however, the resultant lift
of the main rotor is not vertical or directed upslope enough to raise the
downslope gear. Therefore, if the upslope gear rises, the mast causes the
resultant rotor lift to move farther downslope. This increases the downslope
roll tendency, which continues to increase with added collective pitch. The
corrective action is to reduce power at the first sign of a lateral roll
around the downslope skid. Before another lift-off is attempted, appropriate
aviator action may be to--
a
Await different wind conditions.
a
Change CG loading.

a
Dig out 'from under the upslope gear.
a
-Notify operations to send a
recovery crew.

7-14. GENERAL PRECAUTIONS

Certain general rules apply to operations in any type of confined area,


slope, or pinnacle. The more important rules follow.

a. Know wind direction and approximate velocity at all times. Plan


landings and takeoffs with this knowledge in mind.

7-28
b. Plan the flight path, both for approach and takeoff, to take maximum
advantage of forced-landing areas.

c. Operate the helicopter as near to its normal capabilities as the


situation allows. The angle of descent should be no steeper than that needed
to clear existing barriers and land on a preselected spot. Angle of climb
should be no steeper than that needed to clear the barriers in the takeoff
path.

d. Iflow hovering is not hazardous because of terrain conditions, the


helicopter should be hovered at a lower altitude than normal when it is in a
confined area. This practice minimizes turbulence and conserves power. High
grass or weeds decrease efficiency of the ground effect; hovering low or
taking off partially compensates for this loss.

e. Make every landing to


a
specific point and not merely into a general
area. The more confined the area, the more essential it is to land precisely
on a definite point. The landing point must be kept in sight during final
approach, particularly during the more critical final phase.

f. Consider increases in terrain elevation between the point of original


takeoff and subsequent areas of operation.

g. Set the brakes on wheeled helicopters before initiating the approach


for a confined area landing, except for a running landing or when the landing
area is known to be level. This precaution precludes unexpected roll after
touchdown. A slope landing invariably results in a wheel roll unless brakes
are preset.

h. Judge the diameter clearance of main rotor blades before entering


a

confined area; remain alert to prevent possible damage to the tail rotor.
Not only must the angle of descent over a barrier clear the tail rotor of all
obstructions, but the aviator must also avoid swinging the tail rotor into
objects such as trees or boulders. The aviator must ensure that personnel
remain clear of the tail rotor at all times.

7-29
CHAPTER 8

FIXED-WING PERFORMANCE

This chapter presents the elements of applied aerodynamics that relate


directly to the problems of flying fixed-wing aircraft. For simplicity and
clarity. specialized mathematical detail on aerodynamics not applicable to
flight operations has been omitted.

SECTION I

AIRCRAFT DESIGN

8-1. CLASSIFICATION OF AIRCRAFT

Aircraft are classified by distinguishable features of their wings. power


plants. and landing gears and by their design purpose.

a. Components.

(1)
Win~s. The number, location, and design of wings help classify
an aircraft. There may be one, two, or more wings, although most aircraft
are single wing. A high-wing aircraft has the wing attached to the top of
the fuselage; a midwing, at or near the center of the fuselage; and a low-
wing, at the bottom of the fuselage. The wing may have the normal straight-
edge design; the sweptwing design, where the leading and trailing edges are
at an angle to the longitudinal axis; or the delta-wing design, where the
leading edge is swept back with the trailing edge forming the rear of the
aircraft.
(2) Power plant. An aircraft may be referred to by the type of
A
power plant used--reciprocating, gas-turbine, or jet. reciprocating engine
always drives a
propeller, which may further aid in classification. Aircraft
equipped with turbine engines that drive propellers are referred to as turbo-
props. Aircraft with turbine engines and without propellers are referred to
as jets.
(3) Landin~ ~ear. There are several different types of landing
gears. For land planes, they are either retractable or fixed. A conven-
tional landing gear has two main wheels, one on each side of the fuselage,
and a tailwheel. A
tricycle gear has two or more wheels, one on each side of
the fuselage, and one or more nosewheels. A bicycle gear has sets of two or
more wheels in tandem with the fuselage. Seaplanes use floats or the hull
for water operations. Amphibian aircraft are equipped for land and .water
operations.

b. Desi~n Purpose. The purpose for which an aircraft is designed


or used will also help classify it. Types of classification include

8-1
observation, cargo, utility, or trainer. Other classifications of military
aircraft do not apply to Army fixed-wing aircraft.
8-2. AIRCRAFT STRUCTURE

The principal structural units of an aircraft are the fuselage, wings,


control surfaces, and landing gear. Each has a specific function. Collec-
tively, they make up the airframe or aircraft structure.

a. Fuselage. is the main body of an aircraft to which


The fuselage
other structural units are fastened. It contains the crew and cargo, except
external stores; on single-engine aircraft it will usually also contain the
power plant. The three main types of fuselage construction are monocoque,
semimonocoque, and truss.

(1) MonocoQue. A
true monocoque construction--like a tin can--
consists of only the shell, with no internal bracing to help carry stress, as
shown in part A, Figure 8-1. Therefore, it requires heavy metal for the
shell and is not desirable because of its weight.

(2) SemimonocoQue. The semimonocoque--a modified monocoque--is


more suitable for military use, as shown in part B, Figure 8-1. Rings, bulk-
heads, and stringers inside the shell give it shape; these also carry and
distribute stress similar to the design of a tin can but with bracing inside.
The shell of stressed skin is fastened to the internal members and is usually
lightweight metal. Because stress is divided between the skin and internal
bracing, vital or critical points are mostly eliminated.

(3) Truss. The truss is used mostly on fabric-covered aircraft, as


shown in part C, Figure 8-1. The most common truss is the Warren type, which
consists of a rigid framework made of beams, struts, and bars welded together
to form triangles. Its main advantage is that truss members are subjected
only to tension or compression stresses.

8-2
STRINGER
RINGS RING
FRAME
BULKHEAD

A B
MONOCOQUE SEMIMONOCOQUE

Figure 8-1. Fuselage structural styles

b. Wings. Wings provide the lifting force that allows an aircraft to


fly. They also support the weight of the aircraft during flight. Design
depends on the size, weight, purpose, and desired speed for both flight and
takeoff of the aircraft. Figure 8-2 shows the two general types of wing
construction--internally and externally braced cantilever and semi cantilever.
Each type may be covered with cloth or stressed skin. Stressed-skin wings
distribute the load over more of the wing area. Therefore, they can carry
more loads or stress without failing. Most military aircraft are constructed
with stressed-skin wings.

~MI~~- -

~U~-
Figure 8-2. Wing construction

c. Control Surfaces.
Control surfaces are constructed on the same
principle as an airfoil. Normally, they are covered with metal and can be
operated manually, mechanically, hydraulically, or electrically.

8-3
d. The landing gear consists of wheels, shock absorbers,
Landin~ Gear.
and possibly retracting mechanism. Most small aircraft have a fixed gear.
a

However, larger and faster aircraft usually have a retractable landing gear,
which cuts down on overall drag and stress while in flight.

SECTION
II
HIGH-LIFT DEVICES

8-3. PURPOSE

An aircraft's
low-speed characteristics can be as important as its high-speed
performance, if not more so. Army aviators spend much of their time in the
air below 3,000 feet and at airspeeds less than 150 knots because takeoffs
and landings are made at relatively low altitudes and primarily involve low
speeds. For this reason, aircraft designers must turn to "high-lift de-
vices." Hi~h-lift devices increase the maximum value of the coefficient
of lift (CLmaximum) by various means.

a. Lift
Force. The term "high-lift device" is somewhat of a misnomer.
A device is not used to increase lift but to obtain a required lift
high-lift
force at lower velocities. For example, an aircraft flying at 250 knots is
developing 10,000 pounds of lift. When landing, the aircraft still requires
10,000 pounds of lift; however, it might now be flying at 100 knots. Because
the landing approach speed of the aircraft is a function of the stall speed,
reference to the following stall-speed equation readily shows how high-lift
devices can lower landing and takeoff velocities:

L =
Constant -
1/2 p V2 S CL (Equation 8.1)

b. Stall Speed. The slowest velocity that an aircraft can fly depends
on the maximum value of CL attainable. This is shown in the stall-speed
equation. The stall speed is inversely proportional to the square root of
the value of CLma~imum' If
this value is increased, then the stall speed is
lowered or a greater weight can be supported with the same stall speed. In-
creasing the payload of an aircraft is another example of when high-lift
devices are required. All high-lift devices increase the value of CLmaximum'
The two most common ways to increase the value of CLmaximum are by increasing
the camber of the airfoil or by delaying the boundary-layer separation.

8-4. INCREASING THE COEFFICIENT OF LIFT

Increasin~ Camber. Of the two usual methods of increasing


a.
CLmaximwm' increasing the camber of the airfoil is most often used.

(1) A
wing with more camber has a greater velocity differential
between the top and bottom surfaces of the wing. This greater velocity dif-
ferential creates a large pressure differential across the wing. The pres-
sure differential has been previously related to the value of CL for a given

.
8-4
~ngle of attack. Therefore, by increasing the camber of an airfoil, the
value of CL is increased.

(2) The usual method of increasing the camber is through the use of
trailing-edge flaps, as shown in Figure 8-3. This CL curve is shown for
flaps up and flaps down. The basic airfoil is a symmetrical airfoil, and the
wing has its zero-lift point at an angle of attack of zero degrees. With the
flap extended, the airfoil now has a positive camber and the zero-lift point
has snifted to the left. The value of CLmaximumhas increased. The CL curve
of the basic wing shifted up and to the left as the flaps were lowered. In
this manner, all high-lift devices that increase camber effect an increase in
the value of CLmaximum' The angle of attack at which the wing will stall has
been decreased. The basic wing stalled at an angle of attack of about 18
degrees; with increased camber it stalls at 15 degrees. However, the value
of CLmaximum at 15 degrees (flaps down) is greater than at 18 degrees on the
basic wing.

~ Ð G
SAME WING WITH
TRAILING-EDGE FLAPS

-"--

li
. BASIC WING
I I
I I
I I
I I
: I I
I I
I I
I I
I I
I I
I
I I

Ð a
I

150
I

18"

Figure 8-3. Increasing camber with trailing-edge flap

b. Delaying Boundary-Layer Separation. The other ~ommon method of


increasing the value of CLmaximum is by delaying boundary-layer separation.
The maximum value of CL is limited by the separation of the boundary layer.
The basic wing mentioned before stalled at an 1S-degree angle of attack. If
tne energy level of the boundary layer over the wing could be increased, then
the wing could be rotated to higher angles of attack before a stall would
occur. This technique uses boundary-layer control to increase the value of
CLmaximum' The energy level of the boundary layer can be increased by suc-
tion or blowing BLC or by vortex generators. Figures 8-4, 8-5, and 8-6 show
these methods of boundary-layer control. Figures 8-4 and 8-5 also include
the CL curves with the results of BLC. The slight difference between suction
and blowing BLC, because of airflow cnanges on the surface of the wing, is
not considered in this text.

8-5
(1) Suction boundary-layer control. The
suction BLC in Figure 8-4
draws off the low-energy, aerodynamically dead and turbulent
boundary layer, causing the higher-energy air below the
layers above to be lowered closer
to the airfoil surface. This makes the
airfoil effective at angles of attack
where it previously stalled. Part A of Figure 8-4 shows
boundary-layer sepa-
ration when the airfoil is stalled at an 18-degree angle of attack with BLC
off. In part B of Figure 8-4, the a~rfoil is also at an 18-degree
attack; however, with suction BLC on, boundary-layer separation no angle of
longer
occurs. Suction BLC is rather inefficient.
It requires a heavy vacuum pump
or turbine to handle the large volume of air being drawn off the
This increases the weight of the aircraft; the airfoil.
extra weight partially offsets
the advantages gained by the increased value of
CLmaximum'

FLAPS DOWN
BLC ON
0 BOUNDARY lAYER
~
: I
a. =180 II FLAPS
STALLED DOWN
Y
~ Q~!
BlC OFF

BlC OFF
AERODYNAMICAllY
C) ) ':--
\.9lt:>
~ Cl
I
I
I
I
I
I
DEAD AIR (TURBULENT)
"- I I
Holes in top surface J I
of flap connected to J J
vacuum pump. I I
a. =180 I I
NOT STALLED I I
I I
VACUUM
PUMP BlC ON
a. 18" 210

0.-- ANGLE OF ATTACK

Figure 8-4. Suction BLC


(2) Blowing boundary-layer
control. Blowing BLC, as shown in
Figure 8-5, increases the energy level of the boundary
layer by introducing
high-energy air through a nozzle, usually mounted ahead of the
method can be thought flap. This
of as blowing the turbulent air from the top
the surface of
airfoil. Blowing BLC is more commonly used than suction BLC, because it
is more efficient. The compressor section of a turbine engine
can be used to
supply the high-energy air needed. Therefore,
aircraft weight does not
increase.

8-6
FLAPS DOWN
~
I
BlC ON
I
I FLAPS DOWN
BlC OFF

BOUNDARY lAYER
REENERGIZED C1 180 210

Figure 8-5. Blowing BLC

(3) Boundary-layer control by vortex generators. Another method of


reenergizing the boundary layer is to use vortex generators, as shown in
Figure 8-6. These are small strips of metal placed along the wing, usually
in front of the control surfaces or near the wing tips. The turbulence
caused by these strips mixes high-energy air from outside the boundary layer
with boundary-layer air. The effect of the vortex generators on the lift
curve is similar to other boundary-layer control devices.

Figure 8-6. Vortex generators

8-5. TYPES OF HIGH-LIFT DEVICES

In the following paragraphs, various types of high-lift devices are


discussed. These devices increase either the camber of the airfoil or
the energy of the boundary layer.

8-7
a. Trailing-Edge Flaps. Trailing-edge flaps are the most common type
of high-lift device. These flaps have both advantages and disadvantages. A
trailing-edge flap increases the camber of the wing, thereby increasing the
value of CLmaximum. However, in so doing, it moves the lift force toward the
trailing edge of the wing, resulting in a negative or nosedown pitching mo-
ment. This moment limits the use of flaps to aircraft having horizontal
stabilizers and elevators. When a trailing-edge flap is extended, the angle
of incidence is increased because the chord line of the airfoil changes. As
shown in Figure 8-3, the change in angle of incidence changes the zero-lift
line. The nosedown pitching moment on the fuselage results in better forward
visibility during landings and takeoffs, as shown in Figure 8-7. Flaps also
increase drag on the aircraft. This is useful in landing; the aircraft can
make a steeper approach without increasing airspeed. However, this drag
increase is not desired on takeoff. Most aircraft having large and effective
trailing-edge flaps use only partial flaps on takeoff; thus they have the
benefit of increased CL without a large increase in drag. Some of the common
types include the following: the plain flap, the split flap, the Fowler
flap, the slotted flap, and the slotted Fowler flap. These are illustrated
in Figure 8-8, parts A through E.

LONGITUDINAL AXIS

CHORD LINE
LONGITUDINAL AXIS
-. -..!é'./
A'

RELATIVE WIND

Figure 8-7. Angle-of-incidence change with flap deflection


(1) Plain and split flaps. These two basic flaps are shown in
Figure 8-8, parts A and B. Both increase the camber of the airfoil; the
split flap does not produce as large a nosedown pitching moment as the plain
flap. The split flap also creates a greater drag force because of the 10w-
pressure, high-turbulence area between the wing trailing edge and the flap
trailing edge.

8-8
PLAIN FLAP

~O
.

~
.

~(///: 'ZA~
SPLIT FLAP
~

CAMBER INCREASE

~~ 0
----

~~..- -~~õ
~

FOWLER FLAP

~~)
~--- -~~
SLOTTED FLAP

~~.
..--
~
SLOTTED FOWLER FLAP
)i
-
.

~
o.~
~~
0
~
AREA INCREASE

~~C":,.~INCREASE
~
;)0:<),

~~)
~

LEADING-EDGE FLAP
~ ..
0

~~
~~
LEADING-EDGE SLOT

~~~
MOVABLE LEADING EDGE
-

. -
0

=. ~DS:ATlDN

-~~-
~'~-'-' '-'-'~.~ _0
~

(2) Fowler
~

Figure 8-8.

flap.
~

On
-

Types of

lift heavy loads from short C


aircraft that
high-lift devices

fieldst jet transportst


such as the Fowler flapt which is shown in part of
Figure 8-8t is often used. When extended, this type of flap moves rearward
as well as down. This increases CLmaximum because of an increase in camber
and wing area. The Fowler flap then reduces Vs (stall speed) by increasing
CLmaximum and wing area. Although aerodynamically the Fowler flap is the
most efficient flapt it does have disadvantages. With the huge surface ex-
tending so far behind the wing, a
large twisting moment is set up in the
wing. Therefore, the wing must be strong enough to withstand the load. The

8-9
increased structural strength and the more complicated actuating mechanisms
account for large increases in weight and internal wing volume. The Fowler
flap cannot be used on thin, high-speed airfoils.
(3)
Slotted flap. To increase efficiency, most flaps can be
slotted. Using slotted flaps combines the principle of boundary-layer con-
trol with a camber change. Together, the effects are cumulative. A plain
flap curve before and after the slot is added, as shown in Figure 8-9. After
adding the slot, separation over the flap area is delayed so the wing can be
rotated to a higher angle of attack. The increased energy required to delay
the boundary-layer separation comes from the low-velocity and high-pressure
air under the flap. The air is directed through the slot over the top sur-
face of the flap. This increases the energy of the boundary layer over the
flap. A slotted Fowler flap is even more efficient. A slotted Fowler flap
increases both camber and wing area. In fact, the CLmaximum value of a
multiple-slotted Fowler flap may be twice that of the basic wing in part E of
Figure 8-8.

PLAIN SLOnED
FLAP

CL

CI

Figure 8-9. CLmaximum increase with slotted flap

b. Leading-Edge Devices. The leading-edge flap, the leading-edge slot,


and the movable leading edge are all leading-edge devices. These are illus-
trated in Figure 8-8, parts F through H.

(1) Leading-edge flap. Some aircraft use leading-edge flaps, as


shown in part F, Figure 8-8. These increase the camber of the airfoil to
increase CLmaximum' Unlike the trailing-edge flap, the leading-edge flap
does not produce a negative pitching moment. However,
it may create a slight
positive (nose-up) pitching moment, depending on its effectiveness.

8-10
(2) Slots and slats. Boundary-layer control devices have both
advantages and disadvantages. They are usually used with camber-changing
devicest because BLC alone is not as effective as camber change. Suction
and blowing BLC devices and vortex generators have already been mentioned;
howevert the leading-edge and movable leading-edge slots are also forms of
boundary-layer control. They are shown in parts G and H of Figure 8-8.

(a) The slot through the wing (part G) will vent high-pressure
air from the underside of the wing over the top surface. This delays a stall
when the wing is at a high angle of attack. Because the slot is not exposed
to the airstream, drag does not increase much.

(b) Most modern carrier aircraft have movable leading-edge


slats, as shown in Figure 8-8. This is simply a slot that can be opened and
closed. When opened at high angles of attack, the slat moves forward (some
also move downward), increasing the camber and area. This occurs when the
angle of attack is high--whether at low or high speeds--during high-G maneu-
vers. A BLC device does not create any pitching moment. Therefore, aircraft
without horizontal stabilizers use this type of device; it allows aircraft to
rotate to higher angles of attack. Howevert aircraft can be rotated to such
a high degree that the aviator has difficulty seeing the landing area. This
limits effective use of devices that delay boundary-layer separation.

SECTION
III
STALLS

8-6. CHARACTERISTICS

a. In the early years of aviationt the advice was to "fly low and
slow. II Because this condition affords a minimum distance to fall, it seemed
to be sound reasoning. Actually, it is probably one of the most dangerous
conditions of flight. To produce required lift at slow airspeedst aviators
must fly at a high angle of attack--near the angle of attack for the aero-
dynamic stall. When this stall occurst lift decreases and drag increases;
there is almost always a loss of altitude. In addition to loss of altitude,
there can also be a loss of control. Under these conditions, the aircraft
can enter a spin. A
considerable loss of altitude is possible before control
of the aircraft can be regained.

b. Takeoffs and landings involve low airspeeds and altitudes. This


combination makes them hazardous phases of flight. One of the most frequent
causes of takeoff and landing accidents is the stall. When a stall occurst
there is usually insufficient altitude for recovery. Because each flight has
at least one takeoff and landingt an aviator must be prepared to operate the
aircraft at slow speeds and at high angles of attackt a potentially hazardous
configuration. Knowing thist aviators must thoroughly understand stall char-
acteristics. This section defines the stall and discusses the causes, warn-
ings, and characteristics.

8-11
8-7. AERODYNAMIC STALL

An aerodynamic stall occurs when an increase in the angle of attack results


in a decrease in lift coefficient. This is due to separation of the boundary
layer (a thin layer of air near the surface of the wing) from the upper sur-
face of the wing. When this boundary layer separates, turbulence occurs
between the boundary layer and the surface of the wing. This causes static
pressure on the upper surface of the wing to increase. The definition of
stall makes no reference to airspeed. The only thing that causes a stall is
an excessive angle of attack.

a. Stall Angle of Attack. In Figure 8-10, all angles of attack greater


than the angle of attack for the maximum lift coefficient fit the definition
of the stall. An increase in the angle of attack beyond the angle of attack
for CLmaximum (14 degrees) decreases the value of CL' The crosshatched area
is called the stall region. When the aircraft operates at an angle of attack
within this region, it is stalled, whether its airspeed is 60 or 160 knots.

1.6

1.4

STRAIGHT PORTION
BOUNDARY-LAYER
1.2
SEPARATION POINT
t CONSTANT
:; 10
.

~
0
I-
ffi
.8
Õ
ii:
Ifi
.6
0
0

.4

.2

0 2 4 6 8 10 12 14 16 18

ANGLE OF ATTACK (0) STALL ANGLE


OF ATTACK

Figure 8-10. Coefficient of lift curve

8-12
b. Cause of the Stall.
(1) The cause of the stall is relatively easy to understand. The
wing or airfoil is designed with certain camber to give
a definite pressure a

differential between the top and bottom surfaces. As the angle of attack
increases, the CL increases because of an increased pressure differential.
At all angles of attack that correspond to the straight portion of the CL
curve to the left of the stall region, the airflow follows the curvature of
the top surface until it almost reaches the trailing edge. At that location,
a small turbulent wake is formed, as shown
the boundary layer breaks away and
in Figure 8-11.

BOUNDARY -
LAYER SEPARATION POINT

~m1~~~if
-Þa.
20
Þ. 120
..
140
..
200
. .

'-;I

Figure 8-11. Various airfoil angles of attack

(2) The point where the boundary layer separates from the airfoil
stays essentially constant as long as the angle of attack is of value where
a

0 12 degrees, as shown
the CL curve is a
straight line (between degrees and
in Figure 8-10). If the angle of attack increases beyond the straight por-
tion of the CL curve, the point of boundary-layer separation moves forward.
This actually decreases the top surface area of the wing that is producing
lift. The airflow under the boundary layer is turbulent. Therefore, in that
area, the static pressure is increased, compared to the area where no separa-
tion occurs. The increase in the angle of attack increases the pressure
differential on the portion of the wing where no separation exists. This
increase in the pressure differential is partially offset by the loss of some
of the effective area of the wing. This results in a
smaller increase in the
CL' per degree increase in angle of attack. The slope of the CL curve de-
creases and continues to decrease. As the angle of attack increases, the
separation point of the boundary layer continues to move forward. Finally,
a

further increase in the angle of attack results in a


decrease in the value of
the CL. The point where the boundary layer separates has now moved too far
forward; the loss of the effective area of the wing is too large to be offset
by any increase in the pressure differential that may occur. This is the
angle of attack that is defined as the stalling angle of attack. At the
angle of attack for CLmaximum' the slope of the CL curve has reached zero.
Any further increase in the angle of attack develops a negative slope to the
curve (CL decreases as angle of attack increases).

8-13
(3)
The airfoils shown in
Figure 8-11 can be compared to the CL
curve in Figure 8-10. At an angle of
attack of 12 degrees, the curve slope
starts to decrease and boundary-layer
separation starts.
when the boundary
layer lacks the energy to adhere to the Separation results
all the way to the trailing edge. In other surface of the wing
to the sharp bend. When placed words, the airflow cannot conform
at 90 degrees to the airstream, the
plate shown in Figure 8-12 has a turbulent flat
boundary layer will not remain on the
flow behind it. Therefore, the
back of the plate. The same is
true of the wing at high angles of attack. surface
boundary layer no longer There is a
limit where the
remains on the surface of the
the point of boundary-layer wing. That limit is
separation.

Figure 8-12. Boundary-layer separation


8-8. STALL WARNING

a. Aerodynamic Stall Warnin~.


(1)
The turbulent airflow generated when
separates is a sign of an impending the boundary layer
part of the aircraft, it. causes stall. As this turbulence
flows over
buffeting, which is normally felt in the
controls. This notifies the aviator of an
flow is generated before the approaching stall. Some turbulent
which can occur before the
stall actually occurs. Therefore, buffeting,
aircraft actually stalls, is a warning.
(2)Part of the aircraft behind the wing is the
stabilizer. The turbulent flow can pass horizontal
over it to give the warning. The
span of the horizontal stabilizer
is less than the span of the wing.
fore, any turbulent flow coming from the wing There-
wing would not flow over the tips or outer portions of the
wing so the root section stabilizer. This is one reason to design the
stalls before the tip section. The
flow then creates the turbulent air-
aircraft buffet warning before the entire wing is
stalled. Also, when one part of the wing
is not as. abrupt as the entire wing stalls before the other, the stall
stalling at once. Aviators have better
lateral control of an approaching stall
of the wing to the tip on if the stall progresses from the
aircraft with ailerons located toward the wing root
tips.

8-14
(3) Although a root-to-tip stall pattern is
desirable, it is not
always possible to achieve. A rectangular or slightly tapered wing
normally
stalls root first. However, highly tapered, swept, or delta wings exhibit a
strong tendency to stall tip first. Several design techniques can make the
root stall before the tip.

(a) Geometric twist. One method of causing the root to stall


is geometric twist (washout); that is, building a twisted wing. The root
section angle of incidence is greater than the tip section. The twist is
about 3 degrees. If an airfoil section has a stalling angle of attack of
18 degrees, the root section is
at an 18-degree angle of attack when it
stalls. However, the tip section is still at about a ls-degree angle of at-
tack and is not stalled. Thus even if there is aerodynamic buffeting from
the turbulent air around the root section, the aviator can still use ailerons
for lateral control during recovery.
(b) Aerodynamic twist. Another method of stalling the root
section before the tip section is aerodynamic twist. A wing with aerodynamic
twist is not really twisted as with geometric twist. However, the wing re-
acts in the same manner and is said to be twisted. In this case, the air-
craft designer uses two more types of airfoils. In Figure 8-13, the CL curve
for the cambered airfoil and the CL curve for the symmetrical airfoil have
about the same value of CLmaximum; but the angle of attack at which they
attain their CLmaximum is different. In this case, the root section is a
cambered airfoil; toward the tip, the wing will gradually transform into a
symmetrical airfoil. The angle of incidence is the same for both sections.
Therefore, there is no geometric twist to this type of wing. The stall pro-
gression from root to tip is controlled aerodynamically by using different
types of airfoils. If a wing is constructed with the airfoil sections plot-
ted, as shown in Figure 8-13, the root will stall at a ls-degree angle of
attack while the tip would not stall until an 18-degree angle of attack is
reached, as indicated by the curves in the figure.

8-15
1.8

1.6

1.4

1.2

1.0
...
CJ

.8

.6

-4 -2 0 2 4 6 8 10 12 14 16 18 20

ANGLE OF ATTACK a

Figure 8-13. CL curves for cambered and symmetrical airfoils


(c) Stall strip. A third method for stalling root
sections
first or at least creating a buffet on the aircraft is to use a stall strip
on the leading edge of the wing, as seen in
Figure 8-14. This causes the
boundary layer to break away from the
airfoil at an angle of attack lower
than the stalling angle of attack for that
airfoil. The cruise speed, design
load, and general performance requirements of an
aircraft determine the air-
foil section to be used. These design considerations may preclude the use
of twist methods for smooth stall progression. A
stall strip, located in the
root section, detaches the boundary layer to ensure that this
first. This gives adequate warning to allow for a safe recovery with astalls
section
mum loss of altitude. mini-

STALL STRIP

Figure 8-14. Stall strip

8-16
Mechanical Stall Warning. Some aircraft do not have horizontal
b.
stabilizers or, as in the C-12, are designed so that thebyhorizontal stabili-
zer is not in the path of the turbulent wake generated the wing as it is
for the
stalling. These aircraft usually have mechanical a stall warning
a

mechanical stall warning is flapper switch mounted


aviator. The simplest As the wing ap-
on the leading edge of the wing, as shown in Figure 8-15.
switch.
proaches a stall, the relative wind pushes the flapper up, closing
a

the aviator of an impending


This, in turn, activates the device to warn
of attack at which
stall. The flapper can be positioned to vary the angle
the stall warning will occur.

. ~
.
'\,\\r
~_.
~
~
~_._._--_.

Figure 8-15. Flapper switch

8-9. STALL RECOVERY

To recover,
When a warning is received, recovery should be immediate.
stall The only action
the aviator corrects the cause--too high an angle of attack.
This breaks the
the aviator must take is to decrease the angle of attack.
stall, stopping the stall warning immediately.
8-10. STALL-SPEED EQUATION

the basic
This section has been dealing with the angle of attack as
a. the speed
factor affecting the stall. The stalling speed of an aircraft is
at which, for a given set of conditions, the aircraft is at its stalling of
angle of attack. Although the stalling speed varies, the stalling
angle
attack remains constant for any particular airfoil shape.

Certain assumptions must be made to develop


b.
a
stalling speed
equation. First, this section deals only with an aircraft in level flight
which
witb tbe lift vector opposite the weight vector. Climbing fligbt,
Second, the aircraft is
affects stall speed, is discussed in paragraph 8-12. law
considered to be in equilibrium, which allows the use of Newton's first
the of the forces about the center of gravity of the aircraft
of motion; sum
is equal to zero. At this time, only the vertical forces are important;
up

forces must equal down forces. With the aircraft in level flight,
the lift
8-17
equals the weight of the aircraft, as shown in Figure 8-16, and in the
~ollowing lift equation:
L = W =
1/2 P V2 S CL
(Equation 8.2)

LIFT

..
THRUST
~ DRAG
)'

!
WEIGHT

Figure 8-16. Aircraft in equilibrium

c. The lift equation shows that slower velocities require higher angles
of attack to produce higher values of CL so that
lift equals weight. If the
projected area of the airfoil surface (8) and the airstream density ( p) are
assumed to be constant, the minimum of velocity (V) depends on the maximum
value of CL attainable. This value (CLmaximum) occurs at the stall angle of
attack. If an aircraft is operating at the angle of attack for CLmaximum' it
is also operating at its minimum velocity (Vminimum). This is the minimum
velocity at which the aircraft can maintain level flight. This is the stall
speed of the aircraft.

d. To develop the stall-speed equation, the following lift equation is


first solved for velocity:
V -
-

V 2L
CL P 8 (Equation 8.3)

If the lift force is considered equal to the weight, a


direct substitution
can be made, as in the following equation:
V -Vc2W
-

S (Equation 8.4)
LP
In this form of the equation, the velocity for any particular value of CL can
be found. To define the power-off stall speed (Vs)' the value of CL is fixed
as CLmaximum and the equation takes the following form:

V
s
-
-

V 2W

CLmax P
8 (Equation 8.5)

This is the level flight stall-speed equation.

8-18
e. The level flight stall-speed equation shows how weight, altitude,
and configuration affect the stall velocity. These factors are discussed in
the following paragraphs.

(1) Weight. The stall-speed equation shows that changes in weight


vary the stalling speed of an aircraft. As an aircraft flies, its weight
decreases because of fuel consumption. This decrease in weight also de-
creases the stalling speed, because the stalling speed is directly propor-
tional to the square root of the weight. For example, an aircraft that 86 knots
weighs 18,000 pounds at takeoff and stalls at 96 knots will stall at
with a weight of 14,000 pounds at the end of a
flight. As mentioned, this is
due to fuel consumption. However, using fuel is not the only way an aircraft
decreases its weight in flight. Jettisoning of external loads can decrease
aircraft weight enough to change the stalling speed appreciably. The weight
factor--wing loading--is expressed as wIs. Wing loading represents the aver-
age amount of lift required of each square foot of wing area. As weight is
increased, wing loading is increased; the lift required of each square foot
of wing area also increases. Wing loading appears in the stall-speed equa-
tion; the stall speed is directly proportional to the square root of wIs.
(2) Altitude. Air density decreases as altitude increases.
Because air density ( p ) appears in the denominator of the stall-speed equa-
tion, an increase in altitude also causes an increase in the stall velocity.
Previous references to stall speed have been to true airspeed (V). Indicated
airspeed must be corrected for position error, instrument error, compress-
ibility, and density altitude to determine true airspeed. The equation lift
shows that the indicated stall airspeed will be about the same for a given
weight and configuration, regardless of the altitude. An increase in alti-
tude produces a higher stall true airspeed but very little, if any, change in
the indicated stall airspeed.

(3) Configuration.

(a) As flaps are lowered, CLmaximum increases. If CLmaximum


increases, stall speed decreases.

(b) To this point, discussion has concerned an aircraft in


equi1ibrium--with no acceleration. With a simple change to the stall-speed
equation (Equation 8.5), it also applies to an accelerating aircraft, as
shown in Equation 8.6.

(4) Acceleration.

(a) Aircraft do not always require lift


equal to weight--
sometimes they require more, sometimes less. An aircraft pulling out of
a

dive generates an additional force that produces acceleration. For example,


an aircraft is just passing through level flight during a pullout of dive, a

as shown in Figure 8-17. In this case, the weight is acting directly oppo-
site the lift
vector. The lift
now required to produce the acceleration must
be more than the weight of the aircraft. In the stall-speed equation, weight
is substituted for lift; lift must be greater than weight if acceleration is

8-19
to be produced. This is accomplished by introducing a load factor (n) to the
stall-speed equation.

LIFT 20,000 LB

.~
DRAG

Figure 8-17.
~
WEIGHT 10,000 LB

Load
THRUST

factor
.

(b) The load factor is the lift the aircraft is required to


develop, divided by the weight of the aircraft, as shown in the following
equation:

n = L/W
(Equation 8.6)
For example, an aircraft pulling out of a dive weighs 10,000
pounds. The
aircraft needs to develop lift of 20,000 pounds to accelerate (pulled Gs).
The load factor would be n = 20~000/10,OOO
2. The aircraft is in a 2-G
=

condition. Although the aircraft actually weighs only 10,000 pounds, its
apparent weight--because of acceleration--is 20,000 pounds. An
aircraft
under a 2-G load condition needs to develop twice the
lift
developed under
straight and level or 1-G condition. The lift must equal weight times the
a

load factor. In the stal17speed equation, weight is substituted


lift is not equal to just weight but to the weight times the lift.
Now the
for
load
factor, as in the following equation:
L =
nW
(Equation 8.7)
With this load factor, the following equation develops:

Vs =
VC:xPS (Equation 8.8)

Because stalling speed is directly proportional to the square root of the


load factor, an increase in the load factor also increases the
stalling speed
of the aircraft. If an aircraft stalls in straight and level flight
at 100
knots, the same aircraft in a 2-G condition will stall at 141 knots.
If more
lift is needed for an aircraft flying at CLmaximum' it can only be developed
by increasing velocity. CL cannot be increased;
it is already at its
8-20
maximum value. Any further increase in the angle of attack would decrease
CL'

(5) Thrust. Another factor that must be considered in


the develop-
ment of the stall-speed equation is the amount of (T)
thrust developed by the
engine. For simplicity>> the thrust vector can be assumed to act along the
chord line of the wing. The angle between the thrust
vector and the relative
wind is equal to the angle of attack, as shown in Figure
8-18. The figure
shows a
vertical
component of thrust, or a thrust component>> that is
parallel
to the lift vector and acting in the same direction. The value of this ver-
tical thrust component can be expressed as T sin 4. Because it is in the
direction of the lift vector>> it is aiding the lift in supporting the weight
of the aircraft. Summing the forces vertically and equating them
to zero
gives the following equation:

L + T a
sin 4 -
nW =

(Equation 8.9)
Solving the above equation for lift yields the following equation:
L =
nW T
-

sin ø
(Equation 8.10)
This expression of L is now substituted into the
original form of the stall-
speed equation and yields the following equation:

V
S
2(nW
=
T

CLmax
V
P
sin 4 )
S
-

(Equation 8.11)

This is the form of the stall-speed equation when the


effect of thrust is
considered. The more thrust being developed, the larger the value of T
and the lower the stalling speed.
sin 4
The larger the value of the vertical com-
ponent of thrust>> the less lift the wing has to develop.
Therefore, the
aircraft can fly more slowly.

THRUST 4,000 LB
t Uff "WO ~

t T sin 4 =1,030 LB

T'"
.~
t .t..
CL

......-------__L_-
FUGHT PATH

J ~G~1~~

Figure 8-18. Effect of thrust

8-21
SECTION IV

MANEUVERING FLIGHT

8-11. CLIMBING PERFORMANCE

Knowledge of climbing performance is essential because climb is encountered


in every flight. The type of climb performance used for a certain mission is
the aviator's decision. How the desired performance is obtained depends on
the aviator's knowledge of the aircraft and its climb performance. An
air-
craft in a climb is increasing potential energy by increasing its altitude
(potential energy equals weight times height). Generally, potential energy
is increased for an aircraft by an expenditure of kinetic energy (airspeed)
or chemical energy (propulsion power).

a. The exchange of kinetic for potential energy is called a "zoom


climb." This is accomplished by flying straight and level to obtain a high
airspeed, then increasing pitch to a climbing attitude. Velocity is dissi-
pated as altitude is gained.

b. The exchange of chemical energy or propulsion power for potential


energy produces a steady-state climb, which can then be sustained. This type
of climb is used most often and is the one discussed in this section.

8-12. CLIMBING FLIGHT

a. Like a car going uphill, an aircraft climbs at the cruise power


setting with a sacrifice of speed. It can also, within certain limits, climb
with added power and no sacrifice in speed. A definite relationship exists
among power, attitude, and airspeed.

b. Forces acting on an
aircraft go through definite changes when the
aircraft is making the
transition from level flight to a climb. The first
change, an increase in lift, occurs when pressure is applied to the elevator
control. This initial change is a result of the increase in the angle of
attack, which occurs when the pitch attitude of the aircraft is raised. This
results in a climbing attitude. When the inclined flight path and the climb
speed are established, the angle of attack and the corresponding
stabilize.
lift again

c. As airspeed decreases to climb speed, air striking the horizontal


stabilizer is reduced. This creates a longitudinally unbalanced condition;
the aircraft tends to nose down. To overcome this tendency and to maintain a
constant climb attitude, additional pressure must be applied to the elevator
control.

The amount of excess power available is the factor that most affects
d.
an aircraft's ability to climb; that is, the power available above that re-
quired for straight and level flight. During the climb,
lift
operates per-
pendicular to the flight path; it is not directly opposing gravity to support
the weight of the aircraft. With the flight path inclined, lift is acting

8-22
partially rearward, increasing induced drag. This adds to the total drag.
Because weight is always acting perpendicular to the
surface of the earth and
drag is acting in a
direction opposite the flight path of the aircraft during
a
climb, thrust must overcome drag and gravity.

8-13. STEADY-STATE CLIMB

During a sustained climb, two climbing performance factors


a. concern
aviators--the angle of climb and the rate of climb.
(1) The angle of climb (y) is the angle between the
and the horizontal plane.
flight path
The maximum, or best, angle of climb may be
quired to clear an obstacle after takeoff. re-

(2) Rate-of-climb performance is the foot-per-minute gain in


altitude (vertical velocity).
(3) The maximum, or best, rate-of-climb speed (Vy)' is flown
at a
lower climb angle and a higher airspeed than the maximum, or best,
climb speed (v
angle-of-
x), Though the aircraft is flown at a lower climb angle,
higher velocity produces a higher rate of climb than could be obtained
during
a maximum
angle of climb, as shown in Figure 8-19.

d,Ø
c\,.''''.
i.O
s'~
~.,)
~
~G~~~' 3oo-fOO
~

",~O~ ~"')
øES"t\1~~,
SPEEO
oo-f.. þ." of,C\.,~Ø
fóa.f ",.."tE:-
... ~~Jt.'I-'
øES"t

Both aircraft 10 seconds after takeoff

Figure 8-19. Climb angle and climb rate

b. As shown in Figure 8-20, a vector diagram is necessary to understand


the action of the four basic forces acting on the
aircraft during a
stabilized, steady-state climb. Certain assumptions are made: the aircraft
is climbing at a constant velocity (constant true airspeed and straight
flight path) and the thrust force is considered to be acting along the flight
path. Using these assumptions, Newton's first law of motion prevails. The
aircraft is in equilibrium; the sum of the forces acting about the center of
gravity of the aircraft equals zero.

8-23
~~ /RW
LIFT
~

r"'- Y

~ WEIGHT~ny
Assume thrust acts along flight path.

Figure 8-20. Force-vector diagram for climbing flight

c. All forces acting on the aircraft are resolved into components


either perpendicular or parallel to the relative wind. In climbing
weight is not perpendicular to the relative wind. flight,
Therefore, weight must be
resolved into its two components, one parallel and the other
perpendicular to
relative wind, as shown in Figure 8-20. The forces can now be added algebra-
ically, as shown in the following equation:
L = W cos T

(Equation 8.12)
T D + W
=
sin T

d. As shown
in the vector diagram in Figure 8-20 and Equation 8.12, the
lift force is less than the weight in a climb. The steeper the climb
angle,
the less lift required to maintain balanced flight; the
thrust force supports
the portion of the weight that is not supported by (W sin
Equation 8.12, if the aircraft could climb straight lift
up ( T
T). Using
90 degrees), =

lift would be zero and the thrust would support the entire weight of the
aircraft and overcome drag.

8-14. ANGLE OF CLIMB

The amount of excess thrust available


determines the angle of climb that can
be maintained. Flight at Vx is usually
just above stall speed or below mini-
mum control speed (Vmc) in multiengine
aircraft. This places the
critical flight speed; any increase in the angle of attack or a aircraft
a at
loss of
power on one engine could result in a stall or loss of
mended
control. The recom-
airspeed for
a
maximum climb angle--such as obstacle-clearance
speed that is listed in some operator's manuals--is not a
air-
true is a Vx but
"safe best angle-of-climb speed. 11
This airspeed is greater than the true
maximum angle-of-climb speed;
it places the aircraft in a safer flight
envelope, while only slightly sacrificing climb performance. During
when obstacle clearance is the primary
takeoff
concern, Vx should be used. The most
altitude is gained for the horizontal distance covered.

8-24
a. Effects Upon Angle of Climb. Altitude, weight, and wind each affect
angle of climb. These effects are discussed below.

(1) Altitude. As an aircraft gains altitude, the thrust developed


by the engine normally decreases. This is true for both turbine and recipro-
cating engines. The angle of climb must also decrease, because a decrease in
the thrust available (TA) causes a decrease in excess thrust. The thrust
required (TR) remains about constant at all altitudes. Because of this, the
aircraft angle of climb decreases to zero degrees when it reaches its abso-
lute ceiling, where TA equals TR'
(2) A
weight increase adversely affects angle-of-climb
Weight.
performance in two ways. An increase in weight increases both weight and the
thrust required. This means there is more weight to be raised with less
excess thrust, resulting in a shallower angle of climb.

(3) Wind. obstacle clearance is of primary concern, the


When
aircraft's best angle-of-climb speed should be used. The best angle of climb
gains the most altitude for the distance covered. Wind must be considered
because it affects the horizontal distance covered to clear an obstacle.
With the aircraft climbing at Vx' as shown in Figure 8-21, the horizontal
distance covered across the ground in a head wind is less than the horizontal
distance covered with no wind or with a tail wind. This affects the angle
that the aircraft climbs over the ground.

....\~o

~~
~~
..~ ~o

GROUND

(NOTE CHANGES IN CLIMB ANGLE.)

Figure 8-21. Wind effect on maximum climb angle

b. Angle of Attack for Best Angle of Climb.

(1) To determine angle-of-climb performance for a propeller


aircraft, a TR and a TA curve must be constructed. The thrust-required curve
is simply the drag curve for the aircraft. As a propeller aircraft increases
velocity, the thrust force coming from the propeller decreases. The angle of
attack for a propeller aircraft when it is climbing at its best angle of
climb is higher than the angle of attack for LiDmaximum'

8-25
(2) This high angle of attack
required of a propeller aircraft is
near its takeoff angle of attack. If a propeller
aircraft must make an
obstacle-clearance takeoff. it continues to climb at an airspeed
close to its
takeoff airspeed (VIor)'

8-15. RATE OF CLIMB

a. Effects Upon Rate of Climb. Altitude and weight affect rate of


climb. These effects are discussed below. Because the
horizontal and verti-
cal velocities are within the air mass (TAS). wind has no
of climb.
effect on the rate

(1)
Altitude. Altitude affects engine performance. As with angle
of climb. an increase in altitude will decrease the
rate of climb. The rate
of climb at the absolute ceiling of an
aircraft is zero. At this altitude.
there is no excess horsepower (HPA equals HPR). At the altitude called the
service ceiling. an aircraft can maintain a 100-foot-per-minute
climb. When operating on a single engine. the
rate of
aircraft can maintain a
50-foot-per-minute rate of climb.

(2)
Weight. As with angle of climb. weight also affects climbing
performance. As weight increases. horsepower required increases.
the decrease in excess horsepower and the increase Therefore.
in weight will decrease
the rate of climb. As an aircraft burns
fuel. its weight decreases. Because
of this weight decrease. more excess horsepower is
available toward the end
of a
flight.
b. Angle of Attack for Best Rate of Climb. The velocity where a
propeller aircraft can obtain its best rate of climb is close to the velocity
for L/Dmaximum' This point is determined from horsepower curves. Measure-
ments are made on those curves; they are not
calculated. Maximum excess
power produces the best rate of climb.

8-16. CLIMBING STALL SPEED

When an aircraft is in climb. it will stall at a lower speed. Stalling


a

speed depends on the amount of


lift
a
wing produces. Reducing the amount of
lift required of the wing also reduces the stalling speed of the
aircraft.
When an aircraft is in climbing
flight. the lift
required of the wing is not
equal to the weight but only to a portion of the weight. This
is due to the
vertical component of thrust. No lift
force is required when an aircraft is
in vertical flight. Therefore. an aircraft cannot aerodynamically
vertical flight. stall in

8-17. PERFORMANCE OF AN AIRCRAFT IN A CLIMB OR DIVE

a. The full-power performance capabilities of an aircraft in a climb


or dive can be visualized with a full-power polar diagram. The diagram
plotted assuming that weight. altitude. and power or thrust is
are held
constant.

8-26
NOTE: If
any of the three factors mentioned above change, then curve and
performance change.

Figure 8-22 shows the typical polar diagram for full-power operation
b.
at 5,000 feet. This curve represents the plot of vertical and horizontal
velocities obtained by the aircraft at full power with different climb and
dive angles.

BEST ANGLE
OF CLIMB MAXIMUM RATE OF CLIMB
3

+VV
X-AXIS

-VV

AIRPLANE AT
CONSTANT WEIGHT
AND ALTITUDE
WITH FULL POWER
~
z

'II.
'0iii INCREASING Ot

~
a:

TERMINAL VELOCITY

Vv = VERTICAL VELOCITY

Figure 8-22. Full-power polar diagram

(1)
Point 1 on the curve in Figure 8-22 represents the maximum
velocity of the aircraft in straight and level flight (at full power).
(2) As the aircraft starts to climb, velocity decreases; the
aircraft gains altitude and has a vertical velocity, as shown at point 2 in
Figure 8-22. The vertical velocity--rate of climb--can be read on the scale
at the left in Figure 8-22. The angle between the flight path and the hori-
zontal velocity line is the angle of climb ( y ).

Point 3 in Figure 8-22 shows the maximum rate of climb (Vy).


(3)
The curve indicates the maximum vertical velocity obtainable by the aircraft.
A line drawn from the origin to the top point of the curve shows the climb
angle and the TAS when the aircraft is climbing at a maximum rate.

8-27
(4) A
line drawn from the origin tangent to the curve indicates the
maximum angle of climb (Vx) for the aircraft at point 4 in Figure 8-22. At
full power the aircTaft performs somewhere on this curve; there cannot be any
steeper climb angle for this aircraft. The TAS and climb angle can be ob-
tained by drawing a line from the origin tangent to the curve.

(5) Any aircraft that


stalls with excess power at its stall speed
will be in climbing flight when
it stalls at full power. Full-power stalling
speed is shown at point 5 in Figure 8-22. The climb angle that the
aircraft
would be able to attain at its stalling speed can be read from the
graph.
(6) Point 6 in Figure 8-22 is the vertical velocity the
aircraft
would attain if it were diving straight down with full power. Many
aircraft
would break up before reaching this velocity because of their
structural
limitations. This point is shown for information only and to complete the
curve.

c. Any change
in altitude, weight, or power setting affects the
performance of the aircraft and produces changes in the full-power polar.
Curves drawn showing these changes are called the family of polar
curves.
Figure 8-23 could be the polar curve of the aircraft at its absolute ceiling,
the partial thrust polar, or even the sea-level polar where the weight of the
aircraft would not allow the aircraft to climb.
(1) If this is the absolute ceiling polar, the
aircraft cannot
climb. Therefore, the full power that the power plant can produce at maxi-
mum altitude is just enough to maintain the aircraft in straight and level
flight.

+VV

VH

CURVE CAN REPRESENT-

. FULL POWER AT ABSOLUTE ALTITUDE


. PARTIAL POWER (MINIMUM POWER TO
MAINTAIN LEVEL FLIGHT)
. EXCESSIVE WEIGHT AT FULL POWER
-VV

Figure 8-23. Polar curve

8-28
(2) The curve shown in Figure 8-23 could
also represent
power. In this case, the aircraft is operating with minimum power partial
to main-
tain level flight. The aircraft encounters this condition when
operating at
maximum endurance.

(3) The third case would not exist; an


much
aircraft loaded with that
weight could not become airborne. However, a similar curve would show
the angle of climb obtainable with full power in an overweight
condition.

SECTION V

GLIDES

8-18. PERFORMANCE

a. Most aircraft depend on mechanical power plants to sustain flight;


complete engine failure is always possible. Although the
aircraft will not
suddenly fall from the sky if the engines were to
fail, it
descend. When this occurs, an aviator must make rapid and
would begin to
correct decisions
based on knowledge of the aircraft's performance in
power-off gliding flight.
b. If complete engine failure occurs, the following items should be
considered:
0
How far can the aircraft glide?
0
What will the rate of sink be?

0
How long will the aircraft remain airborne?
0
Can a
successful power-off landing be accomplished?
Some of these questions that require immediate answers are discussed in this
section. However, other answers must come from the appropriate operator's
manual. To fully understand the data given in the operator's manual, an
aviator must be thoroughly familiar with the aerodynamics of gliding flight.

c. In descending flight, shallow angles of descent are usually referred


to as glides. Steep angles are normally called dives. Glides or dives can
be accomplished with either power on
or power off. A power-on glide (de-
scent), used during every flight, is considered a normal flight condition.
Power-off glides are usually performed only in emergencies.

8-19. POWER-ON GLIDE

a. When power is reduced during straight and level flight, the thrust
needed to balance the aircraft's drag is no longer
adequate. Because of this
unbalanced condition, drag reduces airspeed with a corresponding decrease
in
lift of the wing. The weight of the aircraft now exceeds the force of
lift.
8-29
The resulting flight path is downward as well as forward.
Because the flight
path is inclined downward, the force of gravity is providing
forward thrust.
In effect, the aircraft is actually going downhill.

Forces acting on an aircraft go through definite changes when the


b.
aircraft is making the transition from level cruising flight to a descent.
When pressure is applied to the elevator
control or when the aircraft's pitch
attitude is allowed to lower, the wing's angle of attack is decreased,
is reduced, and the flight path starts downward. This change in flight lift
path
occurs when lift becomes less than the weight of the aircraft as the angle of
attack is reduced. This unbalance of lift and weight causes the
descend in respect to the horizontal path of level
aircraft to
flight. The initial re-
duction of lift that starts the aircraft downward is momentary. When the
flight path stabilizes, the angle of attack again approaches the original
value and lift and weight stabilize.

c. The downward force on the horizontal stabilizer lessens as airspeed


decreases. This produces an unbalanced condition; the aircraft tends to nose
down. Slight pressure on the elevator control is usually needed to maintain
the desired descent attitude.

d. As
descent begins, airspeed may gradually increase because of a
component of weight now acting forward along the flight path. The overall
effect is that of increased thrust, causing the airspeed to increase if the
power were allowed to remain the same as that used for level
cruise flight.
For the aircraft to descend at the same airspeed as flown in
level cruise
flight, power is reduced as the descent begins.
e. As the descent attitude is steepened, the component of weight acting
forward along the flight path increases. Conversely, weight decreases as the
descent attitude is shallowed. Therefore, the amount of power reduction
for
a
descent at cruising speed is determined by the rate of descent
desired.
8-20. POWER-OFF GLIDE

a.In a power-off glide, there is no thrust force. Therefore, to


obtain equilibrium, a component of the weight must replace the thrust force
to balance the drag force. An aircraft can maintain a straight descending
flight path at a constant velocity without the engine. It can also attain
equilibrium so that the sum of the forces about the center of gravity equals
zero. Figure 8-24 shows this condition. The weight vector has been resolved
into two components--?ne parallel and the other perpendicular to the flight
path.

8-30
","11", ~",,>
~!;J þ

HORIZONTAL PLANE
~If
Figure 8-24. Power-off glide

b. The Greek letter y in the last section represented the climb


angle. However, in Figure 8-24 it represents the glide angle. It is still
the angle between the horizontal plane and the flight path.

8-21. MAXIMUM-GLIDE DISTANCE

Achieving maximum-glide ratio is the main concern of the aviator in gliding


a

flight. The ratio of the horizontal distance to the vertical distance is the
glide ratio. At a maximum-glide ratio, the aviator can cover the greatest
horizontal distance with the least loss of altitude. The aviator who is
experiencing engine failure will naturally want a flight path that gives a
maximum-glide ratio. If the aircraft is of the type that can make a success-
ful power-off landing, the aviator must maintain enough altitude for the
approach over a selected landing area. If altitude is not adequate for the
approach, the aviator can do nothing at this point to regain it.
a. To understand how the maximum-glide ratio is found, the two extremes
of descending flight can be considered. If an aircraft loses its engine at
5,000 feet and can only maintain a flight path with a glide angle of 90
degrees (straight down), then the aircraft travels only 5,000 feet. No hori-
zon~al distance is travelled. On the other hand, if an aviator could glide
his aircraft parallel to the horizon (y 00), he could glide an infinite
=

dis~ance; this example, while impossible in reality, demonstrates that the


shallower the glide angle, the greater the horizontal distance covered by
the aircraft.

b. Because the LID ratio determines the distance an aircraft can glide,
weight does not affect this distance. The glide ratio is based only on the
relationship of the aerodynamic forces on the aircraft. Weight affects only
the length of time the aircraft glides.

c. If the aircraft is gliding at its minimum glide angle (best glide


speed) during a
power-off approach and is going to land just short of the

8-31
selected landing point, the aviator tends to increase the
angle of attack to
stretch the glide. This action, which increases the glide
crease the horizontal glide distance. The
angle, will de-
lower than
aircraft cannot glide any shal-
its minimum angle of glide.
8-22. WIND EFFECT ON GLIDES

Atmospheric winds have an important effect on gliding


performance. For
instance, an aircraft gliding at 100 knots into a laO-knot head wind would
not move horizontally. An aircraft would have to increase
above 100 knots to develop some horizontal
its glide airspeed
velocity.

SECTION VI

TURNS

8-23. PERFORMANCE

a. Unlike automobiles or other ground vehicles, an aircraft can rotate


about three axes. Therefore, it has 6 degrees of
motion. The aircraft can
pitch up or down, yaw left or right, and roll left or
freedom of motion, an aircraft can perform many
right. Because of this
maneuvers. All of these
maneuvers use vertical turns, horizontal
turns, or both. This section dis-
cusses vertical and horizontal turns separately, as well as the
posed on these turns.
limits im-
In this discussion, several terms are defined with
certain assumptions being made.

b. When an aircraft. turns, it is not in static equilibrium. Forces


must be unbalanced to produce the acceleration for
turning. When properly
performed, the turn does not produce any sideward force pulling the
inward or outward from the turn. The net resulting aviator
force (lift) acts toward
the center of the turn. This turn is called a coordinated
"level turn" may be confusing because it refers to a turn. The ter.m
constant-altitude turn,
not to a wings-level.turn. Normally, an aircraft is never
turned with the
wings level. Forces acting on the aircraft must be unbalanced for a
turn to
occur; this does not happen if wings are level.
Incorrect banking during
uncoordinated turns causes slips or skids. This is uncomfortable
crew and passengers.
for the
During uncoordinated turns at slow airspeeds, control
can inadvertently be lost.
c. The force that actually turns the aircraft is the lift force. The
horizontal component of lift is the force that accelerates the
ward the center of the aircraft to-
turn. The rudder counteracts adverse aileron effect
(yaw). The elevator increases the angle of attack to produce the added
required because of the loss of the vertical lift
lift component and an apparent
increase in weight, which is produced by centrifugal
force.

8-32
8-24. TURNING FLIGHT

a. An aircraft, like any moving object, requires


sideward force to
a

make turn. In
a
normal turn, this force is supplied by banking the air-
a

craft so that lift is exerted inward as well as upward. The force of lift
is then separated into two components that are at right angles to each other.
The lift acting upward and opposing weight is called the
vertical-lift com-
ponent. As shown in Figure 8-25, the horizontal-lift component (centripetal
force) is the lift acting horizontally and opposing inertia or centrifugal
force. Therefore, the horizontal-lift component is the sideward force that
forces the aircraft from straight flight, causing it to turn. If an aircraft
is not banked, no force is present to make the turn unless rudder application
causes the aircraft to skid in the turn. Likewise, if an aircraft is banked,
it turns unless it is held on a constant heading with the opposite rudder.
Proper control technique assumes that an aircraft is turned by banking and
that in a banking attitude it should be turning.

b. Banking an aircraft in a level turn does not by itself change the


amount of lift.
However, the division of lift
into horizontal and vertical
components reduces the amount of lift
supporting the weight of the aircraft.
Thus, the reduced vertical component causes a loss of altitude unless the
total lift
is increased by increasing either the angle of attack of the wing
or the airspeed or both. If a level turn with no change in thrust is as-
sumed, the angle of attack is increased by raising the nose of the
aircraft
until the vertical component of lift
equals the weight. When the angle of
bank is greater, the vertical
lift
component is weaker. The angle of attack
is also greater for the lift-weight balance necessary to maintain a level
turn.

8-33
LIFT LVERTICAL
I

-~---

LHORIZONTAL I I (CENTRIFUGAL FORCE)


(CENTRIPETAL FORCE)
I
I
I
I
I
I
I

-1-
---
-

WEIGHT I

Figure 8-25. Effect of turning flight

8-25. RADIUS OF TURN

The radius of turn of an aircraft varies directly with the square of its
velocity (true airspeed) and inversely with the angle of bank. Therefore,
any two aircraft that can fly at the same velocity and angle of bank can fly
in formation, regardless of their weights. Only two variables directly af-
fect the radius of turn. However, certain aerodynamic considerations affect
the velocity; they also affect the radius of turn indirectly. These
considerations--weight, altitude, load factor, angle of attack, and wing
area--affect the velocity and, thereby, affect the turning radius. All of
these aerodynamic considerations playa part in the lift force that must be
produced. To turn an aircraft in the smallest possible radius, an aviator

8-34
flies at the slowest possible speed and the highest possible angle of bank.
The limits on radius-of-turn performance are the aerodynamic, structural, and
power limits. The aviator must be constantly aware of these limits while
maneuvering the aircraft at or near its design limits.

a. Aerodynamic Limit of Performance. Because the horizontal component


of lift is the force that turns the aircraft, a fixed-wing aircraft reaches
its aerodynamic radius-of-turn limit when the aircraft turns at its stall
velocity. The airspeed at which the minimum radius of turn occurs is the
stalling speed where maximum Gs can be pulled without exceeding the design
load-limit factor. This is the maneuvering airspeed. An increase in weight
or altitude or a decrease in CLmaximum requires an increase in velocity,
which increases the radius of turn.

b. Structural Limit of Performance.


(1) The load factor is purely
function of the angle of bank; the
a

weight of the aircraft itself does not affect the G-load imposed on the air-
craft. Both the T-42 and the OV-1 accelerate 2 Gs in a 60-degree bank. The
table in part A of Figure 8-26 shows the load factor at various angles of
bank. The graph in part B of Figure 8-26 shows how the increasing load fac-
tor affects the stalling speed of an aircraft. In the first 60 degrees of
bank, the load factor increases by only one. However, in the next 10 de-
grees, the load factor increases almost one. At higher angles of bank, the
load factor and stalling speed increase rapidly. A steep turn immediately
pfter takeoff is extremely dangerous because of the load factor imposed and
the low velocity of the aircraft.

(2) The stall speed increases as the bank angle increases.


Therefore, acompromise between the bank angle and stalling speed must be
mad~ to obtain the minimum radius of turn of the aircraft. Minimum radius of
turn is found by considering the design strength of the aircraft. An air-
3 Gs develops a
craft that pulls lift force three times its weight. Aircraft
are designed to take certain loads. If these load limits are exceeded, the
aircraft becomes overstressed. Load limits are published in the appropriate
operator's manual. If an aircraft has a load limit of 3 Gs, the minimum
radius of turn would be at its 3-G stalling speed. This occurs at an angle
of bank of about 73 degrees, as shown in part A of Figure 8-26. The maneu-
vering speed of the aircraft is the velocity at which the minimum radius of
turning can be performed at a given altitude without exceeding the load
limit.

8-35
ANGLE LOAD % INCREASE
A
OF BANK FACTOR OF INDUCED DRAG
$ n

O" 1.000 0%

100 1.015 3%

30" 1.154 33%

450 1.414 100%

600 2.000 300%

70" 2.923 761%

80" 5.747 3200%

850 11.473 13163%

90"
0<) 0<).
8
B

5
-

.5.
a:
0
I-
0 4
c(
II.
C
c(
0
...I 3

0 250
50 100 150 200

STALLING SPEED (Vs)

Figure 8-26 .
Effect of load factor on stalling speed

8-36
(3)
Paragraph (2) added a structural limit to turning performance.
The aircraft mentioned above cannot make a coordinated level turn at a bank
angle greater than 73 degrees because of the structural limits of the air-
frame. A 90-degree bank has an infinite load limit, but the structural limi-
tations of the aircraft make a 90-degree banked level turn impossible. At a
90-degree bank, no force develops to oppose weight (Figure 8-27). In this
case, the lift force acts 90 degrees to the weight force. Therefore, the
lift could not support the weight, and the aircraft would not maintain
altitude.
NO
THRUST
t-
o
I
I
I

/'
I

LIFT ...
.. .. ... .. ...... CENTRIFUGAL FORCE
. .

WEIGHT

Figure 8-27. Aircraft unable to maintain altitude in a

90-degree banked turn

c. Power Limit of Performance. The third limit imposed on the turning


performance is the power, or thrust, limit. The amount of induced drag de-
veloped at high-load factors can become quite large. Induced drag is di-
rectly proportional to lift squared. In a 73-degree bank, three times more
lift is produced than in level flight; therefore, induced drag is nine times
that of a level flight at the same velocity. This is a tremendous amount of
drag to overcome. The power available from the power plant is the limiting
factor.
8-26. VERTICAL TURN

a. aircraft can also perform a vertical turn. The aviator uses a


An
vertical turn to pullout from a dive or to conduct vertical maneuvers such
as loops, split-Ss, and Immelmanns. Again, the aviator is concerned with the
radius of turn of the aircraft and whether enough altitude is available to
perform the desired maneuver.

b. Figure 8-28 shows an aircraft just passing level flight when lift is
opposite weight. Because the aircraft is turning about some point in space,
there is also centrifugal force. The radius of turn is greatest at the bot-
tom of the pullout, where forces are directly opposite the lift force.

8-37
LIFT

THRUST 11

,>"'f. i .
.~~)ìi[1tfJrd!f.j~:::::"
.
WEIGHT
.
.
.
.
.

"
CENTRIFUGAL FORCE

Figure 8-28. Vertical balance


8-27. RATE OF TURN

The rate of turn is used primarily during instrument flight and is simply the
rate of change in the heading of the aircraft per unit of time (second or
minute). The angle of bank and velocity are important because they affect
the rate of turn as well as the radius of turn. Higher velocities and shal-
lower bank angles result in slower changes in aircraft headings. The rate of
turn is at its maximum at the maneuvering speed and angle of bank for the
G-limits. This is consistent with the discussion of the radius of turn in
paragraph 8-25.

SECTION VII
TAKEOFF AND LANDING PERFORMANCE

8-28. PROCEDURES AND TECHNIQUES

a. For every successful flight, there must be at least one successful


takeoff and one successful landing. Techniques used in older aircraft during
these phases of flight were essentially the same for each type of aircraft.
Because of design differences in today's aircraft, however, procedures and
techniques differ from aircraft to aircraft. Some aircraft rotate to their
takeoff attitude early in the takeoff roll. Some aircraft make a constant
angle of attack approach and hold that angle of attack until landing; others
flare just before touchdown, which decreases the rate of sink. Aerodynamic
braking is effective in some aircraft during a landing roll. In other air-
craft, aerodynamic braking is dangerous.
b. The design characteristics of an aircraft determine its takeoff and
landing performance. However, a de~ailed discussion of the various aircraft
is beyond the scope of this section. This section is concerned primarily

8-38
with the aerodynamic and physical considerations that determine the length of
runway needed for a successful takeoff or landing.

8-29. TAKEOFF

The takeoff isan acceleration and transition maneuver. During a takeoff


run, the aircraft makes the transition from a
ground-supported to an
air-supported vehicle. At the start of the takeoff roll, the entire weight
of the aircraft is supported by the wheels. As the aircraft gains velocity,
the wing begins to support more and more of the weight. By the time the
aircraft reaches its takeoff velocity, the wing supports the entire aircraft.
The aircraft has then completed the transition from ground to air support.

a. Acceleration Forces.

(1) According to Newton's second law of motion. a body accelerates


only if there is an unbalanced force acting on that body. The acceleration
takes place in the direction of the unbalanced force. For the aircraft to
accelerate during the takeoff run, the sum of the horizontal forces acting on
the aircraft must yield an unbalanced force in the direction of the thrust.
Figure 8-29 shows the horizontal forces that determine the net accelerating
force on the aircraft during its takeoff roll. For a given altitude and RPM,
the thrust from a propeller-driven aircraft decays as velocity increases
during the takeoff roll.

DRAG

Figure 8-29. Net accelerating force

(2) The wing of the aircraft is close to the ground during the
takeoff run. This reduces or cancels out the downwash and wing-tip vortexes
behind the wing. Reduction of downwash also reduces induced drag. This
phenomenon is known as ground effect. It normally reduces induced drag about
1.4 percent at one wingspan, 23.5 percent at one-fourth wingspan. and 47.6
percent at one-tenth wingspan. Therefore, the drag force, as shown in
Figure 8-26. is predominantly parasite. Because parasite drag is directly
proportional to the square of the velocity, this drag force increases as
aircraft velocity increases.
(3) The friction force (Ff).
shown in Figure 8-29, is called
rolling friction. It
results from the rolling action of the tires against
the runway. Like any friction force, this rolling friction is equal to the

8-39
product of the coefficient of rolling friction and
normal (perpendicular)
a

force. The coefficient of rolling friction varies


from 0.02 to 0.3, depend-
ing on the runway surface and the type of
tire. In this case, the normal
force is the aircraft weight not supported by the wing. As mentioned before,
as the aircraft gains speed during the takeoff
and more of the weight.
roll, the wing supports more
This weight reduction on the wheels reduces the
rolling friction force as the aircraft accelerates. When the aircraft
reaches takeoff velocity, the friction force is zero because the normal force
is zero.

b. Takeoff Distance. Takeoff distance is directly proportional to


takeoff velocity squared. Because velocity is squared, its effect on takeoff
distance is significant. The takeoff velocity is a function of the stalling
speed of the aircraft and is usually 1.1 to 1.25 times the power-off
stalling
speed. Some aircraft use high-lift devices to lower the stalling speed,
which decreases takeoff velocity and, thus, takeoff distance. These high-
lift devices are used only to increase CL"

(1) Altitude.
An aircraft taking off from a field at a S,OOD-foot
eleva~ion requires longer takeoff run than the same aircraft taking off at
a

sea level. As altitude increases, air density decreases. This requires an


increase in true airspeed to develop the required amount of
lift. This in-
crease in airspeed increases the takeoff distance. Engine performance is
also affected as elevation is increased. Increases in altitude decrease the
thrust output. Therefore, the net accelerating force decreases at a higher
elevation, and takeoff distance increases because of this thrust loss.
(2)Weight. Weight changes, as well as air density changes, have a
compounding effect on the aircraft's takeoff distance. If an aircraft could
double its takeoff weight, it might be assumed that the takeoff distance
would double. If weight had no effect on the takeoff velocity, this
assump-
tion would be correct. However, doubling the weight doubles the value of the
square of the takeoff velocity. This value, when combined with the increased
value of the weight, yields a takeoff that is four times the original dis-
tance. Another factor is also affected by the weight increase--the rolling
friction force. The increased weight increases the normal force and, there-
fore, the rolling friction force. This results in a decrease in the net
accelerating force, which adds additional distance to the takeoff run.

(3) Wind.

(a) Wind component. Until now, the discussion of takeoff


distance has assumed a no-wind condition. However, wind can be used to
decrease the takeoff distance. For example, an aircraft is assumed to be
taking off into a head wind equal to its takeoff velocity. While the air-
craft is sitting at the end of the runway, the wind velocity over the wing is
enough to support the aircraft. The aviator does not need to accelerate be-
cause the aircraft does not require a takeoff run or takeoff distance. The
ground speed of the aircraft is zero, even though the true airspeed is equal
to the takeoff velocity.

8-40
(b) Runwaydirection. Most runways are built in the direction
of the local prevailing winds. If a runway must be used that has a tail-wind
component, the value of the tail wind must be added to the takeoff velocity;
then the sum of the two must be squared. This means that a large increase in
the takeoff distance is necessary.

c. Takeoff Performance Summary. Various factors influence takeoff


performance and takeoff distance. The interplay between some of these fac-
tors, such as weight and altitude~ makes accurate takeoff distance require-
ments difficult to predict. Additionally, the assumption was made that
acceleration was constant during the takeoff roll, but this is not neces-
sarily true. The aviator can determine the actual takeoff distance required
from the aircraft operator's manual~ which has charts that include the ef-
fects of temperature, pressure altitude, aircraft weight, winds, and runway
slope. These charts, taken from flight tests, show the expected performance
of each aircraft.

8-30. LANDING

In a landing roll, the aircraft must decelerate, not accelerate. As velocity


decreases and the lift force decays, weight shifts from the wings of the
aircraft to the wheels. This is the reverse of the transition and accelera-
tion mentioned in paragraph 8-29.

Distance. During a landing, the aviator is primarily concerned with


a.
dissipating the kinetic energy of the aircraft. Any factor affecting the
mass or velocity of the aircraft must be considered in computing the landing
distance of the aircraft. Because velocity is a squared term in the kinetic
energy equation, the final approach is always flown at the lowest velocity
possible. This condition requires careful planning and execution by the
aviator.
(1) Net decelerating force. Compared to the takeoff, the forces
that compose the net accelerating force of the landing are reversed. The
acceleration force is now in the direction of the drag and friction forces.
Therefore, the aircraft slows down. For this net decelerating force to be
developed, drag and friction must be greater than the thrust force. However,
thrust force must still be considered, even though the engine is usually at
idle. There is a residual thrust force that must be overcome by drag and
friction if the aircraft is to decrease its velocity. Some aircraft are
equipped with propellers that can reverse their pitch. Therefore, the pro-
peller can also develop a thrust force in the direction of the retarding
forces. This thrust increases the retarding force and decreases the required
landing distance.

(a) The net decelerating force can be increased on some air-


craft by aerodynamic braking, which is nothing more than increasing the drag
on the aircraft during the landing roll. The drag force is proportional to
the square of the velocity. Therefore, aerodynamic braking is effective only
during the initial landing roll, when the aircraft is at higher velocities.

8-41
(b) If
short landing roll is required,
a

increase the friction force far above that of any it is possible to


could be applied to the aerodynamic force that
aircraft. This increase in the friction force is
done with the wheel brakes. Brakes must not be applied too
the initial landing roll, most of the early. During
aircraft weight is supported by the
wings. Using brakes at this time is not effective because
the normal force
on the wheels is low and the
resulting friction force developed is small.
Also, the wheels may lock and the tires may blowout
if the brakes are ap-
plied too hard at this time. Velocity should be decreased so that enough
weight is transferred to the wheels for the brakes
to be effective. Some
aircraft have wing spoilers that are used to destroy the lift on the wings.
Weight is then transferred to the
tires, allowing the brakes to be applied
earlier in the landing run. Retracting the flaps on some aircraft also
duces the lift developed by the wings. re-

(c) The condition of the runway surface also


affects the
required stopping distance of an aircraft. On wet or icy
runways, the
ficient of friction is small, resulting ina small decelerating force. coef-
Therefore, a longer stopping distance is required. A runway condition
ing is determined by using a decelerometer. The RCR is given read-
in the remarks
section of weather sequence reports, which are supplied by the
forecaster service and by the air traffic control facility. Thepilot-to-
RCR should
always be considered during a landing on a slick
runway. The landing dis-
tance required may exceed the available runway length.

(2)Hydroplaning. Wet or slippery runways can lead to hydro-


planing. The hydroplaning aircraft rides on a film of water; the
tires have
little or no contact with the runway surface. The aircraft is supported by a
hydrodynamic lift force much like a water skier is supported by
skis. The
aviator must be aware of the conditions that cause hydroplaning and know how
to avoid them.

(a) For a
simplified explanation of hydroplaning, an aircraft
can again be compared to a
water skier. To support the skier, a hydrodynamic
lift force develops that depends on speed. Below speeds where aerodynamic
forces dominate, the faster the speed, the easier
the same way, there is a minimum speed at which it is to hydroplane. In
hydroplaning occurs. Below a
certain speed, however, the drag is so great and the hydrodynamic
so small that the skier sinks. lift force
Likewise, below the speed where hydroplaning
occurs, the tires will directly contact the runway.

(b) The velocity at which


total hydroplaning can occur depends
on the square root of the tire inflation
pressure. Partial hydroplaning may
occur at slower speeds. Under inflated tires can hydroplane
at even slower
speeds.

(c) An assumption might be


that heavier aircraft have to move
at faster speeds than lighter aircraft before hydroplaning
occurs. However,
experiments and classical hydrodynamic theory show that the speed
hydroplaning occurs is independent of weight. Weight only
at which
determines the
size of the "footprint" that the tire makes. The heavier the
bigger the footprint. However, the ratio of weight aircraft, the
per square inch of

8-42
footprint area is the same. Weight has an indirect effect because a
heavier
aircraft must fly at a
faster approach and touchdown speed. Thus the possi-
bility of hydroplaning is greater in heavier aircraft.

Other factors that affect hydroplaning cannot be quantita-


(d)
tively described in a formula. For example, how deep the water must be on
a

runway before hydroplaning develops is not well defined. Tire tread depth
A
and pattern, as well as the runway surface itself, are also factors.
smooth tire can hydroplane in as as .15 inch of water on the runway.
little
A
tire with deep tread has channels for the water to escape, while part of
the tire contacts the runway. This tire may need as much as 2 inches of
water before hydroplaning occurs. Puddles on the runway can also cause in-
termittent hydroplaning. A smooth runway surface, in contrast to coarse
a

surface, may result in earlier hydroplaning. Experiments have been conducted


with a slotted crosswise surface to allow water to escape as the tire rolls
over the runway.

Hydroplaning can occur with any aircraft. Tires should be


(e)
checked during preflight for proper inflation and tread condition. The avia-
tor should avoid crosswind landings when possible andA should fly at minimum
airspeeds when landing on wet or slippery runways. rule of thumb for
determining hydroplaning speeds is to multiply the square root of the tire
pressure by nine.

(3) Deceleration speed alonR landinR roll. Figure 8-30 shows the
speed at various distances from touchdown to a full stop. Again, this as-
sumes a constant deceleration, which is not necessarily true but is easy to
visualize. For example, if the total landing distance requires 4,500 feet
and the touchdown speed is 130 knots, the speed is higher than half the
touchdown speed at half the landing distance. With this in mind, the aviator
can avoid overbraking.

150

125

100

éñ
t-
0
z 75
!!
>
t-
Õ
0 50
...
w
>

25

0
0 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0

LANDING DISTANCE (FEET x 1,000)

Figure 8-30. Landing-roll velocity


8-43
(4) Landing velocity. The final approach airspeed of
an aircraft
is about 1.3 times its stalling speed. Therefore, any high-lift device that
decreases the stalling speed also decreases the landing speed. A decreased
landing velocity decreases both the kinetic energy and the required landing
distance. Longer landing distances are necessary for landings made at high
altitudes because these higher altitudes result in faster true airspeeds for
final approaches.

(5) Wind. A head wind results in a lower final approach and slower
touchdown velocities (ground speed). Reduced kinetic energy, with respect to
the ground, decreases the landing distance.

(6) Weight. The weight of the aircraft affects landing distance


the same as it does takeoff distance. Decreased weight requires less runway,
whether an aircraft is landing or taking off.

b.Performance Summary. As with takeoff performance, landing perfor-


mance is difficult to predict. Each aircraft operator's manual has landing
distance charts, which show the results of flight tests conducted under dif-
fering conditions. These charts consider the factors affecting landing
distance.

SECTION
VIII
FLIGHT CONTROL SYSTEMS

8-31. DEVELOPMENT

After early aircraft designers built surfaces that would yield enough lift to
support an aircraft, the greatest problem was to gain adequate and positive
control of the airborne aircraft. Early gliders flown at the end of the last
century were controlled by shifting the location of the center of gravity in
relation to the aerodynamic center. To do this, aviators shifted their body
weight. This method of control not only proved inadequate but was often
disastrous. The greatest contribution to flying made by the Wright brothers
was the development of an adequate control system. They developed a method
of warping the wings for lateral control. They also added a rudder for use
with each wing. Others had already developed the elevator for gliders. This
section discusses the theory of control surface operation and control re-
quirements. Also covered are the types of control systems in use.

8-32. CONTROL SURFACE OPERATION THEORY

a. The Wright brothers' method of warping the wings changed the value
of the coefficient of lift on each wing. In effect, they changed the camber
of the airfoils. Today, flaps are used to vary the camber of the airfoils,
which varies the coefficient of
lift. If a flap is deflected downward, as
shown in Figure 8-31, the camber of the airfoil is increased. This results
in a higher coefficient of lift.

8-44
Figure 8-31. Using flaps to increase camber

b. Ailerons on conventional aircraft operate in opposite directions.


a

To bank an aircraft to the right, as shown in Figure 8-32, the left aileron
is lowered, while the right aileron is raised. This increases the camber of
the left wing and decreases the camber of the right wing. With increased CL
as compared to the right wing, the left wing has a greater lift force; this
is assuming both wings are at equal or nearly equal velocity. This unbal-
anced lift force between the two wings results in a rolling moment about the
longitudinal axis. Therefore, the aircraft rolls to the right.

~ . SPOILER EFFECT OF RIGHT (UP) AILERON DESTROYS

~ iì)Q)d :;'FT
AND THUS lOWERS THE WING.

.
CONTROL MOVEMENT HAS BEEN
-AW-
===- ~ ----
ï
GROSSLY EXAGGERATED FOR THE
PURPOSE OF DEMONSTRATION.

\ LEFT (DOWN) AILERON INCREASES


............'8 CAMBER AND ANGLE OF ATTACK, THUS
'.. INCREASING LIFT AND RAISING THE

~ '- WING.

RIGHT AILERON UP, -RW


RIGHT WING DOWN
LEFT AILERON DOWN,
LEFT WING UP. RESULT: ROLL TO THE RIGHT

Figure 8-32. Operation of aileron in a


turn

As shown in parts A and B


of Figure 8-33, the elevators attached to
c.
the horizontal stabilizer and the rudder attached to the vertical stabilizer
work in the same manner to develop pitching moments and yawing moments. The
stabilizers are normally symmetrical airfoils. Deflecting the control
8-45
surface changes the airfoil to either a positive- or negative-cambered
foil, depending on the direction of the surface movement. air-

~Q
A RESULTANT ....~
RW
þ
ELEVATOR
(lllffH-HJ<<:'\ I

.+
UP

ELEVATOR ~v]
DOWN

RW
þ
~1
(lllf)Hm~
RESULTANT ~
~()
ELEVATOR UP, TAIL DOWN,
ELEVATOR REACTION
NOSE UP. RESULT: PITCH UPWARD
TO ACTION OF
RELATIVE WIND

RUDDER LEFT, TAIL RIGHT, NOSE LEFT.


RESULT: YAW TO THE LEFT

Figure 8-33. Effect of elevator and rudder on moments

d. Control effectiveness is the term used when discussing


control systems. It does not refer to the ability of the controlaircraft
surface to
maneuver the aircraft but to the amount of change in the
coefficient of lift
per degree of control-surface deflection. Control effectiveness
the change in coefficient of lift and not to the refers to
amount of lift change
produced by deflecting the control surface. Because
the lift formula applies

8-46
to control surfaces, the amount of lift depends not only on the value of CL
but also on the velocity, surface area, and air density.

e. control surface that has 'been deflected 10 degrees produces a


A

greater change in the lift force at 200 knots than at 100 knots. At both
airspeeds, the CL change is the same. Therefore, control effectiveness does
not change. However, the greater lift change at the higher airspeed repre-
sents better control response. This should indicate that if a definite mo-
ment is desired, the control surface deflection must be increased as velocity
is decreased.

8-33. LONGITUDINAL CONTROL

a. Maneuvering Control Requirements.

(1) All Army aircraft can attain CLmaximum' They can fly at any
value of the coefficient of lift that is designed into the aircraft. Army
aircraft can also obtain the maximum lift the airfoil can produce for a
given
airspeed and altitude.

(2) Figure 8-34 shows the effect of center of gravity location upon
the longitudinal maneuvering capability of an aircraft. The slanted lines
represent different locations of the CG of the aircraft. The 30 percent mean
aerodynamic chord denotes that the CG is located 30 percent of the MAC length
back from the leading edge of the airfoil. The mean aerodynamic chord is the
chord of a rectangular wing that has the same pitching moments as the wing
under consideration. The lower the percentage, the farther forward the cen-
ter of gravity is located. More elevator deflection is required to obtain a
certain value of CL as the CG is moved forward. This increases longitudinal
stability and, therefore, would be unsuitable for most military requirements.
To correct this situation, the CG must be moved aft or the elevator must be
designed so that it can produce a larger moment.

z
0
~
0
w
~
~
w UP
C
~
0
~
c
>
w
~ DOWN

Effect of CG longitudinal control


Figure 8-34. location on

8-47
b. Takeoff Requirement.

(1) An aircraft should be able to


reaching takeoff velocity. This rotate to takeoff attitude before
wheels. This is an essential and
rotation occurs about the main landing gear
demanding requirement.
attain takeoff attitude at about 0.9 Normally, it will
Vs, Figure 8-35 represents an
rolling down a runway and shows the forces aircraft
These adverse moments
that produce adverse moments.
be overcome by the
try to pitch the nose of the aircraft down. They
must
elevator control surface as it causes the
aircraft to pitch up. nose of the

(a) First,rolling friction force is generated at the


a

wheels. This force is located below the CG and


In an aircraft with a tricycle landing produces CG negative pitching
a
moment.
ahead of the main wheels. gear, the must be
located
main wheels as
Therefore, if
the aircraft is to rotate about the
it
achieves takeoff attitude, this CG
location produces an
adverse pitching moment that must be
is ahead of the wheels, the overcome. the If lift
force of the wing
the nose of the lift force will assist the elevator in pitching
aircraft up.
(b) Another condition that
ground effect, or the decrease in downwash
is not evident in Figure 8-35 is
the runway. When downwash is because the aircraft is close to
tive and
decreased,
usually more beneficial. In this
lifting surfaces become more effec-
case, however, the elevator
force of the horizontal stabilizer. The tries
to increase the negative
lift
effect makes the horizontal stabilizer less ground
effective in producing negative
lift. Therefore, greater elevator deflection is necessary when the
is being operated in ground effect aircraft
than when the aircraft is being operated
out of ground effect.
Aircraft that use flaps during the takeoff
produce a downwash that might be of run can
benefit to the horizontal stabilizer and
decrease the elevator control
deflection. However, trailing-edge flaps also
generate a negative pitching moment.
The net result might be a
elevator-deflection greater
requirement.

WING LIFT

Figure 8-35. Adverse moments during takeoff

8-48
The elevator must be designed to produce
(2) a
positive pitching
moment--below the aircraft flight speed--that can overcome all adverse pitch-
ing moments created by some or all of the above-mentioned conditions.

Landing-Control Requirement. The landing-control requirement is


c.
essentially a ground-effect problem. The aircraft should have enough eleva-
The
tor control to maintain a landing attitude to the touchdown point.
the runway
elevator must overcome ground effect when the aircraft approaches
surface. If an aircraft is able to takeoff satisfactorily, it usually has
enough elevator control to land.

8-34. DIRECTIONAL CONTROL

Adverse Yaw. Adverse aileron yaw was described in the last section
a. The rudder must
as yaw developed when an aircraft is rolled using ailerons.
yaw created by an aileron
develop enough yawing moments to overcome adverse
wind the nose of the
roll. The rudder is also used to keep the relative on
aircraft so a coordinated turn can be performed.
b. Spin Recovery. An aircraft has
a
large sideslip angle as it is
spinning. The rudder must be used on all Army fixed-wing aircrafta to
decrease the sideslip angle before the aircraft can recover from spin.
This requirement can be critical in some aircraft.

c. Slipstream Rotation and P-Factor.

(1) Most aircraft engines rotate the propeller clockwise, as viewed


from the cockpit. This induces a clockwise airflow about the fuselage that
strikes the left side of the vertical stabilizer. Therefore, the vertical
stabilizer is subjected to a negative sideslip angle, while the rest of the a
aircraft is not. As shown in Figure 8-36, this negative sideslip produces
'

negative yawing moment and tends to move the nose of the aircraft to the
moment when the air-
left. The propeller can also develop negative yawing
a

craft is at a high angle of attack. In Figure 8-37, the propeller disk has
been inclined from the normal to the flight path. Because the propeller disk
is inclined to the flight path, the downgoing blade (right side) has a
greater angle of attack and velocity than the upgoing blade (left side).
As
a
result, the downgoing blade produces
a
larger thrust than the upgoing
added to the
blade, which produces a negative yawing moment that must be
slipstream yawing moment. The propeller disk asymmetric loading is called
the P-factor.

(2) The directional requirement resulting from slipstream rotation


and P-factor is important to propeller-driven aircraft performance. The
both the
rudder must be able to overcome the negative yawing moments of
P-factor and the slipstream rotation. It must also be able to maintain
directional control. The adverse moments created by the slipstream rotation
and the P-factor at high angles of attack are increased as the aircraft slows
down. However, the rudder moment used to counteract the adverse yawing
moments decreases as the velocity decreases. Therefore, the rudder must
be deflected even more. This can also be a
critical control requirement.

8-49
VERTIC~
STABILZER
LIFT COMPONENT

Figure 8-36. Effect of negative sideslip angle on vertical stability

~~
~--= AIRCRAFT RELATIVE WIND

Figure 8-37. Effect of negative yawing moment on nose

d. Crosswind Takeoff and Landin~. An


a
crosswind must have a track over the groundaircraft taking off or landing in
that is parallel to the runway
heading. If the aircraft is going to make the desired track over the ground,
it must sideslip through the air mass that is moving across the runway. The
rudder produces the required sideslip
so the landing or takeoff can be made
in the direction of the runway heading.
The aircraft is at low airspeeds
during these phases of the
lem is again amplified by
flight. Therefore, the directional control prob-
the lack of high dynamic
pressure.
e. Asymmetrical Thrust. A
multiengine aircraft has an additional
directional-control requirement that is not required on a single-engine

8-50
aircraft.
I
This is the yawing moment caused by asymmetrical thrust, which
results from a difference in power on each wing. The difference in the
thrust force developed on each wing produces a yawing moment about the air-
craftls center of gravity away from the thrust. This yawing moment must be
counteracted by an opposite moment from the rudder. While an engine failure
on the critical engine would develop the greatest amount of asymmetrical
thr.ust, it is present any time one engine produces more power than the other.

8-35. LATERAL CONTROL

a. Roll Rate. Each aircraft is designed to perform specific types of


A
missions. These designs require different roll rates. fighter, which is
more maneuverable than a
transport, must have a
fairly high roll rate to
perform its mission. To initiate roll,
a the aviator deflects the ailerons.
One aileron moves up, decreasing the CL and lift on that wing. The other
aileron moves down, increasing the CL and lift on that wing. This difference
in lift between the two wings causes the aircraft to roll toward the wing
that produces less lift. As the aircraft begins to roll, the angle of attack
on the downgoing wing increases while the angle of attack on the upgoing wing
decreases. This is caused by the component of airflow moving opposite the
direction of wing movement. Within usable values of CL (below stall) and as
the wing moves down, CL increases; CL on the upgoing wing decreases. This
difference in the angle of attack between the two wings produces
a
rolling
moment that is counter to the rolling moment developed by the ailerons. As
the roll rate increases, the counterrolling moment increases. When the
counterrolling moment produced by the roll equals the rolling moment produced
by aileron deflection, a steady-state roll is reached. The roll rate is at
maximum for that given aileron deflection. This steady-state roll rate must
be rapid enough to be compatible with the mission requirements of the
aircraft.
b. Crosswind Takeoff and Landing. As mentioned during the directional-
control requirements discussion, an aircraft must sideslip through the air
mass during a crosswind landing or takeoff. Because of the dihedral effect,
a sideslip angle produces a
roll that is away from the sideslip. The
stronger the dihedral effect of the aircraft (positive lateral stability),
the greater the lateral control required during this condition of flight.

8-36. CONTROL FORCES

Control forces refer to the forces that the aviator exerts on the control
column to control the aircraft. These forces must be logical and manageable.
A logical force means that the force must increase as the control surface is
deflected or as the speed of the aircraft is increased. Manageable means the
magnitude of the control force must be within the comfortable physical capa-
bilities of the aviator.
a. Required Control.

Overcoming hinge moments. When a control surface is deflected,


(1)
the pressure differential developed across the control surface creates a
force that tends to streamline the control surface, as shown in Figure 8-38.

8-51
This force creates a moment about the
hinge of the control surface, which is
referred to as the hinge moment. This moment must be overcome by the force
applied by the aviator to the control column. The hinge
moment is directly
proportional to V2 and to the area. of the control surface.
is doubled, the hinge moment for a given deflection If the airspeed
increases four times.
Therefore, the aviator must exert four times the control force on the
column to overcome that hinge moment. control

+
HINGE
MOMENT

Figure 8-38. Hinge moment

(2) MaintaininR
aerodynamic balance. As aircraft size increases,
larger control surfaces are required. Therefore, as aircraft speed and air-
craft size increase, control forces become unmanageable. The control system
contains many levers and bell cranks that give the aviator a
mechanical ad-
vantage over the hinge moments developed by the large
control areas and dy-
namic pressures. However, the magnitude of the
control force is still too
large to overcome. This led to the development of aerodynamic balancing
the control surfaces. Aerodynamic balance does not of
actually balance any-
thing. To achieve this balance, aerodynamic forces aid in deflecting the
control surface. Aerodynamic balancing uses the dynamic pressure of the
airstream to reduce the hinge moments of a control surface. One or more
the devices covered below are used in aerodynamic of
balancing.
b. Types of Devices.

(1) Horn. The horn is probably the


first type of aerodynamic
balancing developed. It is an area located ahead of the hinge
area creates a moment that opposes the hinge moment developed byline.
This
the area of
the control surface behind the hinge. The moment
developed by the horn is in
the same direction as the moment developed by the
aviator and aids in dis-
placing the control surface. It decreases the hinge moment
that the aviator
must overcome. The horn can be either exposed (unshielded)
or hidden
(shielded), depending on the design requirements of the
shown in parts A and B of Figure 8-39.
control surface, as
An unshielded horn is normally more
effective than the shielded horn; however, it is seldom used in modern high-
performance aircraft because of drag considerations.

8-52
~
FIXED SURFACE
~- 0(

HINGE LINE
FIXED SURFACE
~

LI CONTROL SURFACE CONTROL SURFACE

A UNSHIELDED HORN B SHIELDED HORN

Figure 8-39. Aerodynamic balancing using horns

(2) Balance board. Another type of aerodynamic balance is internal


balance. With internal balance, a balance board is contained in the fixed
portion of the airfoil ahead of the control surface, as shown in Figure 8-40.
A flexible
seal separates the balance panel from the wall of the plenum cham-
ber where the balance board is located. Thus the plenum chamber is divided
into two compartments. A slot is located between the control surface and the
fixed portion of the airfoil. When the control surface is deflected as
shown, the increased velocity caused by the increased camber over the top
slot develops a lower static pressure in the compartment at the top of the
plenum chamber rather than in the lower compartment. A pressure differential
develops across the balance panel that moves the panel upward. This assists
in deflecting the control surface downward. The decreased pressure over the
slot on the top of the airfoil decreases static pressure, which then de-
creases the hinge moments of the control surface. Therefore, this becomes a
form of aerodynamic balancing.

FLEXIBLE
SEAL

FIXED
SURFACE

Figure 8-40. Aerodynamic balancing using a


balance board

(3)
Servo tabs. Servo tabs can be considered flaps on flaps, as
shown in Figure 8-41. When it is deflected, the servo tab produces a small
aerodynamic force behind the hinge of the control surface. This small force

8-53
deflects the control surface, which then moves the aircraft. The servo tab
itself does not move the aircraft; it only moves the control
control force required to move a surface. The
small tab is much less than
the force re-
quired to move the entire control surface. Servo
tabs enable the aviator to
use the aerodynamic qualities of the
airstream to reduce hinge moments. Jet
airliners have servo tabs on many control surfaces.

CONTROLSUFACEFORCE
CHANGE MOVES AIRCRAFT

~/~
~(j;~l~
---';:::~~.iERVO TAB

SERVO TAB FORCE CHANGE


MOVES CONTROL SURFACE

Figure 8-41. Aerodynamic balancing using a


servo tab
(4)Balance tabs. Another device used to assist with
forces is the balance tab. This small flap is similar
control
to the servo tab shown
in Figure 8-41. The balance tab moves in the opposite
direction of the
trol flap. If the elevator moves up, the balance tab automatically movescon-
down because of the linkage mechanism
controlling its movement. This down-
ward movement of the balance tab produces a moment
that assists the movement
of the control surface. Sometimes these systems
incorporate springs in the
linkage; they are referred to as spring tabs.

(5) Trim tabs. Trim tabs are identical


in appearance and operation
to servo tabs. Trim tabs are connected to trim
controls in the cockpit and
are used to trim the aircraft for different weights,
airspeeds, and power
conditions. For example, an aircraft has 200 pounds more
than in the other. For the aircraft to fuel in one wing
remain wings-level, the lift force on
the heavier wing must be increased to hold up the added
weight. This is
accomplished by deflecting the ailerons, which
requires the aviator to hold
the ailerons in that position. To decrease the
tabs can be deflected to hold the
aviator's work load. trim
ailerons.
c. Mass BalancinK. Mass balancing of the control surface is
in making the aircraft dynamically stable. important
If the center of gravity of the
elevator is behind the hinge line. a sudden gust of wind that moves the
of the aircraft upward causes the elevator to be tail
deflected downward. This
downward elevator movement increases
movement of the tail.
tail lift, thereby increasing the upward
If, however. the CG of the elevator is ahead of the
8-54
hinge line, the pitchup caused by the air gust deflects the elevator upward,
which dampens the upward motion by developing a negative tail lift.

8-37. CONTROL SYSTEMS

Only few of the various control systems and components will be discussed in
a

this paragraph. Many more exist. However, all of them fall under the cate-
gories of conventional, power boosted, stability augmenter, full power, flap,
or landing gear.

a. Conventional. The control system that has been considered in this


section has been the conventional, mechanical, and reversible control system.
Conventional refers to the type of control system that employs a rudder, an
aileron, and an elevator. Mechanical refers to the method used in the con-
trol system to deflect the control surfaces such as cables, pushrods, and
bell cranks. Reversible means the control system has feedback. When the
aviator moves the stick/wheel, the surface moves; when the surface moves, the
stick/wheel moves. The aviator must feel air loads on the control surfaces.
That will eliminate the aviator's tendency to overcontrol the aircraft, which
can produce overstress. Conventional control is used in all Army aircraft
except the C-12.

b. Power Boosted. As aircraft continue to be designed to go faster,


air loads or hinge moments created on the control surfaces become so large
that aerodynamic balancing is not effective without creating large increases
in drag. Therefore, a new form of control system has been developed called
the power-boosted conventional, reversible control system. This system is
similar to power steering. The aviator supplies part of the control force
through a mechanical linkage with the control surface. However, a power
system in parallel also supplies part of the force to overcome hinge moments.
Usually, a power-boosted system is about a 20:1 or 30:1 ratio. If the avia-
tor puts 1 pound of force in the control system, the power system supplies 20
or 30 pounds of force. Normally, the power system is hydraulic, but pneu-
matic and electrical devices have been used. This type of system is revers-
ible, and the aviator can still feel air loads through the mechanical system.
If the power system fails, the aviator controls the aircraft through the
mechanical system. However, the needed control force is greatly increased.
As mentioned previously, this is similar to power steering and is used in the
C-12 rudder system.

c. Stability Augmenter. When disturbed from equilibrium, an aircraft


exhibiting positive static stability naturally oscillates because of the
moment of inertia caused by the disturbance. A damping force causes the

oscillation to be convergent to equilibrium. However, in some cases, the


aerodynamic damping of the aircraft is insufficient to reestablish equilib-
rium in the desired time. If so, an artificial means of damping must be
used. The stability augmenter system is an auxiliary system designed expli-
citly for this purpose. Pitch and yaw augmenters are common on high-
performance aircraft. Damping forces are overridden when controls for
maneuvering are used, and roll augmenters are normally unnecessary.

8-55
d. Full Power. A
full-power control system is used on supersonic
aircraft when the requirementis to deflect large surfaces against extremely
high dynamic pressures. This type of system is not
necessarily conventional;
it can use spoilers rather than ailerons and slab tails instead of elevators
or rudders. As shown in Figure 8-42, a control surface of this type is much
more effective at supersonic velocities than a flap
control surface. How-
ever, this system is not employed on Army aircraft.

Figure 8-42. Spoiler used as control surface

e. Flap. The wing flap is a movable panel on the inboard trailing edge
of the wing. The flap is hinged so it can be extended downward into the flow
of air beneath the wing to increase lift and drag. The wing flap permits a
slower airspeed and a steeper angle of descent during a landing approach. In
some cases, wing flaps are also used to shorten takeoff
distance.
(1) flap operating control may bean electrical or hydraulic
The
control on the instrument panel, or it may also be a lever located on the
floor or pedestal of the aircraft. As shown in Figure 8-43, the control can
be placed in the following positions: UP, which
raises the flaps if they are
in an extended position; NEUTRAL, which allows the flaps to remain in
an
intermediate position; and DOWN, which lowers the flaps if they are in the
retracted or intermediate position. In addition to the flap operating con-
trol, usually an indicatQr shows the actual position of the flaps. On most
Army aircraft, the maximum extent of flap travel is about 43 to 45
degrees.

FLAP OPERATING CONTROL

~
~ FLAPS RETRACTED-
SHALLOWER APPROACH ~ uP

DfNH
~ FLAPS EXTENDED-

STEEPER APPROACH

Figure 8-43. Wing flap control


8-56
(2) The performance of an aircraft is noticeably affected by
extending and retracting the flaps. For an aircraft with a constant power
setting in level flight. airspeed will be lower with flaps extended because
of the drag created. If power is adjusted to maintain a constant airspeed in
level flight, the aircraft's pitch attitude will usually be lower with flaps
extended.

(3)
When the flaps are extended, airspeed should be at or below the
maximum flap-extended speed (Vfe) of the aircraft. If the flaps are extended
above this airspeed. the force exerted by the airflow can damage the flaps.
If airspeed limitations are exceeded unintentionally with the flaps extended,
they should be retracted immediately, regardless of airspeed.

(4) On the instrument panel, the flap control often has the shape
of an airfoil. It must be correctly identified before the flaps are raised
or lowered. This will prevent inadvertently operating the landing-gear con-
trol and retracting the gear instead of the flaps. particularly when the
aircraft is on or near the ground.
Landing Gear. Some Army aircraft have a retractable landing-gear
f.
control. Because the only purpose of the landing gear is to support the
aircraft on the ground. it becomes excess weight and drag during flight.
Although the weight of the gear cannot be reduced during flight, the landing
gear can be retracted into the aircraft structure and out of the airflow.
This eliminates unnecessary drag.

(1) The control for operating the landing gear is a switch or


lever. This control is often in the shape of a wheel, which differentiates
it from the flap control on the instrument panel. When this control is moved
to the DOWN position, the gear extends. When the control is moved to the UP
position, the gear retracts. In addition to the operating control, an in-
dicator or warning light on the instrument panel shows the present position
of the landing gear.

(2) The landing gear should be operated only when the airspeed is
at or below the maximum landing-gear operating speed (VIa) of the aircraft.
Operation at a higher airspeed can damage the operating mechanism. When the
gear is down and locked, the aircraft should not be operated above the air-
craft's maximum landing-gear extended speed (VIe).
(3)
The landing gear control must be correctly identified before it
is raised or lowered. This will prevent inadvertently operating the flap
control and retracting the flaps instead of the landing gear.

8-38. PROPELLERS

a. Operation.

(1) An aircraft propeller converts the power plant shaft horsepower


into propulsive horsepower. It provides thrust to propel the aircraft
through the air. The propeller consists of two or more blades and a central

8-57
hub to which the blades
are attached. Each propeller blade is an airfoil;
therefore, the propeller is a
rotating wing.
(2)
The engine furnishes the power to rotate propeller blades. The
propeller is mounted on a shaft that is normally an extension of the crank-
shaft on low-horsepower engines. On high-horsepower engines, the propeller
shaft is usually geared to the engine crankshaft. In either case, the engine
rotates the airfoils of the blades through the air at a relatively high ve-
locity. The propeller then transforms the rotary power of the engine into
thrust.
(3) Many different factors govern the efficiency of a
propeller.
Generally, a large-diameter propeller favors a high-propeller efficiency from
the standpoint of a large mass flow. However, compressibility effects and
high-tip speeds adversely affect propeller efficiency. Small diameter pro-
pellers favor low-tip speeds. The propeller and power plant must be matched
for compatibility of both output and operating efficiency.

b. Types of Propellers.

(1) Fixed pitch.

(a) A
fixed-pitch propeller has blade pitch (blade angle)
built into the propeller. For that reason, the pitch angle cannot be changed
by the aviator as it can be on controllable-pitch
propellers. Generally, the
fixed-pitch propeller is constructed of aluminum alloy.
(b)
Fixed-pitch propellers are designed for best efficiency at
one rotational They fit a specific set of conditions
and forward speed.
involving both the engine rotational speed and the forward speed of the air-
craft. Any change reduces the efficiency of the propeller and the engine.
(2) Constant speed.

(a) In automatically controllable-pitch propeller systems, a


control device adjusts the blade angle to maintain a specific preset engine
RPM without the
aviator's constant attention. If engine RPM increases as
the result of a decreased load on the engine, the system automatically in-
creases blade angle of the propeller. This increases the air load until the
RPM
returns to a preset speed. An automatic control system responds to such
small variations in RPM that, for all practical purposes, a constant RPM is
usually maintained. These automatic propellers are termed "constant-speed"
propellers.
(b) An automatic system has a governor unit that controls the
blade pitch angle so that engine speed remains constant. With cockpit con-
trols, the aviator regulates the propeller governor. Thus the desired blade
angle setting (within its limits) and engine operating RPM can be obtained.
This increases the operational efficiency of an aircraft in various flight
conditions. A low-pitch, high-RPM setting obtains maximum power for takeoff.
After the aircraft becomes airborne, a higher pitch and a lower RPM setting
provide adequate thrust for maintaining proper airspeed, as shown in

8-58
Figure 8-44. This can be compared to using low gear in a vehicle to
accelerate until high speed is attained and then shifting into high gear for
cruising speed.

BLADE ANGLE BLADE ANGLE

~ ~
LOW PITCH HIGH PITCH

Blade angle affected by RPM


Figure 8-44.

c. Propeller Feathering.

(1) Ifpower plant malfunctions or fails on aircraft with two or


a

more engines, propeller blades must be streamlined to reduce drag. Flight


can then be continued on the remaining operating engines. This is accom-
plished by "feathering" propeller blades, which stops rotation and reduces
drag on the inoperative engine. When the propeller blade angle is in a
feathered position, parasite drag is at a minimum. On most multiengine air-
craft, the added parasite drag from a single feathered propeller adds rela-
tively little to total drag.
(2) At smaller blade angles near the flat-pitch position, the
propeller, which is windmilling at high RPM, adds a large amount of drag to
the aircraft. This drag may cause the aircraft to become uncontrollable.
The propeller that is windmilling at high speed in the low range of blade
angles can produce an increase in parasite drag that may be as great as the
parasite drag on the rest of the aircraft. An indication of this drag is
best shown by an autorotating helicopter. A windmilling rotor can produce
autorotation rates of descent that approach that of a parachute canopy with
identical disk-area loading. The windmilling of the propeller at a high
speed and a small blade angle produces an effective drag coefficient of the
disk area that compares to that of the parachute canopy of the same size.
The drag and yawing moments caused by loss of power at high airspeeds are
considerable, and the transient yawing displacement of the aircraft may
produce critical loads on the vertical tail. For this reason, automatic
feathering may be a necessity rather than a luxury.

d. Reverse Thrust. The large amount of drag produced by a


rotating
propeller can improve the stopping performance of the aircraft. Rotation of

8-59
the propeller blade to small positive or even negative angle of attack
values with applied power produces a large amount of drag or reverse thrust.
Because of the high-thrust capability of the propeller at low speeds, reverse
thrust alone produces high deceleration.

e. Limitations. Operating limitations of the propeller are closely


associated with those of the power plant. Because of the large centrifugal
loads and blade twisting moments produced by an excessive rotative speed,
overspeed conditions are critical. In addition. propeller blades have vari-
ous vibratory modes. Certain operating limitations may be necessary to pre-
vent resonance.

8-60
CHAPTER 9

FIXED-WING FLIGHT TECHNIQUES

This chapter implements STANAG 3554 (Edition Two).

While this manual does not deal with specific Army aircraft,
an overview of the flight maneuvers and techniques that
it does provide
aviators will use to
accomplish their mission. Aviators are measured by their ability to assess
the variables associated with flying and to act on an acceptable solution to
any given situation. They can do this only as they understand their own
capabilities and limitations and those of the aircraft. By not fully under-
standing the aircraft limitations, aviators may set such a high safety margin
that both aircraft performance and the mission suffer. Conversely, they may
set a low safety margin, which can result in damage to the aircraft, cause
injury to the aircrew, and curtail the mission. To do what is right in a
particular situation, aviators must understand aircraft performance. This
chapter covers the actual mechanics of fixed-wing flight maneuvers and
procedures.

9-1. TAXIING

Taxiing is the controlled movement of an aircraft across the ground under its
own power, except for the takeoff and landing
roll. The taxiing speed is
normally that of a brisk walk, but the taxiing area and other conditions may
cause that speed to be altered.

a. Use of Controls in Taxiing.


(1) To attain and maintain the desired taxiing speed,
aviators use
the throttle and propeller pitch angles and sparingly apply the brakes when
necessary. For directional control, they use the steerable tailwheel or
nosewheel, rudder, and brakes. Effective braking is essential, particularly
with the greater weight of multiengine aircraft. As the
aircraft taxies
forward, aviators should test the brakes immediately by depressing each brake
pedal. If the brakes are weak, taxiing should be discontinued and engines
shut down.

(2) Wind has definite effects on an aircraft moving on the


ground.
It tends to retard or increase speed and turn the aircraft. If an aviator
uses the aileron and elevator controls improperly, a light aircraft
taxiing
in a strong wind can be damaged. Figure 9-1 shows the control positions that
are used to counteract the effects of wind.

9-1

.. ...
USE UP AILERON .. USE UP AILERON
ON LH WING
AND'
NEUTRAL ELEVATOR.
"
ON RH WING AND
NEUTRAL ELEVATOR.

.
.

.
.
"
,
USE DOWN AILERON USE DOWN AILERON
ON LH WING AND , ON RH WING AND
DOWN ELEVATOR. , DOWN ELEVATOR.
.. ...
CODE

WIND DIRECTION.

Figure 9-1. Aileron and elevator control positions

(a)
In a head wind. When taxiing a conventional gear aircraft
into a
direct
head wind, aviators hold back the elevator control to
raise the
elevators and to exert a downward force on the tail and hold on the it
ground. In a tricycle gear aircraft, they hold the elevator and aileron
controls in a neutral position to give a neutral elevator and aileron condi-
tion. If
the wind is quartering off the nose, they position the elevator and
aileron controls to help keep the aircraft level. In a conventional gear
aircraft, they position the controls so that the elevator and the aileron on
the upwind wing are defle~ted up. In this position, the aileron on the
up-
wind wing spoils lift
while the aileron on the downwind wing lowers and
simultaneously creates more on that wing.lift
In a tricycle gear aircraft,
they position the elevator control to give a neutral elevator setting. The
aileron control is held in the direction of the upwind wing. This causes the
upwind aileron to move up, reducing lift; the aileron on the downwind wing
lowers, creating more lift.
The downward deflected aileron also creates more
drag, which helps counteract the tendency of the aircraft to turn into the
wind.

(b)
In a tail wind. When taxiing a conventional gear aircraft
in a
tail wind, aviators hold the ailerons in a neutral position and move the
elevator control forward to lower the elevators, allowing the wind to exert a
downward force on the tail. If the wind is quartering off the
hold the elevator control forward and the aileron control towardtail,
they
the downwind
wing. This deflects the aileron on the downwind wing up and the aileron on
the upwind wing down. The downward deflected aileron helps prevent the

9-2
aircraft from turning into the wind.
In a tricycle gear aircraft, aviators
hold the elevator control in neutral position or slightly aft to raise the
a

elevators and hold the aileron control in a neutral position. If the wind is
quartering off the tail of the tricycle gear aircraft, they move the elevator
control forward and hold the aileron control toward the downwind wing to
prevent the aircraft from turning into the wind and to hold the wings lev~l.
Aviators should never hold the elevator control with force. The propeller
slipstream or taxiing speed may be strong enough to offset the tail wind.
When this condition exists, a slight downward or upward position of the ele-
vator normally allows for wind gusts. Less throttle is required to taxi in a
tail wind, because the wind helps move the aircraft. Aviators use the brakes
as necessary to prevent excessive speed, but they should avoid sudden braking
and power bursts.

b. Precautions in Taxiing. Multiengine aircraft are usually somewhat


heavier, larger, and more powerful than single-engine aircraft, and they
often require more time and distance to accelerate or stop. Aviators are
given a different perspective when they look outside the cockpit while taxi-
ing a multiengine aircraft. Although it usually is not necessary to make S-
turns to observe the taxiing path, aviators must be especially vigilant to
avoid obstacles, other aircraft, or personnel. When making sharp turns, the
aviator should never pivot the aircraft on one wheel; this is dangerous. By
doing so, they can damage the landing gear, tires, and airport pavement. All
turns should be made with the inside wheel rolling. To avoid accidents while
taxiing, aviators should observe the following rules:
0
Keep the aircraft under constant control.
0
When possible, use S-turns (conventional/gear) to afford maximum
visibility of the taxiing area.
0
Match the speed to the taxiing area and the aviator's proficiency.
0
Do not attempt sharp turns at excessive speeds.
0
In confined areas, use wingmen or ground guides or move the aircraft
by hand.

0
Take extra precautions at night.
0
When in doubt about the area, make a thorough ground reconnaissance
before taxiing the aircraft.
0
Position the controls according to the direction and speed of the
wind.
0
Use wing walkers when wind speed is excessive and when directional
control is likely to be too difficult or dangerous.
0
Apply brake pressure smoothly and continuously to stop forward
movement.

9-3
9-"}.. . TRIM TAB USE

a. Trim tabs serve the same purpose in


a
multiengine aircraft as in a
single-engine aircraft. However, their function in a multiengine aircraft is
usually more important for safe and efficient flight because of the greater
control forces, weight, and power. They are also more important because of
the asymmetrical thrust with one engine inoperative, range of operating
speeds, and range of center-of-gravity location. In most multiengine air-
craft, the aviator's strength is taxed in overpowering an improperly set
elevator trim tab on takeoff or go-around. Many fatal accidents have oc-
curred when aviators have taken off or attempted a go-around with the air-
craft trimmed "full nose up" for the landing configuration. Therefore,
prompt retrimming of the elevator trim tab is essential to the success of
a

flight in an emergency go-around from landing approach.


a

All aircraft should be retrimmed in flight for each change in


b.
attitude, airspeed, power setting, and loading. Without such changes, firm
forces must constantly be applied on the flight controls to maintain any
desired flight attitude.

9-3. TAKEOFFS

Actual flight begins with the takeoff. The takeoff roll begins as throttles
are advanced to start the roll down the active runway. It continues until
the aircraft becomes positively airborne.

Normal Takeoff. As shown in Figure 9-2,


a.
a
normal takeoff is made as
directly into the wind as the runway will allow. In this manner, aircraft
become airborne with a shorter ground roll.

~
~
l'II/i!;IIIII~/I1iJ1I IIIII I III II I II III II
II
Figure 9-2. Normal takeoff

(1)
After completing the before-takeoff check and upon receiving
appropriate clearance, the aviator taxies the aircraft onto the runway. He
aligns the aircraft with the runway and selects runway marks or other ground
reference points to use as aids for directional control during the takeoff.
The aviator holds the brakes and performs a lineup check. The brakes are
released, and the throttles are advanced to takeoff power. Directional

9-4
control is maintained by smooth and positive rudder application. Because of
torque, a sudden application of power can cause the aircraft to yaw sharply
to the left. Engine instruments are checked early in the takeoff run to
confirm that the engines are developing takeoff power and that no malfunction
exists. Fuel pressure is adjusted, if applicable. At the start of the take-
off run, sufficient aileron is established to keep the wings level and the
elevator control is held slightly aft of neutral. Positioning the elevator
control as described will minimize a "wheelbarrow effect" by keeping the
weight of the aircraft off the nose gear. Directional control with rudders
is maintained, as necessary, to keep the aircraft going straight down the
runway. Use of the brakes for directional control is avoided; braking
lengthens the distance required for takeoff and can result in a blown tire as
aircraft speed increases. As the aircraft approaches rotation speed (Vr)'
aft pressure is increased on the elevator to establish the attitude at which
the aircraft will fly off the ground at lift-off speed (Vlof)' As necessary,
the ailerons are used during the takeoff roll to keep the wings level. When
flight is assured, the landing gear is retracted and the proper initial climb
attitude (best- rate-of-climb speed [VyD is established. The flaps are re-
tracted, if applicable, at the best single-engine rate-of-climb speed (Vyse)
or at minimum single-engine control speed (Vme)' At a single-engine maneu-
vering altitude that is not less than 500 feet above ground level, the pitch
attitude is adjusted to obtain cruise climb airspeed. As cruise climb air-
speed is attained, the power for the climb is adjusted using maximum continu-
ous power or as specified in the operator's manual and the mixture control is
set if applicable. The after-takeoff check is then completed.
NOTE: aircraft that is equipped with a manual fuel enrichment system
When an
is taking off at an airport where the elevation is fairly high above sea
level (Denver, Colorado), the aviator may need to set takeoff power before
starting the takeoff roll and then lean the engine to achieve maximum engine
performance.
(2) A multiengine aircraft differs little from
normal takeoff in a

one in single-engine aircraft. The controls of each class of aircraft are


a

operated the same. The throttles of the multiengine aircraft are normally
treated as one compact power control. They can be operated simultaneously
with one hand.
NOTE: For the demonstration and practice of normal takeoffs, refer to the
appropriate aircrew training manual.

b.Crosswind Takeoff. While it is usually preferable to take off


directly into the wind, circumstances or wisdom may often indicate otherwise.
Consequently, an aviator must be familiar with the techniques for crosswind
takeoffs as well as those for normal takeoffs. A crosswind will affect the
aircraft during takeoff much as it does in taxiing. In taxiing and during
takeoffs, the technique for crosswind correction is quite similar. The tech-
nique in taxiing is explained in paragraph 9-1a(2).

(1)
The technique used during an initial takeoff roll in a crosswind
is generally the same as that used during a normal takeoff. One difference
is that the aileron control must be held into the crosswind. This raises the

9-5
aileron on the upwind wing to impose a downward force on the wing. The down-
ward force counteracts the lifting force of the crosswind and prevents that
wing from rising.
(2) During the crosswind takeoff roll, sufficient aileron must be
held into the wind to keep the upwind wing from rising and to hold that wing
down so that immediately after lift-off the aircraft will slip into the wind
enough to counteract drift. In effect, the aircraft is slipped into the wind
only enough to offset wind drift. This results in a straight ground track
parallel to the longitudinal axis and prevents a side stress on the gear if
the aircraft returns to the runway. After the aircraft becomes positively
airborne, the slip is removed and a crab is established to correct for drift
and allow a faster rate of climb.

(3) Crosswind takeoffs are performed in multiengine aircraft in


basically the same manner as those in single-engine aircraft. Less power
.
may
be used on the downwind engine to overcome the tendency of the aircraft to
turn into the wind at the beginning of the takeoff. Then full allowable
power is applied to both engines as the aircraft accelerates to
a speed where
better rudder control is attained.
NOTE: For the demonstration and practice of crosswind takeoffs, refer to the
appropriate ATM.

c. Takeoff With Obstacle Clearance Climb. If an obstacle must be


cleared after takeoff, aviators perform climbout at the best-angle-of-climb
speed (Vx) until the obstacle has been cleared. After reaching this point,
they may accelerate to the best-rate-of-climb speed (Vy)' retract the flaps
and gear, and then reduce the power to the normal climb setting.

NOTE: For the demonstration and practice of takeoffs with obstacle clearance
climb, refer to the appropriate ATM.

d. Minimum-Run Takeoff. Performance data might not be available for


a

minimum-run takeoff. Since the takeoff is performed below Vme and power-off
stalling speed (Vs) control can be lost if an engine fails at or immediately
,

following lift-off before the best single-engine angle-of-climb speed (Vxse)


can be attained or before power on the operating engine is reduced. This
procedure may be used when it becomes necessary to take off from short
a

field or a soft field covered with water, mud, sand, or snow. Itmay also be
used to take off from a
hard, rough surface that could cause damage to the
landing gear.
.

(1) Without obstacle. The before-takeoff check is completed,


including the use of flaps as prescribed in the operator's manual. The air-
craft is aligned with the desired takeoff track using all of the available
area. If the surface is adequate for the use of brakes, the brakes are
held
firmly and the throttles are advanced to the takeoff setting. Then, the
brakes are released. (If the brakes will not hold, they are released and
power is adjusted during the roll.) The engine instruments are kept within
limits, and the fuel pressure is adjusted if applicable. Directional control
is maintained with nosewheel steering and rudder. Wings are kept level with
9-6
..

the ailerons. The takeoff roll is started with the elevator control full
aft. As the elevator becomes effective, the nosewheel is raised clear of the
surface. As the aircraft accelerates, elevator pressure .is adjusted as nec-
essary to maintain a nose-high attitude until the aircraft becomes airborne.
When lift-off occurs, the angle of attack should be reduced gradually to
maintain the wheels just clear of the surface until Vlof is achieved. After
is retracted and the
Vlof is attained and flight is assured. the landing gear
aircraft is accelerated to Vy. Flaps are retracted, if applicable, at Vyse
or Vmc. Vy speed is maintained until single-engine maneuvering altitude is
reached, and then the aircraft is accelerated to cruise climb speed.

With obstacle. This is a military contingency maneuver for


(2)
which performance data might not be available. Since the takeoff is per-
formed below Vs and Vmc' control of the aircraft can be lost if an engine
fails after takeoff before the aircraft attains safe speed. If the engine
a

be
fails while the airspeed is below Vs and Vmc' aircraft control can only
maintained by reducing power on the operating engine. If the airspeed is
just above Vs' the only way to maintain control may be to establish
a g~ide.
there is insufficient alti-
If an engine fails above Vmca but below Vxse and would
tude to descend and attain safe airspeed that result in a single-
engine positive rate of climb, a controlled descent to a landing area should
be maintained.

9-4. AIR WORK

The four fundamental flight maneuvers are straight and level flight, climbs,a
glides and descents, and turns. Mastery of these fundamental maneuvers is
prime requisite for aviator proficiency, because any other flight maneuver is
a combination of two or more of these.

a.StraiRht and Level FliRht. Perhaps the most demanding maneuver is


straight and level flight, because the aviator has to apply continuous cor-
rections to maintain a constant heading and altitude. Many factors affect
the flight attitude of an aircraft, and continual small corrections are
required.
(1) In flight, the aircraft must be kept steady in relation to its
lateral, longitudinal, and vertical axes and the horizon. These axes are CG
three imaginary lines perpendicular to each other and passing through the
of the aircraft. The aircraft rolls about the longitudinal (nose-to-tail)
about the
axis, yaws about the vertical (up-and-down) axis, and pitches
lateral (wing tip to wing tip) axis. The aviator uses reference points or
instruments in the aircraft to detect movement about these axes.

The longitudinal (or pitch) reference point for level flight is


(2)
usually some portion of the nose cowling or a bolt on the windscreen. In
level flight, this reference point may not rest directly in line with the
horizon, but it will serve that purpose if the variation from the horizon is
A relative increase of distance
clearly established in the aviator's mind.
between the eye and the reference point permits the aviator to easily and
quickly notice the changed relationship between the point and the horizon.
When this reference point falls below the desired relationship, back pressure

9-7
on the elevator control may be required to return to straight and level
flight; or, if this reference point rises above the established relationship,
forward elevator control pressure is required. With variations of attitude
caused by changes in atmosphere, power, and CG, the aviator has to change the
reference point or the relationship to fit the new attitude for level flight.
(3) Lateral level flight is established by equalizing the distance
between each wing tip and the horizon and is controlled by the ailerons.
Checking the lateral axis for straight and level flight also serves to verify
the clearance of the area. This is an important safety habit for the avia-
tor. Furthermore, the head motion eases aviator tension and lessens fixation
of attention to a
limited field of vision.

(4) Straight flight is best achieved by establishing a flight


relationship between the nose (longitudinal axis) of the aircraft and a
ground reference point such as a section line, road, or fence line. To main-
tain straight flight, aviators use ailerons and the rudder. As training be-
comes more advanced, they can determine direction by referring to the
magnetic compass or the directional indicator.

(5) Cross-checks with the turn-and-slip indicator should be made to


ensure that the aircraft is in coordinated flight.

(6) Trim tabs should be adjusted for all conditions of flight to


relieve aviator control pressures.
NOTE: A affects the other two. Therefore, additional
change to one axis
control pressures are required for the other axes. In attempting to maintain
straight and level flight, however, the aviator must not constantly alter
flight controls. An aircraft will roll and pitch with air turbulence; if
practicable, the aircraft should be allowed to level itself by inherent
design stability. Aviator fatigue can result from constant control
corrections.

b. Climbs. Climbs are used for ascent to a higher altitude. The type
of climb that an Army ~viator will most often use is cruise climb; other
types of climbs will be limited to short duration.

(1) Cruise climb. The cruise climb is made at the angle, airspeed,
and power that will give a desirable lift/drag ratio. For aircraft with
RPM should be increased to the climb setting
variable-pitch propellers,
before the manifold pressure or torque is increased. The recommended climb
airspeed and power setting are given in the operator's manual.

(a) Cruise climb attitude is entered from straight and level


flight by applying back pressure to the elevator control. As the aircraft's
nose starts to rise, airspeed begins to dissipate and power should be slowly
and smoothly added to obtain climbing power and speed simultaneously. Con-
tinuous movement of the throttles is required to maintain a constant power
setting' as the aircraft climbs to higher altitudes. The aviator enters the
climb pitch attitude by relating reference points on the aircraft to the
horizon.

9-8
NOTE: Flight controls feel quite different and react more slowly to aviator
pressures at slower climbing airspeeds.

(b) During a climb, increased power and decreased speed result


in a
left turn or yaw. To compensate for this, aviators must apply pressure
on the right rudder pedal throughout the climb. If rudder trim is available
by cockpit control, it should be used to relieve the pedal pressure. If
available, elevator trim should be used to relieve elevator-control pressure
during the climb.

(c) When the desired altitude is reached, the power will be


adjusted to maintain the desired airspeed in level flight. This is best
accomplished by lowering the nose of the aircraft as the desired altitude is
approached and by maintaining climb power. When altitude and cruising air-
speed are reached, level-off is completed by reducing power to normal cruise.
When power is being reduced, manifold pressure or torque is reduced
first,
then RPM. The aircraft should then be retrimmed for straight and level
flight at normal cruise.
(2) Best angle of climb. The best angle of climb yields the
greatest amount of altitude for a given distance across the ground; it is
used to perform an obstacle-clearance climb. The best angle of climb is
performed the same as the cruise climb, except the pitch attitude is adjusted
to attain and maintain Vx as shown in the operator's manual. Those airspeeds
have been selected to give adequate margins above the flaps-up stall speed
and Vmc. If climbing is necessary to clear an obstacle when only one engine
is operating, the best angle of climb becomes Vxse.
(3) Best rate of climb. The best rate of climb yields the greatest
amount of altitude for a given time. It is used for the initial climb seg-
ment after takeoff and, if desired, to expedite the climb to altitude. The
best rate of climb is performed the same as the cruise climb, except the
pitch attitude is adjusted to attain and maintain Vy as shown in the opera-
tor's manual. If climbing is necessary for other than clearing an obstacle
when only one engine is operating, maximum Vyse is used.

c. Glides and Descents. A glide is used to descend to a lower altitude.


A
normal glide is at an angle and airspeed permitting the greatest forward
distance with a given loss of altitude. Refer to the operator's manual to
determine the best glide airspeed.

(1) To enter a glide from straight and level flight, the aviator
should reduce power smoothly to either the idle or required setting. Alti-
tude is maintained until the airspeed dissipates to normal glide airspeed for
the particular aircraft. Then the nose is lowered to maintain that airspeed.
With reduced power and airspeed, control pressures on both the elevator and
the rudder should be coordinated to maintain glide attitude and direction.
Without power, the aircraft tends to turn to the right and the left rudder
pedal has to be used to counteract this tendency. As airspeed decreases,
back pressure should be applied to the elevator control to keep the nose in
level attitude. When gliding airspeed is reached, back pressure is reduced

9-9
to lower the nose of the aircraft and to maintain the gliding attitude. If
available, trim is used to relieve aviator-control pressures.
(2)As in the climb, initial glide training requires reference
points to maintain the correct glide attitude until the ability to sense and
feel the proper speed and attitude is developed. The level-off from a normal
glide should begin about 50 feet before the desired altitude is reached. At
this time, the aviator should increase power slowly and smoothly to that
desired. While straight and level flight is maintained, the airspeed will
stabilize at an indication corresponding to the power setting. As power
increases, the rudder is adjusted to maintain the heading. After the air-
craft has stabilized in the straight and level flight attitude, the aviator
retrims to relieve control pressures.
(3) Wind affects a glide by varying ground track and distance
covered, as shown in Figure 9-3. In a head wind, the glide path will be
steeper if the airspeed remains unchanged because the effective head wind
reduces the ground speed of the aircraft by the amount of that component. An
aircraft gliding at 60 knots into a head wind of 60 knots will have a ground
speed of zero and will make a vertical descent. By the same analysis, the
same aircraft gliding with a 60-knot tail wind will have a ground speed of
approximately 120 knots and a proportionately shallower glide path.

Figure 9-3. Effect of wind on a glide


(4) When en route descents are performed, the descent should be
established by reducing power and adjusting pitch attitude to maintain the
cruise airspeed and the desired rate of descent. During the descent, the
airspeed is controlled by adjusting the pitch attitude and power. The rate
of descent will depend on the amount of power reduced. The mixture is ad-
justed as necessary, and the aircraft is trimmed as required throughout the
maneuver.

(5) When slow-cruise descents are performed, power is reduced to a


setting below that required for level flight at slow cruise. Altitude is
maintained while the aircraft is decelerating to slow cruise. As slow-cruise
airspeed is approached, pitch attitude and power are adjusted to maintain the
slow-cruise airspeed and desired rate of descent. The rate of descent will

9-10
depend on the amount of power reduced. The aircraft is trimmed as required
throughout the maneuver.

(6) maximum rate descent is performed, the descent is


When a

established by reducing the power to idle (or minimum allowable) and the air-
craft is configured as recommended in the operator's manual. Pitch attitude
is adjusted to maintain maximum operating speed. To maintain positive
G-forces and properly clear altitudes below, the aviators should establish a
25- to 45-degree bank in the initial descent for at least a 90-degree heading
change. During descent, airspeed is controlled by adjusting pitch attitude.
The aircraft is trimmed as necessary throughout the maneuver.

d. Turns. Turns are used to change the direction of the flight path.
The three general classifications of turns are shallow, medium, and steep.

(1)
Turns are made by moving the aileron control to incline the
A
aircraft's lateral axis toward the horizon, as shown in Figure 9-4. dif-
ference of lift developed on each wing causes the aircraft to bank. This
directs the lift
being developed by the wing toward the center of the turn.
After the desired angle of bank is established, the ailerons should be re-
turned to the neutral position. If the ailerons are not neutralized, the
angle of bank and the rate of turn will continue to increase.

LIFT LIFT

t
w::...-
. -=2

Figure 9-4. Lift


(2) Since drag increases with lift and the ailerons cause a
difference in lift, an aircraft being banked in one direction tends to yaw in
the opposite direction. This yawing tendency, commonly referred to as "ad-
verse yaw," is more noticeable when large, brisk movements are made with the
ailerons. A slow, deliberate movement of the aileron control in the direc-
tion of the turn will reduce the yawing tendency. However, it will consider-
ably increase the time required to establish bank.

(3) Rudder is used to prevent yaw when the aircraft is entering or


recovering from a bank. However, it is not used to turn the aircraft. For
example, if the aileron control is moved to the right, the aviator uses right
rudder until the bank is established. The rudder pressure is then released,
and the rudder is allowed to streamline (part A of Figure 9-5). If bottom
9-11
(inside) rudder pressure is held during the turn, the aircraft skids (part B
of Figure 9-5). If top (outside) rudder pressure is applied, the aircraft
slips (part C of Figure 9-5).

-- -- -- "
~
A
"
" TURN-AND-SLIP
INDICATOR
,
"

.-,\
I
I ,
I ,
B I
.
,
,
,

~
,
.
.
'
'--)
-- \
--,
'\
. ,
...."
---
).,.--,. ...." NEtll
FI.IGI,"t'"
~~
/JF;""-".)
------~- P~"'11

\"
"
I TURN-AND-SLIP

\'
,
-'" INDICATOR
'
\
\ \
, \

'.)
SKID

c
.-,\
I
I
.
I
I
.
.
,
,
~
. .
NEW TURN-AND-SLIP
: '-:) FLIGHT PATH INDICATOR
,'\ \.... -..i
---
-- :-
I , -

,. ..-
/'-;------'''--
~~ -,....---) ,
ORIGIN~I.
"-
'URN
,I
,I
" .-
.'"
\' \
,
,
,
I
,
,
.

,.J SLIP

Figure 9-5. Effect of rudder in turns


(4) An aircraft is correctly banked by coordinated use of ailerons
and rudder. The relative use of these controls is determined by the
9-12
characteristics of the particular aircraft and the airspeed at which the
maneuver is executed. After the aircraft has been banked for a turnt the
angle of bank should be held constant without the aircraft skidding or slip-
ping. The aileron, rudder, and elevator pressures will require some adjust-
mentt depending on the aircraft's characteristics and speed.

(5) turn is started, the aviator begins to increase the back


Once a

pressure on the elevator control to maintain level flight. This increases


the angle of attack and results in more lift; the increase in lift compen-
sates for the loss of vertical lift and the additional weight caused by the
centrifugal force of the turn. The amount of back pressure on the elevator
control is in proportion to the steepness of the bank.
(6) General rules cannot be set for determining in advance the exact
amount of control pressure needed for a specific aircraft. The pressure must
be learned from actual practice.

(7) The stalling


speed of an aircraft is increased in a turn because
of the increase of aerodynamic load from centrifugal force. The increase of
stalling speed is proportional to the square root of the load factor. There-
fore, an aircraft with a normal stalling speed of 60 knots will stall at 120
knots with a load factor of 4 Gs.

(8)
Climbing turns are generally made with a shallow bank. With
the shallow angle, the bank is not steep enough to cause the resultant of
lift to drop appreciably. Since airspeed is less than at normal cruise with
power greater, torque normally tends to turn the aircraft to the left. In
aircraft without trim control, entry in a left climbing turn does not always
require left rudder pedal. Instead, a slight relaxation of pressure on the
right rudder pedal will prevent skidding. When entering a right climbing
turn, the aviator must increase rudder pressure beyond that of a straight
climb.

(9)
Gliding turns generally should be limited to no more than a
medium bank. If the situation requires a steeper bank, the aviator must
remember that the stalling speed increases as the bank increases.

9-5. BANKS

a. Banks With Turns.

(1) This type of bank is shallow at first; as proficiency develops,


the angle is increased. The aircraft is allowed to turn while the aviator
maintains continuous coordinated flight.

(2) A reference line on the ground, normally a straight road or


fence line, is chosen and the aircraft is flown above and in alignment with
it. The aviator starts the first turn in either direction and allows it to
continue until the heading is approximately 45 degrees from the original
heading; at that time he starts a turn in the opposite direction and allows
it to continue for 90 degrees (45 degrees from the original heading in the
opposite direction).

9-13
(3) The entry of the turn, or roll-in, should be smooth and steady
until the desired degree of bank is attained. The change from one bank to
another in the opposite direction is continuous, without pause, at the wings-
level position. The new turn is continued for 90 degrees (45 degrees from
the original heading in the opposite direction). Altitude should be held
constant, and all turns should be fully coordinated. The rollout from each
turn is lead to ensure that the desired heading is not overshot.

b. Banks Without Turns.

(1) To perform banks without letting the


aircraft turn is sometimes
referred to as "Dutch Roll," or "control timing. II Although the aircraft
slips during the maneuver, the aviator develops a perceptual motor skill.
That skill is the ability to perceive flight characteristics or faults and to
respond instinctively with corrective action.

(2) To perform this maneuver, the aviator maintains a constant


heading while rolling the aircraft back and forth from right to left banks
without a pause at the wings-level position. The maneuver is entered by
applying either left or right aileron and opposite rudder pedal in the de-
sired direction of bank. As the degree of bank is reached, a reversal of the
procedure is followed. Altitude is kept constant. The degree of bank may be
increased progressively as the aviator becomes more skilled in performing the
maneuver, or a specified degree of bank may be used. The rate at which con-
trol pressures are applied will affect the coordination of the maneuver. By
combining this maneuver with climbs and glides, the aviator further develops
control touch, timing, and alertness.

9-6. ATTITUDE FLYING

a. Attitude flying is based on the concept that attitude and power will
equal performance. In other words, the attitude of the aircraft in relation
to the horizon and the power applied are the only two elements of control.
Proper use of these two elements of control will produce any desired maneuver
within the capability qf the aircraft. Therefore, all maneuvers must be
based solidly upon attitude and power control references. When flying con-
tact, the aviator uses both inside references (instruments) and outside
references (horizon). When flying under instrument conditions, the aviator
uses inside references only. In either case, the attitude and power settings
are identical for each maneuver. The operation of the controls also remains
the same.

b. The following are some basic facts that each aviator must learn.
(1) For a given power setting, there is a pitch attitude that will
maintain a constant altitude/vertical rate and constant airspeed. A decrease
in pitch attitude results in a loss of altitude and an increase in airspeed.
Conversely, an increase in pitch attitude results in a gain of altitude and a
decrease' in airspeed.

9-14
(2)
For a given pitch attitude, there is a power setting that will
maintain constant
a
altitude/vertical rate and constant airspeed. A decrease
in power results in a loss of altitude and airspeed, whereas an increase in
power results in a gain in altitude and airspeed.

(3) For a given angle of bank, there is a power setting and pitch
attitude that will maintain a constant altitude/vertical rate and constant
airspeed. This pitch attitude will be greater than that required for wings-
level flight. It will increase with an increase in the angle of bank because
of the loss of the vertical component of lift and the apparent increase in
weight that is due to centrifugal force.

c. Attitude and power must be coordinated when aviators perform those


maneuvers that require a constant airspeed and constant altitude/vertical
rate. Power is used to control the airspeed, whereas attitude is used to
control the altitude/vertical rate. For maneuvers requiring a fixed power
setting, such as takeoff, climb, and all power-off maneuvers, the power is
kept constant and attitude is used to control the airspeed or vertical rate,
whichever is required. Either airspeed or vertical rate must be a variable.

d. Through knowledge gained from their experience, aviators learn the


approximate power setting and attitude for each maneuver; they should use
these initially. Then, after the aircraft has stabilized, they make further
adjustments as needed. The following four factors are involved when aircraft
attitude and power are modified or adjusted by the aviator:
0
Time of the adjustment.
0
Rate of the adjustment.
0
Amplitude or extent of the adjustment.
0
Length of time that the adjustment is maintained.

e. Keeping the basic control elements and modifiers in mind, aviators


cross-check to maintain a constant awareness of what the aircraft is doing
at any given moment. A thorough and rapid cross-check of references is an
absolute necessity in good aircraft control. Aviators must also be able to
correctly interpret what they see during the cross-check. From this inter-
pretation, they project what the aircraft is going to do, based on the power
setting and attitude that are being maintained. If a maneuver is being prop-
erly performed, the aviator simply continues to cross-check and interpret.
If he determines changes are necessary, he makes those changes to attitude
and power that cause the aircraft to correctly perform the desired maneuver.

9-7. SLOW FLIGHT

Slow flight serves two purposes. It teaches the beginning student that the
aircraft flies with sufficient control at reduced airspeeds; for the experi-
enced aviator, it
serves as a review and coordination exercise of power-
approach techniques and as a foundation for other flight activities.

9-15
a. Slow flight
may be at any airspeed between
normal cruise and
stalling, but it is usually flown about 5 knots above stalling airspeed.
Pitch and power are used, as necessary, to control
altitude.
b. Slow flight is
entered by reducing power and letting the airspeed
dissipate. Altitude is maintained by increasing the back pressure on the
elevator control as airspeed dissipates. At the point where attitude alone
will not maintain altitude, the required pitch and power are used to control
altitude and airspeed. Turns can then be executed, but the angle of bank
should be shallow because of reduced airspeed. Slow
flight should be prac-
ticed in the cruise and landing configurations during straight and level
flight, climbs, descents, and turns.
c. Maintaining lift and control of an aircraft in flight depends on
keeping a certain minimum airspeed. This critical airspeed depends on the
various circumstances of flight such as gross weight, G-loads imposed by
maneuvering, and density altitude. The closer the airspeed is reduced to
this critical airspeed, the less effective the flight controls
are. The
order in which the control surfaces lose effectiveness is normally the
elevator, ailerons, and rudder.

d. An important
feature of aviator training is the development of the
ability to estimate the margin of safety above the stalling speed by the
diminishing effectiveness of flight controls.

(1)
Slow flight in the cruise configuration needs a higher
pitch
attitude but less power than slow flight with the flaps down. The higher the
pitch attitude for any configuration, the greater the power required. Some
aircraft, because of weight and power available, cannot maintain altitude
with full power. Any attempt to maintain altitude or climb in this
condition
will result in a stall. The proper correction is to reduce pitch attitude
and increase airspeed.

(2)
Flight at minimum airspeed requires positive use of rudder and
ailerons to counteract the asymmetrical loading of the propeller, the action
of the corkscrewing slipstream, and the torque reaction. Rolling in and
out
of turns requires more rudder than rolling at normal airspeeds because of the
greater displacement of the ailerons required. However, banking to the left
normally requires only the relaxing of right rudder pressure, unless trim has
been used to relieve all rudder pressure.

(3) The slower the airspeed, the faster the


rate of turn for any
given bank. Aviators need to remember this, since proper rate of turn
bank angle is part of coordination. As airspeed
for
slows, coordination becomes
more important in avoiding loss of control and a possible
spin. Stall speeds
increase with bank angle; when the aircraft is operating at minimum airspeed,
positive use of power is required in turns to prevent a stall.
(4) As flaps are extended, lift is rapidly increased and pitch
attitude is lowered to prevent a gain in altitude. The faster the airspeed
when flaps are extended, the greater the change in pitch. Conversely, as
flaps are retracted, the pitch attitude is raised to prevent a loss of

9-16
altitude caused by a
loss of lift. The slower the airspeed, the greater the
change in pitch. Flaps should not be retracted at or below a flaps-up
speed.
stall

9-8. STALLS

a. An
aircraft stalls only when it reaches an excessive angle of attack.
The stall
can occur at a reduced airspeed or at any airspeed following an
abrupt change of attitude which causes a high load factor. Figure 9-6 shows
the power-off, power-on, and high-speed stalls. When a
stall occurs, the
aviator must be able to recognize it and take tmmediate and proper corrective
action. Therefore, practice stalls become confidence maneuvers.

(1) An aircraft will stall at the same angle of attack without


regard to pitch attitude, as shown in parts A, B, and C of Figure 9-6. To
initiate a power-off stall (part A, Figure 9-6), aviators reduce the power to
idle and enter a normal glide. The nose is raised to an approximate landing
attitude, which is maintained until the stall occurs. At that time, the nose
of the aircraft will drop below the level flight attitude if aft elevator
pressure is released. Flying speed is regained, and the aircraft tends to
assume a normal glide attitude. Throughout the stall sequence, proper rudder
and aileron coordination is imperative to prevent slipping
or skidding durin~
recovery. If the aircraft stalls with either the left or right wing low, the
rudder application has been improper.

(2)
For the power-on stall (part B, Figure 9-6), power and resultant
thrust require a greater pitch attitude. Normally, a given pitch attitude is
maintained until the stall occurs. As airspeed starts dissipating, the avia-
tor should compensate for the left yaw with right rudder and continue to
increase aft pressure on the elevator control. As the stall occurs, elevato~
pressure is relaxed and the nose of the aircraft starts dropping; as airspeel
increases, the aviator releases the right rudder pressure.
(3) Initially, stall recovery should be effected in a positive
manner by lowering the nose (decreasing angle of attack) and applying power
to regain flying speed. Then, the flight controls are adjusted to attain
straight and level flight at cruise setting. A constant heading should be
maintained throughout the recovery.

(4)
As proficiency in stall recognition and
recovery develops, the
aviator can complete the recovery with little or no loss of altitude. This
should be practiced until recovery can be made at the moment the
initial
stall buffet occurs. As the aircraft stalls, the aviator should relax the
back pressure on the elevator and smoothly apply
full power to increase the
thrust, thereby breaking the stall.
(5) The aircraft can stall in both gliding and climbing
turns.
Therefore, these maneuvers should be practiced. Stall recovery is accom-
plished by lowering the nose of the aircraft to lessen the angle of attack.
The wings are then leveled, and the nose of the aircraft is pulled back
to
level flight attitude. Power can be used to help effect recovery. Recover~

9-17
from turning
a
stall should also be practiced and completed while
the turn is
maintained.

~
RELATIVE W'IIID ~

'-'
.
WIND

~
A. POWER-OFF STALL B. POWER-ON STALL

'-

~-
-

C. HIGH-SPEED STALL

Figure 9-6. Relative stall attitudes

b. As with single-engine aircraft, the aviator should be familiar with


the stall and minimum controllability characteristics of the
multiengine
aircraft being flown. Because of their weight, these larger and heavier
aircraft have slower responses in recovering from a stall and in maneuvering
at critically slow speeds.
NOTE: For the demonstration and practice of various stalls and
recoveries,
refer to the appropriate ATM.

9-18
9-9. SPINS

A spin is described as
a. an aggravated stall that results in what is
termed lIautorotation.1I In autorotation, the aircraft follows a spiral path
in a downward direction. The wings produce some lift,
but the aircraft is
forced downward by gravity; the aircraft wallows and yaws in this spiral
path, as shown in Figure 9-7. It
is assumed that many factors contribute to
a
spin. In fact, the spin is not suited for theoretical analysis.

\
\
'\
'\ ,
,
"

Figure 9-7. Spins

9-19
b. Many aircraft have to be forced to spin; considerable
judgment and
technique are required to start a spin in these
aircraft.
that have to be forced to spin may accidentally be put However, aircraft
into a spin when the
aviator mishandles the controls in turns, stalls, and flight
trollable airspeeds. at minimum con-

c. Once a wing is allowed to drop at the beginning of a the nose


will attempt to move (yaw) in the direction of the low wing. stall,
The aircraft
begins to slip in the direction of the lowered
wing. As it does, the air
meeting the side of the fuselage, the vertical
fin, and other vertical sur-
faces tends to turn the aircraft into the relative wind.
This accounts for
the continuous yaw present in a spin.

(1) At the same time, rolling is also


occurring about the aircraft's
longitudinal axis. This is caused by the lowered wing having
an increasingly
greater angle of attack because of the upward motion of the
against its surfaces. This wing is then well beyond the
relative wind
stalling angle of
attack and, accordingly, has an extreme loss of
wind is striking it at a smaller lift. Because the relative
angle, the rising wing has a smaller angle
of attack than the opposite wing. Thus the rising wing has more
the lowered wing, and the aircraft begins to lift than
rotate about its longitudinal
axis. This rotation, combined with the effects of centrifugal force and the
different amount of drag on the two wings, then becomes a spin. The
descends vertically, rolling and yawing
aircraft
until recovery is effected.
(2) The first corrective action taken during
any power-on spin is to
close the throttles. Power aggravates the spin characteristics,
causing an
abnormal loss of altitude in the recovery. As power is reduced,
full
site rudder is applied. Brisk, positive, straight, forward movement of the
oppo-

elevator control (forward of the neutral position) is then applied. Ailarons


should be neutral, and the controls should be held
firmly in this position.
The forceful movement of the elevator will decrease the
excessive angle of
attack and will, therefore, break the stall. When the stall is broken,
ning stops. This straight, forward position should be
spin-
maintained, and the
rudder should be neutralized as the spin rotation
stops.
(3) If the rudder is not neutralized at the proper time, the ensuing
increased airspeed acting on the fully deflected rudder causes
an excessive
and unfavorable yawing effect. This yawing
effect places tremendous strain
on the aircraft and can cause a secondary spin
in the opposite direction.
(4) Slow and overly cautious control movements
during spin recovery
must be avoided. In certain cases, such movements have caused the aircraft
to continue spinning indefinitely, even when full opposite
controls have been
applied. Brisk and confident operation results in a more positive
recovery.
(5) After the spin rotation has stopped and the
rudder has been
neutralized, back elevator pressure is applied to raise the nose to level
flight. Aviators must be careful not to apply excessive back pressure after
rotation stops. To do so will cause a secondary stall and may result in
another spin that is more violent than the
first.
9-20
d. Accidental stalls and spins can result from improperly executed steep
turns or from increases in the load factor and stalling speed caused by an
increase in bank. When the aircraft is close to stalling speed, a slight
application of rudder may cause an aircraft to spin. If top (outside) rudder
is applied, the aircraft will spin opposite the direction of the turn (over-
the-top spin). If bottom (inside) rudder is applied, it will spin in the
direction of the turn (under-the-bottom spin).

e. Probably the most disastrous of all inadvertent spins occurs when the
aviator turns from the base to the final leg of the traffic pattern. Being
close to the ground, the aviator may be dubious about using a steep bank to
accomplish the necessary rate of turn to align with the runway. He may try
to tighten the turn with the bottom rudder without increasing the bank. This
causes a skidding turn that leads to a violent under-the-bottom spin. Con-
versely, if outside rudder is used to decrease the rate of turn, a slip re-
sults. If a stall occurs during this slip, an over-the-top spin can result.
For a safe turn to be accomplished, airspeed must be kept well above stalling
and controls must be coordinated at all times.

NOTE: Accidental stalls and spins are not limited to turning situations;
they may occur in any flight attitude.

f. Anytime
spin is encountered, regardless of the conditions, the
a

normal spin recovery sequence should be as follows:


0
Retard power.
0
Apply opposite rudder to slow rotation.
0
Apply positive forward-elevator movement to break the stall.
0
Neutralize the rudder as spinning stops.
0
Return to level flight.
0
Do not use power.

9-10. TRAFFIC PATTERNS

a. traffic pattern expedites


The the flow of air traffic about an
airfield with safety and regularity. It
is the designated flight path for
landings and takeoffs at an airfield. The traffic pattern requires the avia-
tor to simultaneously watch for other aircraft, receive and acknowledge tower
signals, and maintain aircraft control. Normally, the flight path is
rectangular-shaped, as shown in Figure 9-8. The active runway forms one
side, and all turns made to the left form the other three sides. The sides
are referred to as legs--takeoff, crosswind, downwind, base, and final
approach. The pattern is usually flown 1,500 feet above the terrain. The
first turn after takeoff may be started before the aircraft reaches 1,500
feet, but it should never be started before 500 feet is reached or the field
boundary is crossed, whichever comes last. When other traffic is in the

9-21
pattern, the takeoff leg may be extended to acquire adequate spacing between
aircraft. The climb is continued to the designated altitude, and climb power
is maintained until the aircraft accelerates to the prescribed downwind
air-
speed. The turn from crosswind to downwind should be executed so that the
aircraft is placed at the proper distance from the runway. Depending on the
aircraft and the traffic pattern altitude, the aircraft's proximity to the
runway may range from 1,000 to 5,000 feet; 2,000 feet is
usually adequate.
b. Entry into a traffic pattern normally is made by flying toward the
field at a 45-degree angle to the downwind leg. This entry reveals the avia-
tor's intentions and affords excellent visibility for checking the area for
other aircraft in the pattern. As the downwind leg is reached just short of
the field, the aircraft is turned 45 degrees to fly the downwind
leg.

/
~~y
f(,~~i,ß

DOWNWIND
/~
LEG

c
z
w
0 j
~ W
III
0
W
...I
a:I III ...I
0
a:
(J

FINAL TAKEOFF
ACTIVE RUNWAY
APPROACH

. WIND

Figure 9-8. Traffic pattern


9-11. APPROACHES

The 180-degree side approach is normally used.


It
involves two turns of 90
degrees each. On the downwind leg of the traffic pattern at a point opposite
the point of intended landing, the aviator reduces power and establishes a
descent. Multiengine aircraft characteristically have steeper gliding angles
because of their relatively high wing loading with greater drag of wing flaps
and the landing gear when extended. For this reason, power normally is used
throughout the approach to shallow the approach angle and to prevent a high
rate of sink. The turn to the base leg should be executed at an altitude
that will assure the turn to final can be completed at or above 500 feet AGL.

9-22
The wind condition and the distance from touchdown determine the point of
initiating the turn to final and the angle of bank.
9-12. SLIPS

A slip iscombination of forward and sideward movement; the lateral


a
a.
axis is inclined forward and the longitudinal axis moves sideways toward the
lower side. Slips can be used for crosswind landings and for dissipating al-
titude without increasing forward speed. Aviators control them to make the
aircraft move sideways or to continue the initial flight path over the
ground. They can also be used during a turn.

b. To establish a slip from normal glide, the aviator lowers the wing
toward the desired direction of slip by using the ailerons. Opposite rudder
is applied simultaneously to control the desired ground track, which mayor
may not require a change of heading. Normal glide speed should be maintained
during this slip. Because of the stress placed on the aircraft, prolonged
flight in this out-of-trim condition should be avoided.
9-13. LANDINGS

a. The landing is accomplished into the wind to effect a slower ground


speed at touchdown. Aircraft attitude is changed from normal glide to land-
ing attitude, as shown in Figure 9-9. Rate of descent is decreased until the
aircraft flies parallel with the surface of the runway. This change of both
attitude and flight path is called the roundout.

WIND t
~
~
..,...,

I//I////~//////////////////////////////////////I///I//11I1
Figure 9-9. Normal landing

b. The roundout must be started and timed to attain a landing attitude


and touchdown almost simultaneously. The application of back pressure on the
elevator control increases pitch attitude and dissipates airspeed. Back
pressure is continued smoothly, with only enough force applied to prevent the
aircraft from striking the ground before a landing attitude is obtained. At
this time, the aircraft should be only inches from the runway and almost
instantaneously making ground contact. Aircraft with conventional gear
should be stalled at the same time ground contact is made.

9-23
Because touchdown is made on the main
c. gear, a tricycle landing gear
generally simplifies a
landing. The nosewheel is off the ground when the
m~in gear touches. The location of the CG
forward of the two main wheels
combines with the forward momentum of the
aircraft to help hold it on the
ground. This condition tends to keep the
path.
aircraft rolling in
a
straight

d. If the roundout is too fast, the


gain of altitude. Since airspeed is
aircraft will "balloon," causing a
dissipating, the aircraft may stall at a
considerable height above the ground.
If ballooning occurs, the aviator must
compensate by relaxing the back pressure to
bring the aircraft back to the
runway in a landing attitude before a stall
occurs. Power may need to be
applied to prevent a stall at an unsafe
altitude.
e. If
the roundout is too slow, the aircraft
strikes the ground while
descending. The aircraft either bounces back into the
damaging the landing gear, or crashes. air, after possibly
If the aircraft touches the runway
before reaching the desired landing attitude with
descent unchecked, it will
ricochet off the ground with results similar to ballooning.
Recovery is
effected by the application of power.

f. The aviator must maintain directional control of the aircraft after


wheel contact with the runway. Ground
steering is accomplished by means of
the rudder, steerable nosewheel or
tailwheel, and brakes. If the available
runway permits, the speed of the aircraft is allowed
to dissipate with normal
ground friction and drag. Aviators use the
brakes, if needed, to help slow
the aircraft. When they use the brakes, they
apply back
vator control to prevent the nosing over of conventional pressure on the ele-
keep excessive weight from being exerted gear aircraft or to
on the nose gear of tricycle
aircraft.
Successful landings under crosswind conditions are
g.
on the largely dependent
aviator's ability to fly
by sensory perceptions. The aviator must be
able to recognize drift and then make
corrections. The crosswind-landing
technique in multiengine aircraft is not very different
from that required in
single-engine aircraft. The only significant difference
the greater weight of multiengine is that, because of
aircraft,
must be maintained before touchdown.
more positive drift correction
Although crosswind capabilities differ
for single-engine and multiengine aircraft, corrective methods
The two methods used are the wing-low are identical.
(slip) and crab or a combination of the
two.

(1) Wing-low (slip) method. The wing-low method


is a
method is accomplished by lowering the upwind wing into the windslip. This
to compensate for drift, as shown in Figure 9-10. The just enough
tablished during final approach. slip is normally es-
out the approach and roundout. The
It is then maintained as necessary through-
aircraft will initially touch down on the
upwind main wheel. As speed decreases, the
main wheel.
aircraft settles onto the other

9-24
~fliI
------------ ~
Figure 9-10. Crosswind landing--slip
(2) Crab method.
The crab method, as illustrated in Figure 9-11, is
more difficult,
since better timing is needed during the roundout. The avia-
tor maintains required crab for the wind condition throughout the final ap-
proach. Just before touchdown, he removes the crab and aligns the aircraft
with the runway. If the crab is not completely removed before touchdown,
side loads will be imposed on the landing gear that can result in gear damage
or loss of control. If the crab is removed too soon, the aircraft will drift
off course.

~~
(2)

~~ (3)
Figure 9-11. Crosswind landing--crab

Combination of methods. A combination of the methods in (1) and


may be employed--the crab for the approach and then a slip for roundout
and touchdown.

NOTE:
This combination is usually preferable.

For the description and practice of normal landings and power


approach/precision landings, refer to the appropriate ATM.

9-25
9-14. GO-AROUNDS

procedure for remaining airborne after an intended landing


A go-around is a

is discontinued. Go-arounds are used--

control tower landing clearance is withheld.


0
When

0
When the approach does not develop as planned.
0
When other traffic does not permit a
safe landing.

Go-arounds in a controlled traffic pattern may be initiated upon


a.
instructions from the tower. They are limited to maintaining or regaining
Alertness for
pattern altitude and spacing according to local traffic rules.
other traffic is of prime importance.

b. Most go-arounds are on the timely decision of the aviator


initiated
landing
at any time before completion of the A roll. Both the decision and the
implementing action must be timely. late start may result in failure to
clear a barrier or in an accidental stall in the climb.

c. Because go-arounds frequently occur in


limited or confined areas,
an accurate awareness of aircraft and aviator performance
limitations is
and altitude
required. The problem is basically one of building up
speed
with either
within the limitations imposed by available power. Preoccupation
factor will result in a loss of the other. Maximum performance requires a
thorough familiarity with the aircraft's performance under
a
variety of
atmospheric conditions, flap configurations, and airspeeds.

d. Go-arounds should not be used to salvage


a bounced, dropped-in, or
crabbed landing attempt. Such situations sometimes involve directional con-
considerable space before
trol and stall recovery problemS that may consume damage
the climbout can be started. Serious structural is possible in any
hard ground contact. Therefore, may be it
wiser to land rather than risk
further flight in a weakened or uncontrollable aircraft.
NOTE: For the description,and practice of go-arounds, refer to the
appropriate ATM.

9-26
CHAPTER 10

FIXED-WING MULTIENGINE OPERATIONS

Several types and models of light twin-engine aircraft are used


Army
in performing
training and operational missions. Primarily, this chapter explains the
most prominent flight characteristics of
twin-engine, fixed-wing aircraft.
The appropriate aircraft operator's manual should be
consulted for rotary-
wing, multiengine operations.

10-1. LIGHT-TWIN AIRCRAFT PERFORMANCE

The term light twin is used to define Army


propeller-driven aircraft that
have a maximum certificated gross weight of less than 12,500
pounds and one
engine mounted on each wing. The basic difference between a
light twin-
engine and a single-engine aircraft is the potential problem of
one of the light twin engines.
failure of
The following information is about that
difference.

a. Performance and Operating Speeds. Certain aircraft performance


operating limitations are based on airspeed. These airspeeds
(A listing of V speeds are called V
speeds. is found in the Glossary). In addition to
the performance and operating speeds common to single-engine and
engine aircraft, the mu1tiengine aircraft aviator must become
light twin-
familiar with
some additional V speeds. They are Vme, Vx,
Vxse' Vy' and Vyse and are de-
fined in the paragraphs that follow:

(1) Vme. Vme is the minimum airspeed at which an


controllable
aircraft is
whenthe critical engine suddenly becomes inoperative and the
remaining engine produces takeoff power. The FAR, under which the
was certificated, states that at Vme the aircraft
certificating test pilot must be
able to stop the turn that results when the
critical engine suddenly becomes
inoperative. Using maximum rudder deflection and no more than
bank into the operative engine, the as-degree
test pilot must stop the turn within 20
degrees of the original heading. The FAR also states that
the certificating test pilot must maintain the after recovery,
aircraft in straight flight
with not more than a 5-degree bank (wing lowered toward the
operating
engine). This means that the aircraft must maintain a heading,
must be able to climb or hold altitude. The
not that it
principle of Vme is that at
airspeeds less than Vme' air flowing along the rudder is such
that the
plication of rudder cannot overcome the combined effects of asymmetricalap-
yawing caused by takeoff power on one engine and a
powerless windmilling
propeller on the other engine.
(2) Vx. Vx is the speed that provides the best angle of climb. At
this speed, the aircraft gains the greatest height for a given
distance. Vx
is used for obstacle clearance when all engines are operating. However, when
one engine is inoperative, Vx is referred
to as Vxse (best angle of climb,
single engine).

10-1
(3) provides the best rate of climb. This
Vy is the speed that
speed
~.
provides the for a given period when all engines
maximum altitude gain
Vy is referred to as Vyse
are operating. When one engine is inoperative,
(best rate of climb, single engine).

Single-Engine Operation. Many aviators erroneously believe


that a
b.
light-twin engine aircraft will continue to perform at least half as well
when only one of its engines is operating. Part 23 of the FAR that governs
that the
the certification of light twin-engine aircraft does not require
takeoff configuration with one engine
aircraft maintain altitude while in the not
inoperative. In fact, some civilian light twin-engine aircraft are
in any configura-
required to maintain altitude with one engine inoperative
the operation of light twin-
tion, even at sea level. This is important in 23 of the FAR. Regarding
engine aircraft that are certificated under part in a
performance rather than controllability, most light twin-engine aircraft
takeoff or landing configuration are merely
a single-engine aircraft with its
power divided into two individual units.

When one of the twin engines fails, aircraft


c. ~limb Performance.
performance is reduced by 80 percent or more. The loss of performance is
function
percent because the aircraft's climb performance is
a
more than 50
which is available power in excess of that required
of the thrust horsepower,
is added to the engines than is needed for
for level flight. When more power The of climb depends on
straight and level flight, the aircraft climbs. rate
above that required
the amount of excess power that is added, which is power
is lost and drag increases
for level flight. When one engine fails, power
the full bur-
because of aSYmmetric thrust. The operating engine must carry
This leaves the
den by producing 75 percent or more of its rated power.
climb performance. When one of its
engine with very little excess power for 1,860 FPM
of climb of
engine fails, an aircraft that has an all-engine rate
190 FPM loses almost 90 percent of its
and a single-engine rate öf climb of
climb performance. During straight and level flight, the light twin offers
However, the
obvious safety advantages over the single-engine aircraft.
by that second engine in the takeoff
aviator must know the options offered
and approach phases of flight.

10-2. ASYMMETRIC THRUST

The asymmetric thrust, or unequal engine thrust, in multiengine aircraft


is
To achieve
the principal flight characteristic that must be counteracted.
most
the desired stability during power changes, manufacturers position
engines so that the thrust line passes through or near
single-engine aircraft
the CG. In conventional twin-engine aircraft, only the resultant
thrust of
both engines provides this stability. When both engines are not operating
at
equal power, aSYmmetric thrust results and causes movement
about the vertical
axis, or yaw. The rudder is used to prevent this movement. If yaw occurs,
the aviator must
the aircraft may also roll or bank. To regain level flight,
apply both the rudder and. the aileron.

10-2
10-3. CRITICAL ENGINE

a. The P-factor, or asymmetric propeller thrust, is present in the


light-twin aircraft just as it is in single-engine
the dissimilar thrust of the rotating aircraft. It is caused by
propeller blades during certain flight
conditions. The P-factor Occurs when the relative wind
striking the blades
is not aligned with the thrust line, as it is with a nose-high
attitude. As
a
result, the downward-moving blade has a greater angle of attack than the
upward-moving blade.

b. In Army light twin-engine aircraft, both engines rotate clockwise


when they are viewed from the
rear and both engines develop equal thrust.
With positive angle of attack, low airspeed, and high-power conditions, the
a

downward-moving propeller blade of each engine develops


more thrust than the
upward-moving blade. In part, this explains why conventional fixed-wing
aircraft pull to the left on takeoff. The center of thrust of the propeller
disk shifts to one side when the thrust line is
tilted upward.
c. As indicated by lines D1 and D2 in Figure 10-1, the P-factor
in a center of thrust at the right side of each engine. The yawing results
the right engine is greater than that of the force of
left engine. As indicated by
line D2 in Figure 10-1, the center of thrust of the right engine has a longer
lever arm and is farther away from the centerline of the fuselage. There-
fore, when the right engine is operative and the left engine is inoperative,
the yawing force is greater than
it is when the left engine is operative and
the right engine is inoperative. In an engine-out
demand on the rudder is made when the
situation, the greatest
operative engine is the one on which
the downward-moving blade is farther from the fuselage (the
right engine).
Therefore, the left engine is the critical engine; its loss presents the
greatest controllability problem.

ARM # , ARM
~ /
OPERATIVE INOPERATIVE INOPERATIVE OPERATIVE
ENGINE ENGINE ENGINE ENGINE
(CRITICAL ENGINE)

Figure 10-1. Forces created during single-engine operation

10-3
10-4. MINIMUM SINGLE-ENGINE CONTROL SPEED

Asymmetric Thrust Control. Maximum asymmetric thrust


is created
a.
when the critical engine, which is usually the left one on Army aircraft, is
adequate airspeed is
inoperative and the other one is at takeoff power. If Below this air-
maintained, yaw can be prevented by applying the rudder.
be maintained by reducing power. Each
speed, directional control can only
aircraft's critical airspeed is identified as Vme' Vme is the minimum speed
at which the critical engine can inoperative and straight flight
be rendered
maximum available
continued with the remaining engine operating at takeoff or
and the aircraft is
power. When the critical engine is rendered inoperative
that the aviator
in the most unfavorable flight configuration, Vme ensures
can stop the turn and maintain the new heading. Vme applies only to the
can be main-
control of asymmetric thrust; it does not ensure that altitude low a Vme as pos-
tained or that a climb can be accomplished. To achieve as
To determine
sible, a 5-degree bank is always used in flight testing. so that when an
Vme,
airspeed that is low enough
the test pilot arrives at an
Full
engine is cut an immediate bank into the operative engine is required.
bank provide the necessary control to keep
rudder deflection and the 5-degree
dead engine. Avia-
the aircraft from turning more than 20 degrees into the
tors should refer to the appropriate
ATM to practice flight at Vme'

Certification. The configuration required for obtaining


a
b. Vme
manufacturer's Vme certification follows:

Landing gear is retracted.


0

Aircraft is trimmed for takeoff.


0

Flaps are set to takeoff position.


0

0
Takeoff or maximum available power is attainable.
0
Rearmost allowable center of gravity exists.
0
Maximum sea level takeoff weight is maintained.
0
Cowl flaps on piston-engine aircraft are in the position normally used
for takeoff.

Propeller is windmilling or feathered if the aircraft has an auto-


0

feather system on the inoperative engine and if full power exists


on

the other engine.

Additionally, rudder control force required to maintain control must not


exceed 150 pounds.

are in a position less than


c. Bank Angle. If the wings of an aircraft the value shown in
a 5-degree bank angle, Vme is substantially higher than
the difference
the flight manual. On most Army light twin-engine aircraft,
in Vme between the 5-degree bank condition and the wings-level condition may

10-4
be as high as 15 knots. The complex reasons for this large increase in Vrne
with varying bank angles are discussed below.

(1) The effect of bank reduces the amount


of -rudder power required
to overcome the asymmetric thrust condition. As the
wings brought to a
level position, more rudder is necessary. At a given rudderare
rudder-pedal force, a higher airspeed is required. Although deflection, or
this charac-
teristic applies to all light twin-engine aircraft, it is accentuated in the
latest designs because of the amount of power or thrust
available for take-
off. In addition, the thrust lines of the engines are located
the wingspan, which increases the turning moment farther out on
caused by the unbalanced
thrust condition.

(2) To achieve the best


performance when an engine
takeoff, climb, or any other flight condition when high powerfails
during
the aviator must keep the aircraft in a is required,
5-degree bank attitude with the
inoperative engine on the high side. The normal takeoff
procedure ensures
that the airspeed will be above Vrne when the most
critical engine is inopera-
tive. However, this is true only if the 5-degree bank angle is maintained.
d. Control Problems. The following paragraphs discuss
associated with engine failure. Aviators must be aware of control
problems
these problems and
learn proper procedures to correct them.
(1) When an engine stops, many aviators instinctively
the ball. They do not understand how a light twin functions
try to center
with asymmetric
thrust.
(2) Drag normally acts around a
point along the
centerline of the
aircraft fuselage. When the propeller is windmilling or feathered, the cen-
ter of drag moves toward the dead engine. The operative engine
pull along a line several feet to the side of the center of exerts its
drag. This
causes the aircraft to rotate toward the
inoperative engine. The aviator
can prevent this rotation in one of two ways, which are discussed below.

(a) The aviator can cut the power on the


quickly regain control of the aircraft because operative engine and
it is in a symmetrical power-
off glide. Unless the aircraft is about to go out of
control, this is not a
desirable option immediately following takeoff or during low altitude flight.
(b) The aviator can use as much power
from the operative engine
as possible to maintain a safe single-engine flying speed. This
stopping the rotational movement with the rudder, which requires
causes the aircraft
to skid toward the inoperative engine. To
correct the skid, the aviator must
bank into the good engine to maintain
the longitudinal axis parallel to the
relative wind.
(3) A
variety of rudder and
aileron combinations can be used to
maintain heading. Most aviators
try to center the ball in a
attitude by raising the aileron on the operative engine side wings-level
of the aircraft.
This compensates for the additional
lift produced by the propeller slipstream
passing over that wing. When it is viewed from outside
the aircraft,

10-5
however, the fuselage is not aligned with
the direction of flight, or

relative wind; it is yawed toward the operative engine.


(4) aviators believe that during coordinated flight their This
Many
aircraft flies straight through the air without slipping or skidding.
may be true in a single-engine aircraft or
a light twin with equal power on
both sides. However, when one engine stops and power is off-center,
this is
not true.

(a) the ball is precisely centered during wings-level


When
coordinated flight, light twin-engine aircraft with one engine out a will fly
a

with a
large sideslip. This is shown in part
A of Figure 10-2. If piece
windshield of the aircraft, the string
of string were taped to the nose or
Single-engine rate of climb declines
would lean toward the operative engine.
or disappears and Vme increases.
(b) When manufacturers run a performance test, they use precise
and maximum perfor-
sideslip-indicating instruments to assure zero sideslip
has no way of knowing the
mance. Without these instruments, the aviator
sideslip occurs when
sideslip angle. Most aviators mistakenly assume zero
the wings are level and the ball is centered.

(c) As shown in part B of Figure 10-2, zero sideslip occurs in


3 to 5 degrees
most light twins when the aircraft is banked approximately
into the operative engine. Although it is disturbing to many aviators, the
A yaw string will show,
ball will be off-center toward the good engine.
however, that airflow is straight along the nose,
which is the proper airflow
for minimum drag and maximum performance.

B
A ~ ~
!! ~!! j1 l_!f ~ 11
+

~
~
Figure 10-2. Sideslip

10-6
10-5. SINGLE-ENGINE CLIMBS

a. Climbs are made with reserve, or excess, power. Reserve power is the
power available that is not required to maintain level flight. With one
engine shut down, a twin-engine aircraft will not have an abundance of
reserve power under the most favorable circumstances. Any change from the
best rate-of-climb and angle-of-climb speeds above or below the best
single-
engine climb speed rapidly decreases climb performance. The operator's
manual establishes the best angle-of-climb and rate-of-climb speeds.

b. To provide adequate power for single-engine climbs, drag should be


reduced to a minimum. Drag can be reduced by retracting the landing
gear,
raising the flaps, and feathering the propeller of the inoperative engine.
Single-engine best rate-of-climb speed is the most efficient single-engine
operating, or Vyse, speed. If altitude cannot be gained at this speed, more
power must be obtained or drag or weight must be reduced.

10-6. SINGLE-ENGINE LEVEL FLIGHT

a. Maintaining level flight with one engine inoperative is possible only


below the single-engine absolute ceiling. This ceiling is based on standard
atmosphere at sea level conditions. The operating engine must be at maximum
continuous power, the aircraft must be at maximum gross weight, the gear and
flaps must be up, and the inoperative propeller must ,be feathered. In addi-
tion~ the aircraft must be at a zero sideslip angle. High density altitude
or failure of the propeller to feather reduces the ceiling. Airspeed is also
a
factor in maintaining level flight; the best single-engine rate-of-climb
speed provides maximum efficiency. The operator's manual contains the power
settings and ceilings for both normal and single-engine cruise flight.
b. Power charts normally provide cruise information for aircraft
operating at 45 to 75 percent power. To supply the necessary power for con-
tinued flight, one engine may be required to operate above the recommended
cruise range. The loss of one engine creates an emergency; therefore, flight
should not be continued beyond the nearest suitable
airfield. Normally, trim
controls can relieve control pressures when the aircraft is operating with a
single engine. Some operator's manuals recommend a bank of no more than 5
degrees toward the operating engine during straight flight. Bank reduces the
amount of rudder required to counter drag and asymmetric thrust and to reduce
sideslip. The degree of bank should be confined to recommended amounts.

10-7. SINGLE-ENGINE DESCENTS

Usually, descent in aircraft powered by reciprocating engines is performed at


a
specified power setting and airspeed. Correct power setting and airspeed
help retain minimum engine operating temperature and reduce plug fouling or
engine loading. Aviators should avoid making descents at idle power for
prolonged periods. One inch of manifold pressure for each 100 revolutions
per minute is the general rule for descent power. Because of the low power
requirements involved, en route descents with one engine inoperative are not
a
problem. However, descent for landing is more involved and requires

10-7
caution. For specific information, the engine power charts in the operator's
manual should be consulted.

10-8. SINGLE-ENGINE APPROACH AND LANDING

When both engines are operating normally, no special technique or skill is


required to perform an approach and
a
landing. However, performing the
approach and landing with one engine inoperative demands more skill
and
be used for
judgment. When possible, normal patterns and speeds should
To preclude loss of directional control and
single-engine approaches.
provide for the best rate of climb, speeds above Vyse. should be maintained
during an approach until the landing is assured or during
a
go-around.
according to the aircraft configuration and
Approach and landing speeds vary
the type of approach. The operator's manual supplies this information.

10-9. PROPELLER FEATHERING

When an engine fails in flight, the aircraft's movement through the


a.
air keeps the propeller rotating. failedThe engine no longer delivers power
the which produces thrust. Now the propeller is absorbing
to propeller,
The drag of
energy to overcome friction and the compression of the engine.
the windmilling propeller is significant. As shown in Figure 10-3, the drag
of the windmilling propeller causes the aircraft to yaw toward the failed
engine. To minimize the yawing tendency, all Army multiengine aircraft are
equipped with full-feathering propellers.

b. The aviator can position the blades of


a
feathering propeller to such
a high angle that they are streamlined in the direction of flight. In this
feathered position, the blades are streamlined with the relative wind; there-
fore, they stop turning. This significantly reduces drag on the aircraft.
A feathered propeller creates the least possible drag on the aircraft and
reduces its yawing tendency. As a result, multiengine aircraft are easier to
control in flight when the propeller of the inoperative engine is feathered.
In addition, a feathered propeller causes less damage to the engine. Feath-
ering a propeller should be demonstrated and practiced in all aircraft
equipped with propellers that can be safely feathered and unfeathered during
flight.

~
'\

Figure 10-3. Windmilling propeller creating drag

10-8
10-10. ACCELERATE-STOP DISTANCE

a. During the two or three seconds immediately following the takeoff


roll, an aircraft accelerates to a safe engine-failure speed. This is the
most critical time fortwin-engine aircraft should an engine-out condition
a

occur. Army twin-engine aircraft are controllable at a speed close to the


engine-out minimum control speed. However, their performance is often so far
below optimum that continued flight following the takeoff may be marginal or
impossible. A more suitable and recommended speed, which some aircraft manu-
facturers call minimum safe single-engine speed, is the speed at which
altitude can be maintained while the landing gear is being retracted and the
propeller feathered.

b. When one engine on a twin-engine aircraft fails on takeoff after


having reached the safe single-engine speed, it loses approximately 80 per-
cent of its normal power. The twin-engine aviator, however, has an advantage
over the single-engine aviator. If the twin-engine aircraft has the single-
engine climb capability at the existing gross weight and density altitude,
the aviator can either stop or continue the takeoff. The single-engine
aviator has only one choice--Iand the aircraft.

c. If
one engine fails before the aircraft reaches Vme' the aviator has
no choice but to close both throttles and bring the aircraft to a stop. If
engine failure occurs after the aircraft becomes airborne, the aviator must
immediately decide whether to land or to continue the takeoff. If the avia-
tor decides to continue the takeoff, the aircraft must be capable of gaining
altitude with one engine inoperative. If no obstacles are involved, the
aviator must accelerate to Vyse. If obstacles are a factor, the aviator must
accelerate the aircraft to Vxse.

d. To make a correct decision in this type of an emergency, the aviator


must consider runway length, field elevation, density altitude, obstruction
height, head wind, and the aircraft's gross weight. For simplicity, runway
contaminants, such as water, ice, snow, and runway slope, are not discussed
here. The flight paths shown in Figure 10-4 indicate that the area of
decision is bounded by the point where Viof is reached and the point where
Vyse is reached. An engine failure in this area demands an immediate deci-
sion. If engine failure occurs beyond this decision area, the aircraft can
usually be maneuvered back to a landing at the departure airport if it is
within the limitations of engine-out climb performance.

10-9
ENGINE FAILURE ON TAKEOFF

k-
G
I $
"~I~ .\

Figure 10-4. Decision area

and the accelerate-after-lift-off


e. The accelerate-stop distance
distance are based on the assumption that an engine fails at the instant Vlof
is attained. The accelerate-stop distance is the total distance required to
accelerate the twin-engine aircraft to Vlof and bring it to a stop on the
remaining runway. The accelerate-after-lift-off distance is the total dis-
tance required to accelerate the aircraft to Vlof and continue the takeoff
50 feet.
on the remaining engine to a height of

10-11. ENGINE-OUT PROCEDURES

of Army light
a. Engine Failure. Flight performance characteristics At times,
twin-engine aircraft with one engine inoperative are excellent.
however, their performance can be marginal. As long as sufficient airspeed
is maintained, light twins can be controlled and maneuvered safely. However,
to use the safety and performance characteristics of light twins effectively,
the aviator must understand single-engine performance and limitations that
result from an unbalance of power.
(1) When an engine fails after the aircraft becomes airborne, the
aviator should hold the heading with the rudder and simultaneously roll into
a bank of at least 5 degrees toward the operating engine. In this attitude,
the aircraft tends to turn toward the operating engine. At the same time,
the asymmetrical power that results from the engine failure tends to turn the
aircraft toward the inoperative engine. The result is
a
partial balance of
the tendencies that provide an increase in aircraft performance and easier
directional control.
NOTE: In the above situation, the ball in the turn-and-bank indicator will
be approximately one ball width off center toward the operative engine.

(2) The best way to identify an inoperative engine is to note the


To
direction of yaw and the rudder pressure required to maintain heading.
counteract the asymmetrical thrust, extra rudder pressure must be exerted on
the operating engine side. To help identify the failed engine, some aviators

10-10
use the terms "best foot forward" or "dead
foot, dead engine." They must
never rely on tachometer or manifold pressure readings to
~etermine which
engine has failed. After an engine loses power, the tachometer
often indi-
cates the correct RPM and the manifold pressure gage
indicates the approxi-
mate atmospheric pressure.

(3) Experience shows that the biggest problem


is the aviator's
actions after the inoperative engine has been identified. For
example, an
aviator may identify the inoperative engine but attempt to shut down the
wrong one. This results in no power at
all. To avoid this mistake, the
aviator should verify which engine is dead by retarding the throttle of the
suspected engine before shutting it down.

b. Engine Failure During Takeoff.

(1) If failure occurs during the takeoff roll before the


engine
aircraft becomes
airborne, the aviator must close both throttles immediately,
which will bring the aircraft to a
stop. The same procedure is recommended
if, after the aircraft becomes airborne, an engine fails before reaching
Vyse. Because of the altitude loss that is required to
increase the speed to
Vyse, an immediate landing is usually inevitable.

(2)Before takeoff, the aviator determines what


and
altitude, airspeed,
aircraft configuration must exist to permit the flight to continue in the
event of an engine failure. If engine failure occurs before
these factors
are established, the aviator must realize that both throttles must be closed
and the situation treated the same as an
engine failure on a single-engine
aircraft. If the engine-out rate of climb under the existing circumstances
will be at least 50 FPM at 1,000 feet above the airport and Vxse has been
attained, the takeoff can continue. However, if airspeed is below Vxse and
the landing gear has not been retracted, the takeoff
must be abandoned
immediately.

(3)
If Vxse has been attained and the landing gear is in the retract
cycle, the aviator should climb at Vxse to clear any obstructions. After the
decision is made to continue flight, airspeed should be
the flaps retracted, and all appropriate systems
stabilized at Vyse,
reset. Vyse results in the
slowest rate of descent and provides the most time
for executing an emergency
landing. Therefore, even if altitude cannot be maintained, the
should continue to hold aviator
Vyse,' After the decision is made to continue the
flight, the landing gear should be retracted as soon as practical.
(4) If
the aircraft is barely able to maintain altitude and
airspeed, turn requiring a bank greater than 15 degrees should not be
a

tempted. When such a turn is made under these conditions, both at-
airspeed decrease. Therefore, the aviator should continue lift and
straight ahead
until the aircraft reaches a safe maneuvering altitude. Then a steeper bank
can be made safely in either direction.
If a safe speed and zero sideslip
are maintained, the aircraft can be banked toward the inoperative
For demonstration and practice of engine engine.
failure during takeoff, refer to
the appropriate ATM.

10-11
c. Engine Failure During Cruise Flight.

(1) Normally, when engine failure occurs during cruise flight, the
Under these less
situation is not as critical as engine failure on takeoff.
the cause
stressful circumstances, the aviator should take time to determine be
of the failure and correct the condition. If the condition cannot cor-
procedure should be accomplished and a
rected, the recommended single-engine
landing made as soon as practical.

(2) During engine an aviator tends to perform the


failure,
This error results in
engine out identification and shutdown too quickly.
improper identification of the failed engine or incorrect shutdown proce-
with actual
dures. The element of surprise, which is generally associated
engine failure, may result in confused and hasty reactions.

(3) When an engine fails during cruise flight, an aviator's main


problem is maintaining sufficient altitude to continue flight to
a suitable
depends on
airfield for landing. Whether the aviator can maintain altitude
density altitude, aircraft gross weight, terrain elevation,
and obstructions.
When the aircraft is above its single-engine service ceiling, it loses al-
The single-engine service ceiling is the maximum density altitude
titude. 50 FPM
at which the single-engine best rate-of-climb speed will produce a

rate of climb. The manufacturer determines this ceiling based on the maximum
gross weight of the aircraft with its flaps and landing gear retracted, the
critical engine inoperative, and the propeller feathered.

(4) Although en route engine failure in normal cruise conditions may


not be critical, maximum permissible power should be added to the operating
engine before securing or shutting down the failed engine. If
the aviator
on the operating engine is
later determines that maximum permissible power
not needed to maintain altitude; power can be reduced. Conversely, if maxi-
mum permissible power is not applied, the airspeed may decrease much further
and more rapidly than expected. This condition could present a
serious per-
below
formance problem, especially if the airspeed drops Vyse.

(5) is within the capability of the aircraft, the aviator


If it
should maintain altitude. 'If an aircraft with an inopérative engine cannot
should be maintained
maintain altitude under existing circumstances, airspeed
at Vyse to conserve the remaining altitude and reach
a suitable landing area.

After the landing gear and the flaps are retracted, the failed
(6)
engine is shut down and heading and altitude are under control,
the aviator
should advise the nearest ground facility that flight is being conducted with
one engine inoperative. FAA air traffic control facilities can give valuable
IMC or when a
assistance, particularly when the flight is conducted under
landing is to be made at a tower-controlled airport. Rather than continuing
the flight, good judgment dictates that a landing be made at the nearest
suitable airport as soon as practical.
(7) 'During engine-out practice in aircraft with reciprocating
engines, the engine may cool to temperatures below
the normal operating range
when the
if zero thrust power settings are used. This requires caution
10-12
.

aviator advances power at the termination of single-engine practice.


power is advanced rapidly, the engine might not respond and
theIf
an actual engine
failure could result. This is particularly important when engine-out ap-
proaches and landings are being practiced. A good
procedure is to advance
the throttle slowly, which allows the engine to respond and
stabilize. This
procedure also results in less wear on the engines of training
aircraft.
(8) To avoid engine damage, restarts after the
propeller has been
feathered require the same amount of care. Following a
should be maintained at, or slightly above, the idle
restart, engine power
setting until the engine
is warm and lubricated. For demonstration and practice of engine failure
during cruise flight, refer to the appropriate ATM.

d. Engine-Out Approach and Landing.

(1)
An engine-out approach and landing is about
the same as a normal
approach and landing. Long, flat approaches with high power output and
ex-
cessive threshold speed on the operating engine result in floating and
unnecessary runway use. These types of approaches should be avoided. Limi-
tations and variations in performance do not allow a specific flight path
or
procedure to be established for all engine-out approaches in Army light
twin-
engine aircraft. However, single-engine approaches in most Army light
twins
can be accomplished using the same flight path and procedures as a
normal
approach and landing. A recommended single-engine landing procedure is
included in the aircraft operator's manual. '

(2) During transition training, the'aviator performs


approaches and
landings with the power of one engine set to simulate the drag of a
feathered
propeller. With the inoperative engine feathered, or set to zero thrust,
normal drag is reduced, which results in a longer landing
roll.
(3) Until the landing is assured, the final approach speed should
not be less than Vyse. Thereafter, the approach speed should be commensurate
with the flap position until the aviator begins the roundout for a
landing.
Under normal conditions, the approach should be made with
full flaps or as
stated in the operator's manual. However, full flaps should not be extended
until the landing is assured. With full flaps, the approach speed should be
Vref or as recommended by the manufacturer. Vref is 1.3 times the stalling
speed (V,so) or the minimum steady flight speed in the landing
configuration.
(4) The aviator should use caution when lowering the
flaps. Once
the flaps have been fully extended on final approach, a go-around should
not
be attempted. Most Army light twins
cannot make a single-engine go-around
with full flaps. For demonstration and practice of engine-out approach and
landing, refer to the appropriate ATM.

NOTE:
a
If rudder trim has been used to counteract aSYmmetric thrust following
power reduction, the aviator should retrim to neutral during the approach.

10-13
When it becomes doubtful that
a safe
e. Single-Engine Go-Around.
landing can be made, the aviator should--
the pitch
Apply maximum controllable power and simultaneously increase
0

to stop the descent.

Retract the landing gear.


0

Retract the flaps to the best lift/drag position.


0

0
Adjust the pitch attitude to avoid loss of altitude.

Accelerate to Vyse.
0

the approach.
Retract flaps after reaching the Vref speed used for
0

0
Trim as required.

10-12. PRESSURIZED AIRCRAFT

less fuel for a given


When an aircraft is flown at high altitude, it consumes
lower altitude. In other
airspeed than it does for the same speed at
a

a higher altitude. In addition, bad


words, the aircraft is more efficient at
is flown in the rela-
weather and turbulence can be avoided if the aircraft
advantages of flying at
tively smooth air above storms. Because of the configured with cabin pres-
higher altitudes, some Army aircraft have been
systems are limited to
surization systems. Aircraft without pressurization
and the operating altitude of
lower altitudes. The degree of pressurization
For example, the fuse-
the aircraft are limited by critical design factors.
lage is designed to withstand
a
particular maximum cabin differential pres-
pressurized aircraft should
sure. Aviators who are being trained to fly those aircraft.
become familiar with the basic operating principles of

A cabin pressurization system


a. Cabin Pressurization System. cabin pressure
safety. It maintains
a
contributes to passenger comfort and
the designed cruising alti-
altitude of no more than 10,000 feet at maximum
cabin pressure
tude of the aircraft. It also prevents rapid changes of
to passengers and crew. In
altitude that may be uncomfortable or injurious and eliminates odors by
addition, the pressurization system removes stale air
Cabin pressurization also
quickly exchanging inside air with outside air.
They can be transported
protects personnel from the effects of hypoxia.
comfortably and safely for long periods in
a pressurized cabin. This is
10,000 feet or
particularly true when the cabin altitude is maintained at the flight crew must
below where oxygen equipment is not required. However,
and be prepared
be aware of the danger of accidental loss of cabin pressure
to meet such an emergency should it occur.

10-14
b. Operating Principles. To understand the operating principles of
pressurization systems, aviators must become familiar with certain terms and
definitions. The following is a list of these pressurization terms and
definitions:
0

Aircraft altitude is the actual height above sea level at which the
aircraft is flying.
0
Ambient temperature is the temperature in the area immediately
surrounding the aircraft.
0
Ambient pressure is the pressure in the area immediately surrounding
the aircraft.
0
Cabin altitude is used to express cabin pressure in terms of
equivalent altitude above sea level.
0

Differential pressure is the difference in pressure between the


pressure acting on one side of a wall and the pressure acting on the
other side of the wall. Differential pressure is the difference
between cabin pressure and atmospheric pressure.

Components.
c. In a typical pressurization system, the cabin, flight
compartment, and baggage compartments are all in a sealed unit that bo1ds air
at pressure higher than outside atmospheric pressure. Cabin superchargers
a

pump pressurized air into the sealed fuselage. These superchargers deliver a
relatively constant volume of air at altitudes up to the manufacturer's de-
signed maximum. A device called an outflow valve releases air from the fuse-
lage. The superchargers provide a constant inflow of air to the pressurized
area; however, the outflow valve is the major controlling element in the
pressurization system because it regulates the releasing of the air.

d. Cabin Pressure Control System. The cabin pressure control system


provides cabin pressure regulation, pressure relief, and vacuum relief.
It
provides a way to select the desired cabin altitude in the isobaric and dif-
ferential ranges. In addition, the cabin pressure control system dumps cabin
pressure. The following paragraphs discuss the different components of the
cabin pressure control system.

(1) Cabin pressure regulator.


The cabin pressure regulator controls
cabin pressure to selected value in the isobaric range. It also limits
a

cabin pressure to a
preset differential value in the differential range.
When the aircraft reaches the altitude at which the difference between the
pressure inside and outside the cabin is equal to the maximum differential
pressure for which the fuselage structure is designed, a further increase in
aircraft altitude results in an increase in cabin altitude. Use of the dif-
ferential control prevents the maximum differential pressure. Differential
pressure is determined by the structural strength of the cabin and the rela-
tionship of the cabin size to the probable areas of rupture, such as windows
and doors.

10-15
Cabin air pressure safety valve. The cabin air pressure safety
(2)
and dump valve. The
valve is a combination pressure relief, vacuum relief,
exceeding a predetermined
pressure relief valve prevents cabin pressure from
differential pressure above ambient pressure. The vacuum relief prevents to
ambient pressure from exceeding cabin pressure by allowing the A outside air
cockpit
enter the cabin when ambient pressure exceeds cabin pressure.
the dump valve. When this switch is properly posi-
control switch activates
solenoid valve opens and dumps cabin air into the atmosph~re.
tioned, a

Cabin differential pressure gage. The cabin differential


(3)
pressure gage is used along with the pressurization controller. It
indicates
be
outside pressure. This gage should
the difference between inside and
monitored to ensure that the cabin does not exceed the maximum allowable
differential pressure.
(4) Cabin altimeter. The cabin altimeter is also used with the
pressurization controller to provide
a check on the performance of the sys-
tem. In some cases, the cabin differential pressure gage and the cabin al-
timeter are combined.
(5) Cabin rate-of-climb or descent gage. This instrument indicates
the cabin rate-of-climb or descent.

Decompression is defined as the inability of the


Decompression.
e.
aircraft pressurization systema to maintain its designed pressure differen- if
tial. This can be caused by malfunction in the pressurization system or
structural damage to the aircraft causes decompression. Physiologically,
They are explosive decompression
decompressions fall into two categories.
and rapid decompression.

when
Explosive decompression. Explosive decompression occurs
(1)
than the lungs can decompress. Lung damage can
cabin pressure changes faster
occur during explosive decompression. Normally, .the time required to release
exist, is 0.2 seconds.
air from the lungs, when no restrictions such as masks explosive
seconds is decompression
Decompression that occurs in less than 0.5
and is potentially dangerous.
(2)
Rapid decompression. During rapid decompression, the lungs can
decompress faster than the cabin. The likelihood of lung damage during rapid
decompression is slight.

f.Hazards of Decompression. Noise may occur during decompression and,


The cabin air
for a split second, the aircrew and passengers may feel dazed.
with fog, dust, and flying debris. The rapid drop in temperature
will fill Ears will pop, but they will
and the change of relative humidity cause fog.
normally clear automatically. However, belching or passing of intestinal gas
may occur. Because the air must escape from the lungs, it rushes from the
mouth and nose. Some major hazards of decompression are discussed in the
following paragraphs.

Hypoxia. The primary danger of decompression is hypoxia. If


(1)
unconsciousness
oxygen equipment is not used quickly, the crew may experience

10-16
in a short time. Rapid decompression shortens the period of the aviator's
useful consciousness. Because decompression rapidly reduces the pressure on
the body and forces oxygen out of the lungs, the partial pressure of oxygen
in the blood is reduced. Therefore, the aviator's effective performance time
is reduced by one-third to one-fourth its normal time. For this reason, the
aviator should wear an oxygen mask on his face when flying at very high
altitudes. If the aircraft is equipped with a demand or pressure-demand
oxygen system, crew members should select the 100 percent oxygen setting on
the oxygen regulator at high altitude.

(2) Other hazards. During decompression, individuals may be tossed


or blown out of the aircraft. Therefore, those that are seated near openings
in the aircraft should wear safety harnesses or seat belts at all times when
the aircraft is pressurized. Other possible hazards are evolved gas decom-
pression sicknesses, exposure to windblast, and extremely cold temperatures.

g. Decompression Warning Systems. "If the above hazards are to be


minimized after decompression, rapid descent from altitude is necessary.
Automatic visual and aural warning systems are included in the equipment of
all pressurized aircraft. These systems will ensure that slow decompressions
do not occur and overwhelm the crew before they can be detected.

10-17
GLOSSARY

Section I

TERMS AND DEFINITIONS

Accelerated stall--a stall caused by increasing an aircraft's weight because


of centrifugal force in a turn or in an abrupt pullout from a dive.

Accelerate-go distance--(See After-lift-off distance.)


Accelerate-stop distance--the distance required to accelerate an airplane to
a
specified speed and, assuming failure of the critical engine at the
instant that this speed is attained, to bring the airplane to a stop.

Acceleration--the rate of change of velocity with respect to time.

Advancing blade--the rotor blade experiencing an increased relative wind


because of airspeed.

Adverse yaw--yaw in the opposite sense to that of the roll of an aircraft;


for example, a yaw to the left with the aircraft rolling to the right.

Aerodynamic center--the point along the chord where all changes in lift are
considered to take place.

Aerodynamics--l: the science that treats the motion of air and other gaseous
fluids and the forces acting on bodies when the bodies move through such
fluids or when such fluids move against or around the bodies. 2a: the
actions and forces resu~ting from the movement or flow of gaseous fluids
against or around bodies. b: the properties of a body or bodies with
respect to these actions or forces. 3: the application of the prin-
ciples of gaseous fluid flows and their actions against and around
bodies to the design and construction of bodies intended to move through
such fluids.

Aerodynamic twist--the twist of an airfoil having different absolute angles


of attack at different spanwise stations.

After-lift-off distance--the total distance required to accelerate the


aircra~t to Vlof and, assuming failure of an engine at the instant that
this speed is attained, to continue takeoff on the remaining engine to a
height of 50 feet. Also known as Accelerate-go distance.

Aileron--a movable control surface or device, one of a pair or set located


in or attached to the wings on both sides of an airplane. Its primary
usefulness is to control the airplane laterally or in roll by creating
unequal or opposing lifting forces on opposite sides of the airplane.
An aileron commonly consists of a flap-like surface at the rear of a
wing, although other devices are sometimes used.

Glossary-l
Aircraft axis--an aircraft six directions of motion about three mutually
has
perpendicular lines (axes). The three axes are the longitudinal axis
about which the airframe rolls, the lateral axis about which the air-
craft pitches, and the vertical axis about which the aircraft yaws.

Airflow--a flow or stream of air. An take place in a wind


airflow may
tunnel, in the induction system of an engine, and so on, or a relative
airflow can occur past the wing or around other parts of a moving
aircraft.
Airfoil--a structure, piece, or body originally likened to a foil or leaf
in being wide and thin and designed to obtain a useful reaction upon
itself in its motion through the air. An airfoil may be no more than a
flat plate, but usually it has a cross section carefully contoured in
accordance with its intended application or function. Airfoils are
applied to aircraft, missiles, or other aerial vehicles or projectiles
to develop lift (as a wing), for stability (as a fin), for control (as
an elevator), and for thrust or propulsion (as a propeller blade).
Certain airfoils combine some of these functions.

Airfoil characteristics--l: any aerodynamic quality peculiar to a particular


airfoil, especially to an airfoil section or profile, usually a
speci-
fied angle of attack. Airfoil characteristics are expressed variously
as the coefficients of lift or drag, the pitching moment, the zero-lift
angle, the lift-drag ratio, and so on. 2: a feature of any particular
airfoil or airfoil section such as the actual or'relative amount of
span, taper, or thickness.

Airfoil section--l: section of an airfoil, especially a cross section,


a

taken at right angles to the span axis or some other specified axis of
the airfoil. 2: the form or shape of an airfoil section; an airfoil
profile or the area defined by the profile.
Airframe--the structural components of an aircraft including the framework
and skin of such parts as the fuselage, wings, empennage, landing gear
(minus tires), and engine mounts.

Airspeed--the speed of an aircraft in relation to the air through which it is


passing.

Airspeed velocity--the component of the relative wind produced by forward


movement of an aircraft.

Angle of attack--the angle at which a body, such as an airfoil or fuselage,


or a system of bodies, such as a helicopter rotor, meets a flow.
Usually expressed as the acute angle between the chord line of an air-
foil and the resultant relative wind.
Angle of climb--the angle between a
horizontal plane and the flight path of a

climbing aircraft.

Glossary-2
Angle of incidence--fixed airfoils (wings, horizontal and vertical fins,
stabilizers): the acute angle between the chord line of the airfoil and
a selected reference plane, usually the longitudinal axis of the
aircraft. Rotating airfoils (helicopter main and tail rotors,
propellers): the acute angle between the chord line of the airfoil and
the tip-path plane. Twisted airfoils: the root chord is commonly
chosen to measure the angle of incidence. Angle of incidence is
normally called pitch angle for main rotor, tail rotor, and propeller
blades.

.Angle of sweep--the acute angle between a reference line in swept or


a

tapered airfoil and some other chosen reference line or plane. For
fixed airfoils, the angle is measured from a plane perpendicular to the
longitudinal axis of an aircraft to the reference line of the airfoil.
The angle is positive if the outboard end of the airfoil reference line
is aft of the inboard end.

Angular acceleration--a simultaneous change in both speed and direction of


movement. An example of this is an airplane in a spin.

Anhedral--a negative dihedral. (See Dihedral.)

Antitorque--a method used to counteract torque reaction.

Articulated rotor system--a rotor system in which the hub is mounted rigidly
to the mast and the individual blades are mounted on hinge pins, allow-
ing them to flap up and down and move forward and backward. Individual
blades are allowed to feather by rotating about the blade grip retainer
bearing.

Attitude--the position of a body as determined by the inclination of the axes


to some frame of reference. If not otherwise specified, this frame of
reference is fixed to the earth (horizon).

Autorotation--the action of turning a rotor system by airflow and not by


engine power. The airflow is produced by movement through the air.
(See Spin.)

Axis--l: a
line passing through a body about which the body rotates or may
be assumed to rotate. Any arbitrary line of reference such as a line
about which the parts of a body or system are symmetrically distributed.
A
line along which a force is directed; for example, an axis of thrust.
2: specifically, anyone of a set or system of mutually perpendicular
reference axes--usually intersecting at the center of gravity of an
aircraft, rocket projectile, or the like--about which the motions,
moments, and forces of roll, pitch, and yaw are measured.

Balancing tab--a tab so linked that when the control surface to which it is
attached is deflected the tab is deflected in an opposite direction,
creating a force which aids in moving the larger surface. Sometimes
called a Servo tab.

Glossary-3
Bank--to roll about the longitudinal axis of the aircraft.
Bernoulli's law--a law or theorem stating that in a flow of incompressible
fluid the sum of the static pressure and dynamic pressure along a
streamline is constant if gravity and frictional effects are
disregarded.

Blowback--the tendency of the rotor disk to tilt aft in forward flight as a

result of flapping.

Boost--additional power, pressure, or force supplied by a


booster; for
example, hydraulic boost.

Boundary layer--a thin layer of air against the surface of the airfoil.
The boundary layer forms the transition from the
airfoil to the air
above. An excessive angle of attack causes the boundary layer to sepa-
rate from the airfoil, and a stall occurs.

Boundary-layer control--the control of the flow in the boundary layer about a


body or of the region of flow near the surface of the body to reduce
or
eliminate undesirable aerodynamic effects and hence to improve
performance.

Buffeting--the beating effect of the disturbed airstream on an aircraft


structure during flight.

Camber--the curvature of the surfaces of an airfoil or airfoil section from


leading edge to trailing edge.

Cantilever--a projecting beam or member supported at only one end; for


example, a wing supported at the fuselage with no external bracing.
Center of gravity--the point within an aircraft through which, for balance
purposes, the total force of gravity is considered to act.

Center of pressure--a point along the chord line of an airfoil through which
all aerodynamic forces are considered to act.
Center-of-pressure travel--the movement of the center of pressure of an
airfoil along the chord with changing angle of attack; the amount of
this movement is expressed in percentages of the chord length from the
leading edge.

Centrifugal force--the apparent force acting on a body moving in a curved


path that is directed away from the center of curvature or axis of rota-
tion. It is not a recognized force in the study of physics.
Centripetal force--the accelerative force acting on a body moving in a curved
path. It is the component of force that is directed toward the center
of curvature or axis of rotation. Centripetal force causes a change in

Glossary-4
of a body in motion, result-
the direction of the linear velocity vector
force is the out-of-
ing in an acceleration of the body. Centripetal
is the horizontal
balance force that causes an aircraft to turn. It of the turn and
the center
component of lift that is directed toward
toward the ,center of the
results from tilting the total lift vector
turn.
blade; the reference line from
Chord--the width of an aircraft wing or rotor measured; a
which the upper and lower contours of an airfoil are edge to the
straight line directly across an airfoil from the leading
trailing edge.

the leading and trailing edges of an


Chord line--a straight line intersecting
airfoil.
the chord, as in chordwise axis
Chordwise--moving, located, or directed along
or chordwise distribution.
number indicating the drag
Coefficient of drag (Cc)--a dimensionless by angle of attack and
inefficiency of an airfoil which is determined
wind tunnel testing.
airfoil design. It is derived from
number indicating the efficiency of
Coefficient of lift (~)--a dimensionless
by angle of attack and airfoil design.
the airfoil whichis determined
It is derived from wind tunnel testing.
mechanical change of the angle
Collective feathering--the simultaneous equal
blades in a rotor system.
of incidence, or pitch, of all rotor
from the advancing blade
Compressibility effects--a phenomenon resulting because of excessive forward
approaching Mach 1, or the speed of sound,
Mach number, a shock wave is
speed. As the blade reaches the critical
density of the air and causes sepa-
formed. This shock wave changes the
the shock wave. The most adverse
ration of the airflow rearward of from the first third of the
effect is a shift of center of pressure moment on the blade.
chord position, causing a severe twisting

high enough that density changes in the


Compressible flow--flow at speeds
fluid can no longer be neglected.

result of lift and centrifugal


Coning--an upward sweep of rotor blades
as a

force.
designed to be rotated or otherwise moved
Control surface--a movable airfoil
to change the speed or direction of an aircraft.
would most adversely affect the
Critical engine--the engine whose failure an
performance or handling qualities of aircraft.
number at
Mach which a local Mach
Critical Mach number--the free-stream
point on the body under consideration.
number of 1.0 is attained at any

Glossary-5
Crosswind--l: the offset of the wind in the
flight of an aircraft that, if
uncorrected, will result in a drift to the
course. When the wind is coming from either
left or right of the intended
like a free balloon; that side, an aircraft behaves
is, it drifts with the wind uriless corrective
action is taken. 2: when used
word means "crosswind leg" which
pertaining to the traffic pattern, the
is normally 90 degrees to the
leg. takeoff

Cyclic feathering--the mechanical change of the


angle of incidence, or pitch,
of individual rotor blades independently
of the other blades in the
system.

Delta wing--a wing shaped in planform


sUbstantially like the Greek capital
letter delta ( ~ ), or like an isosceles triangle, with the base
forming the trailing edge.

Differential ailerons--ailerons geared so that when they


aileron moves through a greater angle than the down are deflected the up
used to reduce adverse yaw aileron. They are
or to lessen the control force necessary
deflection. for

Dihedral--the spanwise inclination of a wing


or other surface, such as a
stabilizer, or of .a part of a wing or other surface to the horizontal or
to a line or plane equivalent to the
horizontal; specifically, a posi-
tive, or upward, inclination.
Dihedral effect--the rolling moment of an
aircraft because of sideslip,
resulting principally from the actual dihedral
The dihedral effect is said of the wing or wings.
to be positive if it tends to raise the
forward wing in the sideslip.

Dissymmetry of lift--the difference in


half and the retreating half of a lift existingA between the advancing
rotor
that is normally compensated for through disk. transitory phenomenon
cyclic feathering and blade
flapping.

Dive--a steep descent with or without power at a


to level flight. greater airspeed than normal

Dive brake--an air brake designed especially to slow down an


dive. A dive brake normally consists of a airplane in a

located in the wing or fuselage. flap, plate, or the like

Double-slotted flap--a flap consisting of two


sections jointed together and
providing two slots for the passage of
air.
Downwash--the induced downward flow of
air resulting from the passage of an
airfoil (induced flow).

Glossary-6
Downwash angle--the angle, measured in plane parallel to the plane of
a

symmetry of an aircraft, between the direction of downwash and the di-


when the
rection of the undisturbed airstream. This angle is positive
deflected stream is downward. (See Upwash angle.)

Drag--the aerodynamic force in a direction opposite that of flight


and caused
by the resistance to movement brought to bear on an aircraft by the
atmosphere through which it passes.

Dynamiç pressure--the pressure of


a
fluid resulting from its motion; it is
equal to one-half the fluid density times the fluid velocity squared
(q 1/2,p V2).
=
In incompressible flow, dynamic pressure is the differ-
ence between total pressure and static pressure.

Dynamic stability--the property that causes a body, such as an aircraft or a


rocket, to dampen the oscillations set up by restoring moments and to
when disturbed from the original
return gradually to its original state
state of steady flight or motion.

Effective translational lift--at approximately 16 to 24 knots airspeed


(depending upon the size, blade area, and rotor RPM of rotor system),
a

the recirculation of old vortexes and the


the rotor completely outruns
short air supply of the hover. At this point, the required air molecule
supply of the rotor system is met and supplied by forward flight. It
now receives a sufficient volume of free undisturbed air to create
a new

aerodynamic environment. Lift improves noticeably.

Elevator--a control surface, usually hinged to


a horizontal
stabilizer,
deflected to impress ~ pitching moment; that is, to make an aircraft or
other flying body of which it is a
part to rotate about its lateral
axis. An elevator may be one of
a
pair, with each one of the pair being
situated to either side of the centerline (hence the frequent use of the
surface running from end
plural "elevators"); or, it may be a continuous
~o end of the stabilizer.

Empennage--the assembly of stabilizing and control surfaces at the tail of an


aircraft.
Endurance--the time an aircraft can continue flying under given conditions
without refueling.
and
Engine cowling--a shroud placed around an aircraft engine for directing
regulating a flow of cooling air, for streamlining, or for protection.

Engine nacelle--a housing for an engine and its accessories.


slung
Engine pod--a streamlined structure or nacelle on an aircraft, usually
b~neath the wings or attached to the wing tips, that houses one or more
jet engines.

Glossary-7
Equivalent airspeed--calibrated airspeed of an aircraft corrected
for
adiabatic compressible flow for the particular altitude.
Equivalent
airspeed is equal to calibrated airspeed in standard atmosphere
at sea
level.

Face curtain--a sheet of heavy fabric, installed above an ejection seat, that
is pulled down
to trigger the ejection seat and to protect the
oxygen mask, and the like against wind blast. face, the

Feathering--a mechanical change in the angle of incidence, or pitch, of an


airfoil segment.

Feedback--the transmittal of forces, which are initiated by


aerodynamic
action on control surfaces or rotor blades, to the cockpit
controls.
Fence--a stationary plate or vane projecting from the upper surface of a wing
(sometimes continuing around the leading edge) and
substantially paral-
lel to the airstream; it is used to prevent spanwise flow. Sometimes
called a stall fence.

Fin--a fixed airfoil that aids directional stability.


Final approach--that portion or left of an approach pattern after the
last
turn in which an aircraft is in line with the runway in the landing
direction and is descending.

Fixed slot--a slot that remains open at all times.


Fixed surface--a stabilizer, fin, or wing, that is either
fixed in position
during flight or adjustable but is distinguished from a movable
surface
or rotary wing.

Flap--a hinged, pivoted, or sliding airfoil or plate, or a combination of


such objects regarded as a single surface;
it is normally located at the
rear of a wing and extended or deflected for increasing camber, espe-
cially at takeoff or during landing.
Flapping--the movement of the rotor blades on an upward (upflap) or a
downward (downflap) path during
rotation.
Flapping velocity--the component of the relative wind produced by blade
flapping.
Flat spin--a spin in which an airplane remains in a more
level attitude than
that of a
normal spin, with centrifugal force holding the
from the axis of the spin.
airplane away

Flight path--the line or way connecting the continuous positions occupied


to be occupied by an aircraft or a missile as or
it moves through the aero-
space with reference to the vertical or horizontal
planes.

Glossary-8
from a surface; the condition of
a flow
Flow separation--the breakway of flow
body and no longer following its
separated from the surface of
a

contours.

Flutter--a vibration or oscillation of a definite period set up in anby the


an
forces and maintained
aileron, a wing, or the like by aerodynamic
and inertial forces of the object
aerodynamic forces and by the elastic
itself.
Fowler flap--a type of extensible trailing-edge
flap that increases both the
camber and wing area.

Frontal area--the projection of


a body, such as a complete aircraft or a

of the body.
nacelle, on a plane perpendicular to the fore-and-aft axis
and indicates the rate of fuel flow
Fuel flowmeter--a flowmeter that measures Also called a fuel-flow
to an aircraft's engines, in pounds per hour.
indicator.
and tail are attached.
Fuselage--the body to which the wings, landing gear,

geometric angles
Geometric twist--the twist of an airfoil having different
of attack at different spanwise stations.
maintained only by the loss
Glide--sustained forward flight in which speed is
of altitude.
and downward path along
Glide angle--the acute angle between the horizontal
which an aircraft descends.
to the vertical
Glide ratio--the ratio of the horizontal distance traveled
distance descended in a
glide. Also called gliding ratio.
following a decision to
Go-around--a procedure for remaining airborne
discontinue an intended landing.
that depends on their
Gravity--an attraction of two objects for each other
mass and the distance between them.

Gross weight--the total weight of an aircraft and its contents.

Ground effect--l:the effect of the ground or surface in turning the


of an air-
downwash vortexes, or induced flow, from the wings or rotor
induced drag and
craft hovering or flying near it, thus reducing at altitudes
increasing lift. 2: the temporary gain in lift very low
because of the compression of the air between
the wings or rotor of an
aircraft and the surface.
phenomenon in rotating systems that makes all forces
Gyroscopic precession--a the direc-
react with a movement 90 degrees from the point of force in
tion of rotation.
Glossary-9
Headwind--a wind blowing from directly ahead
or from a forward direction such
that its principal effect is to reduce ground
speed.
High-lift device--any device, such as a
flap, slat, or boundary-layer control
device, used to increase the CL of a
wing.
Horizontal tail--the entire horizontal part of
extending on both sides of the plane of aircraft's empennage
an
symmetry and, in most forms,
compromising both fixed and movable
surfaces.
Horn--a projection or a
projecting part, as from an aircraft control
surface.
Horsepower--a unit of power equal to the power
necessary to raise 33,000
pounds 1 foot in 1 minute. Thus a 1,000-horsepower
1,000 times 33,000 foot-pounds of work engine develops
per minute.
Hypersonic flow--flow at very high supersonic
speeds; as arbitrarily defined,
flow ata Mach number of 5
or greater.
~

Induced drag--airfoil drag induced by the


production of aerodynamic force.
Induced flow (or Induced
velocity)--flow drawn or sucked in such as the
flow induced by the vortex system of a
rotor or the flow of air or mix-
ture sucked into an engine by the action of the
pistons.

Kinesthesia--the sense which detects and estimates motion


to vision or hearing. without reference

Kinetic energy--the energy of a


system because of motion.

Laminar flow--a smooth flow in which no


cross flow of fluid particles occurs,
hence a flow conceived as made up
of layers.
Laminar-flow airfoil--an airfoil specially designed
to maintain an extensive
laminar-flow boundary layer from a body.

Laminar separation--the separation of a


laminar-flow boundary layer from a
body.

Landing gear--the components of an


on land, water, or other
aircraft that support and provide mobility
surfaces. It may consist of wheels, floats,
skis, or other devices together with
and supporting components. all associated struts, bracings,

Lateral axis--an axis going from side to side of an


It is usually the side-to-side body aircraft,
and so on. rocket missile,
axis passing through

Glossary-IO
the center of gravity. The axis about which pitching action occurs.
Sometimes called a Transverse axis.

the rolling movement of an


Lateral control--control over lateral attitude or
aircraft, rocket, and so on.

oscillation or any
Lateral oscillation--a rolling, yawing, or sideslipping complex lateral
combination of these oscillations; the Dutch roll is a

oscillation.
body, such as an aircraft, to resist
Lateral stability--the tendency of
a

sometimes, lateral displacement; the tendency of an aircraft


rolling or,
to remain wings-level, either in flight or at rest.

Lead and lag--the movement of the rotor blade forward


(lead) and aft (lag) of
through the axis
the radial line from the center of the main rotor shaft
of the drag hinge.
and the like. The edge
Leading edge--the forward edge of an airfoil, blade,
which normally meets the air or fluid first.

Leading-edge flap--a flap installed at the


leading edge of a
wing. It may be
a
split flap, extensible flap, or other kind of flap.

of the total, or resultant, aerodynamic force acting


on
Lift--that component
relative to the body.
a body perpendicular to the undisturbed airflow

Lift oomponent--a force acting on an airfoil perpendicular to the direction


of its motion through the air.

Lift-drag ratio--the ratio of to induced drag, obtained by dividing the


lift
lift by the induced drag or the coefficient of lift by the coefficient
of drag.

Linear acceleration--acceleration along


a
line or axis.

the forces acting on a structure. These may be static (as with


Load--l: combination of
gravity), dynamic (as with centrifugal force), or
a

static and dynamic. 2: used to describe an aircraft's cargo.


Load factor--the of the loads on a structure, including the static and
sum
dynamic loads; expressed in units of G.

member in a fuselage.
Longeron--the principal longitudinal structural
along the longitudinal
Longitudinal acceleration--acceleration substantially
axis of an aircraft,
a
rocket, or the like.
of an
Longitudinal axis--a straight line through the center of gravity
aircraft fore and aft in the plane of symmetry.

Glossary-ll
Longitudinal control--control over the rolling movement of an aircraft, a

rocket, or the like.

Mach number--the ratio of the velocity of a body to that of sound in the


surrounding medium. Thus a Mach number of 1.0 indicates a speed equal
to the speed of sound; 0.5, a speed one-half the speed of sound;
speed five times the speed of sound, and so on.
5.0, a

Mach wave--l: a shock wave theoretically occurring along a common line of


intersection of all the pressure disturbances emanating from an
infinitesimally small particle moving at supersonic speed through a
fluid medium; such a wave is considered to exert no changes in the con-
dition of the fluid passing through it. The concept of the Mach wave is
used in defining and studying the realm of certain disturbances in a
supersonic field of flow. 2: a very weak shock wave appearing, for
example, at the nose of a very sharp body where the fluid undergoes no
substantial change in direction.

Maneuver--any planned motion of an aircraft in the air or on the ground.

Maneuverability--the property of an aircraft that permits it to be maneuvered


easily and to withstand the stresses imposed by maneuvers.

Mass--the amount of material in a body; normally expressed in slugs.

Mass balance--l: weight that brings about a desired balance under static or
dynamic conditions. 2: a
weight or counterpoise used to effect a de-
sired condition of equilibrium, especially about the hinge axis of an
aircraft control surface.
Mean aerodynamic chord--the chord of an imaginary rectangular
airfoil that
would have pitching moments throughout the flight range the same as
those of an actual airfoil or combination of airfoils under considera-
tion, calculated to make equations of aerodynamic forces applicable.
Mean camber line--a line drawn halfway between the upper and lower surfaces
of an airfoil. The curvature of the mean camber line in relation to the
chord line is very important in determining the aerodynamic characteris-
tics of an airfoil section. The maximum camber (displacement of the
mean line from the chord) and the location of the maximum camber help
to
define the shape of the mean camber line. These quantities are ex-
pressed as fractions or a percent of the basic chord length. A typical
low-speed airfoil may have a maximum camber of 4 percent located 40
percent aft of the leading edge. On sYmmetrical airfoils, the mean
camber line and the chord line are the same.

Minimum flying speed--the lowest steady speed at which an aircraft can


maintain altitude under the given conditions and without ground effect.

Minimum gliding angle--the shallowest or flattest gliding angle an aircraft


can steadily maintain with no engine thrust.

Glossary-12
such as that of an aircraft fuselage, in
Monocoque--a type of construction,
which most or all the stresses are carried by the covering or skin.

Movable surface--an aileron, rudder, elevator, flap, or other primary control


surface.

on an aircraft,
Nacelle--a streamlined structure, housing, or compartment
such as the housing for an engine.

by a footward acceleration.
Negative G--the headward inertial force produced when
The force occurs in a gravitational field or during an acceleration
the human body is so positioned that the force of inertia acts on it ~n
a
foot-to-head direction.
a body such that after it is disturbed it
Neutral stability--the stability of from
tends neither to return to its original state nor to move further
neither increase nor decrease
it; that is, its motions or oscillations
in magnitude.

No~ axis--(See Vertical axis.)

Normal spin--a noninverted spin with the nose pointed


steeply downward.

Nosewheel--a turnable or steerable wheel mounted forward


in tricycle-geared
aircraft.

an oblique
Oblique shock wave (or Oblique shock) --a shock wave inclined at
angle to the direction of flow in
a supersonic flow field.

Orientation--the act of fixing position or attitude by visual, artificial, or


other reference.

OVershoot--to fly beyond a designated area or mark.

not contributing
Parasite drag--drag incurred from components of an aircraft
to lift.
of
Pitch (or propeller or rotor blade angle of incidence)--the angle
propeller or rotor blades measured from the plane of rotation.

Pitching mcment--a moment about


a
lateral axis of an aircraft, rocket,
airfoil, and so on.This moment is positive when it tends to increase
the angle of attack or to nose the body upward.
and nonsplit)
Plain flap--the elemental flap {nonextensible, nonslotted,
hinged to a wing and forming part of the trailing edge.

Glossary-13
Plenum chamber--a chamber in certain ducting systems, such as that
in a

gas-turbine engine, that receives ram air for the compressor.

Positive G--the footward inertial force produced by a headward


The force occurs in a gravitational
acceleration.
field or during an acceleration when
the human body is so positioned that the force of
inertia acts on it in
a
head-to-foot direction.

Positive lift--lift acting in an upward direction.


Potential energy--the energy of a
system derived from position.

Power--the rate of doing work; often expressed in units of horsepower.

Pressure gradient--a change in the pressure of a


gas or fluid per unit of
distance.

Profile drag--drag incurred from frictional or parasitic resistance of the


blades passing through the air. It does not change significantly with
the angle of attack of the airfoil section,
as airspeed increases.
but it increases moderately

Propeller--a device for producing thrust in a


fluid such as water or air.
Pylon--l: prominent mark or point on the ground used as a fix in precision
a

maneuvers. ~: a
rigid structure on the outside of an aircraft for
supporting something such as an engine, tank, or rocket.

Rate of climb--the rate at which an aircraft gains altitude; that


is, the
vertical component of its airspeed in climbing.
Rate of descent--the rate at which an aircraft descends; that
is, the
vertical component of its airspeed in descending; the rate at which a

parachute and its burden descend.

Relative wind--air in motion with respect to an airfoil.

Resultant aerodynamic force--the resultant of all the forces acting on an


airfoil segment as a result of relative airflow.
Resultant relative wind--airflow from rotation that is modified by induced
flow.

Retreating blade--the rotor blade experiencing a


decreased relative wind
because of airspeed.

Retreating blade stall--a stall that begins at or near the tip of a blade
because of the high angles of attack required to compensate for
dissym-
metry of lift.

Glossary-14
Reverse thrust--thrust applied to
a
moving object in a
direction opposite the
direction of the object's motion.

Reynolds number--the product of


a
typical length and the fluid speed divided
by the kinematic viscosity of the fluid. It expresses the ratio of the
internal forces to the viscous forces.
Rigid rotor system--a rotor system in which the rotor blades are fixed
The
rigidly to the hub and are not allowed to flap or lead and lag.
only action allowed is pitch change.

Roll--movement around the longitudinal axis of an aircraft.

Rotational velocity--the component of the relative wind produced by rotation


of the rotor blades.

Rotor--a system of rotating airfoils.

Rotor disk--(See Tip-path plane.)

Roundout--a change of aircraft attitude and flight path from that used on
final approach to that used for landing.

Rudder--an upright control surface that is deflected to impress


a yawing
moment; that is, to make an aircraft or other body of which it is a part
rotate about its vertical axis.

Rudder peda1s--controls within an aircraft by means of which the rudder is


actuated.

Semimonocoque--a type of construction, such as that of


a fuselage or nacelle,
in which longitudinal members, as well as formers, reinforce the skin
and help carry the stresses.

Semirigid rotor system--a rotor system in which the blades are connected to
the mast by a trunnion that allows blades to flap. Pitch change (feath-
ering) is allowed at the hub about the blade grip retainer bearing.

Separated flow--flow over or about


a body that has broken away from the
surface of the body and no longer follows its contours.

Service ceiling (single-engine service ceiling)--the maximum altitude


an
of obtaining with one engine inoperative under stan-
aircraft is capable
dard conditions.

Servo tab- (See Balancing tab.)


-

Settling with power--a transient condition of downward flight


(descending
through air after just previously being accelerated downward by the
rotor) during which an appreciable portion of the main rotor system is
being forced to operate at angles of attack above maximum. Blade stall

Glossary-1S
starts in near the hub and progresses outward as the rate of descent
increases.

Shock--(See Shock wave.)

Shock stall--a sudden reduction of lift on an airfoil brought about at


supersonic speeds by airflow separation aft of a shock wave.

Shock wave--a surface or sheet of discontinuity (abrupt changes in


conditions) sets up a supersonic field of flow through which the fluid
undergoes a finite decrease in velocity accompanied by a marked increase
in pressure, density, and temperature, as occurs in a supersonic flow
about a body. Sometimes called a Shock.

Sideslip--a movement of an aircraft such that the relative wind has a

velocity component along the lateral axis.


Sinking speed--the rate at which an aircraft loses altitude; especially, the
rate at which heavier-than-air aircraft descends in
a a
glide in still
air under given conditions of equilibrium.
Skid--rate of turn is greater than normal for degree of bank established.
Skin friction--the friction of a fluid against the skin of an aircraft or
other body; friction drag.

Slat--any of certain long and narrow vanes or auxiliary airfoils; for


example, a
venetian blind flap; specifically, the vane used in an auto-
matic slot.

Slip--the rate of turn is less than normal for the degree of bank
established.

Slipstream--the current of air driven astern by the propeller.

Slot--a long and narrow opening, such as that between a wing and a deflected
fowler flap; specifically, a long and narrow spanwise passage in a wing,
usually located near the leading edge, that improves flow conditions at
high angles of attack.

Slotted aileron--an aileron having a specially contoured leading edge which,


along with the wing, forms a slot for the smooth passage of air over the
aileron upper surfaces when the aileron trailing edge is deflected
downward.

Slotted flap--a flap that exposes a slot between itself and the wing when
deflected, or a flap consisting of a number of slim surfaces, or slats,
fastened together with slots between them.

Slug--a unit of mass that is accelerated at the rate of 1 foot per second
per second when acted upon by a force of 1 pound. To convert weight (in

Glossary-16
(in pounds) by the gravitational
pounds) to slugs, divide the weight
constant, 32.2 feet per second.
the large increase in drag that acts upon
Sonic barrier--a popular term for of sound. Also called the Sound
an aircraft approaching the speed
barrier.
which is
sound heard when a shock wave,
Sonic boam--an explosion-like
speed, reaches the ear.
generated by an aircraft flying at supersonic
The principal shock waves are approximately
conical in shape and origi-
or object. The shock wave
nate at the front and rear of the aircraft
cone angle depends upon the speed
of the aircraft and the speed of
shock
To the observer, who senses the
sound in the surrounding medium. manifested as a
with his
wave the arrival of each pressure wave is
ear,
booming sound.

the speed of sound.


Sonic speed--pertaining to sound or

Sound barrier--(See Sonic barrier.)


from tip to tip, or
Span--~: the dimension of an airfoil from end to end,
an aircraft, measured between
from root to tip. b: the dimension of
of an airfoil from tip to tip,
lateral extremities. 2: the dimension or elevators extend beyond
measured in a straight line. Where ailerons
is included in the span.
the tips of the airfoil proper, their extension the span.
Sweeping an or giving it dihedral decreases
airfoil
moves in relation to time and distance.
Speed--the rate at which an object
which sound travels in a given medium under
speed of sound--the speed at
specified conditions.
of an airplane, controlled or uncontrolled,
Spin--a maneuver or performance helical (corkscrew) path while fly-
in which the airplane descends in
a

the angle of maximum lift. The


ing at an angle of attack greater than
nose of an airplane in
a spin is usually, though not necessarily,
pointed sharply downward.
which an aircraft ascends or descends in
Spiral--a maneuver or performance in an angle of attack that is within the
a
helical path while flying at
The flight path of an aircraft so as-
normal range of flight angles.
cending or descending.
the bottom of a wing, usually near
Split flap--a plate or surface hinged to increased camber and drag.
the trailing edge, and deflected downward for

comb, tube, bar, or other device that


Spoiler--a plate, series of plates, body break up or spoil the
to
projects into the airstream about
a

such a device that projects from the


smoothness of the flow, especially decreased
It gives an increased drag and
a

upper surface of an airfoil.


lift.

Glossary-17
Spring tab--a tab attached to a
control surface and actuated through a
control linkage that is spring-loaded
in such a manner that the tab
supplies a certain amount of the
force necessary to move its
surface. Its action and use are similar control
to those of the servo
However, the springs incorporated
make the tab supply
in the control system work so tab.
as to
only a part, rather than
to move its surface. all, of the force required
Stabilator--a horizontal surface that pivots as a
the usual combination of whole; it is distinct from
fixed and movable surfaces.
Stability--the property of an aircraft to maintain
displacement and, if displaced, to its attitude or to resist
develop forces and moments tending
restore the original condition. to

Stabilizer--a fixed or adjustable airfoil or vane


an aircraft; that that provides stability for
is, a fin or more specifically the horizontal
lizer on an aircraft. stabi-

Stall--la: a condition in which a wing or other


flies at an angle of attack greater than thatdynamically lifting body
ing in a loss of lift and an for maximum lift, result-
increase ofdrag. b:
an increase of drag brought on
by a shock wave; that
a
loss of and lift
2: the flight condition or behavior is, a shock stall.
of an aircraft flying at an angle
greater than the angle of maximum
mances involving a lift;
any of various aircraft
perfor-
stall.
Stall fence--(See Fence).
Stalling angle of attack--l: the minimum angle of attack
airfoil section or other dynamically of an airfoil or
curs; that lifting body at which a
stall oc-
is, a
critical angle of attack. 2: the angle of maximum
lift.
Stall speed--the airspeed at which, under a
given set of conditions, an
aircraft will stall.
Static pressure--the atmospheric
pressure of the air through which an
aircraft is flying.
Steady-state--the condition of a substance
chemical properties do not
or system whose local physical and
vary with time.
Streamline--the path of a
particle or small portion of
noneddying fluid, commonly taken
a
fluid, usually a
with respect to a body in
path whose tangent at any
point is in the direction of the
it; such a
vector at that point. velocity

Subsonic flow--flow at a
velocity less than the speed of sound in the
under the prevailing conditions. medium

Glossary-18
Supersonic flow--flow at a speed greater than the speed of sound in
the
medium under the prevailing conditions.

SWeepback--the backward slant from root to tip (or inboard end to outboard
end) of an airfoil or of the leading edge or other reference line of an
airfoil. Sweepback usually refers to a design
a in which both the leading
and trailing edges of the airfoil have backward slant.

Tab--a small auxiliary airfoil set into the trailing edge of an aircraft
control surface (or something set into or attached to another surface
such as a rotor blade) and used for trim or to move or assist in moving
the larger surface.

Tail rotor--the antitorque device of a


single-rotor helicopter. Control of
this rotor is through the foot pedals.

Tail rotor drift-- (See Translating tendency.)


Tailwheel--a turnable or steerable wheel mounted at the aft end of the
airframe.

Tandem rotor system--a main lifting rotor is used at each end of the
helicopter. The rotor systems rotate in opposite directions to counter-
act torque.

Taxi--l: the operation of an airplane or helicopter under its own power on


the ground, except that movement incident to actual takeoff and landing.
~: the forward movement of a helicopter at a hover is referred to as
a

hover taxi.

Terminal velocity--the hypothetical maximum speed that could be obtained in


a

prolonged vertical dive.

Thrust--l: the resultant force in the direction of motion caused by the


components of the pressure forces in excess of ambient atmospheric pres-
sure acting on all inner surfaces of the vehicle propulsion system
parallel to the direction of motion. 2: the force that opposes drag.
In rotary-wing aircraft, it is often expressed as the total lift of the
rotor system.

Thrust axis--a line or axis through an aircraft, rocket, and so on alonga


which the thrust acts; an axis through the longitudinal center of jet
or rocket engine along which the thrust of the engine acts; a
center of
thrust.

Thrust borsepower--the actual amount of horsepower that an engine-propeller


combination transforms into thrust. It is brake horsepower multiplied
by propeller efficiency.

Glossary-19
Tip-path plane--a plane defined by the circle scribed by the
average flight
path of the blade tips in a rotor system.
Rotor disk.
It is sometimes called the

Tip vortex--a vortex springing from the tip of a wing because of the flow of
air around the tip from the high-pressure region below the surface to
the low-pressure region above
it.
Torque--any turning or twisting force applied to the rolling force that
imposed on an aircraft by the engine in turning the
is
propeller.
Torque effect--the reaction to the turning of the rotor system.
system turns counterclockwise, the fuselage reacts by If
the rotor
turning clockwise.
Trailing edge--the rearmost edge of an airfoil.
Trailing-edge flap--a flap, especially a
plain flap, installed at the rear of
a
wing.

Trailing vortex--a vortex that is shed from a wing or other lifting body and
is trailing behind it, especially such a
vortex trailing from a wing tip
or from the end of a bound vortex. It is sometimes referred to as Wake
turbulence.

Translating tendency--the tendency of the single-rotor helicopter to move


laterally during hovering flight. Also called Tail rotor drift.
Translational flight--any horizontal movement of a
helicopter with respect to
the air.

Translational lift--additional lift obtained because of translation.

Transonic speed--the speed of a body relative to the surrounding fluid


which the flow is in some places subsonic and
at
in other places
supersonic.

Transverse axis--(See Lateral axis.)

Transverse-flow effect--a condition of increased drag and decreased


the aft portion of the rotor disk caused by the lift in
air having a greater
induced velocity and angle in the aft portion of the
disk.
Trim--the condition of a heavier-than-air aircraft in which
it maintains a
fixed attitude with respect to the wind axes, with the moments about the
aircraft axes being in equilibrium. The word "trim" is often used with
special reference to the balance of control forces.

Trim speed--a speed at which an aircraft maintains a


given trim.
Trim tab--a tab that is deflected to a position where
it remains to keep the
aircraft in the desired trim. Adjustment of a trim tab on a rotor blade
causes the blade to maintain a given track or plane of
motion.

Glossary-20
Tropopause--the upper boundary of the troposphere.
characterized
Troposphere--the lowest layer of the earth's atmosphere,
especially by a relatively steady temperature lapse rate, varying humid-
ity, and turbulence.
True airspeed--equivalent airspeed corrected for error
that is due to air
density (altitude and temperature).

Turbulence--an agitated condition of the air or other


fluids; a disordered,
fluid or fluid flow such as that about
a

irregular, mixing motion of


a

body in motion through the air.

by random
Turbulent boundary layer--a boundary layer characterized
fluctuations of velocity and by pronounced lateral mixing of the fluid.

Turbulent flow--a flow characterized by turbulence;


that is, an irregular,
eddying, fluctuating flow; a flow in which the velocity of a given point
varies erratically in magnitude and direction with time.

destructive
Ultimate load--a load that causes, or is calculated to cause,
failure of a structural member or part.
parallel and
UnifoDn flow--an idealized flow in which the streamlines are
the velocity is constant throughout.

Unsteady flow--a flow whose velocity components vary


with time at any point
the velocity at any
in the fluid. Unsteady flow is of fixed pattern if
poïnt changes in magnitude but not direction and of variable pattern if
the velocity at any point changes in direction.

Upwash--a flow deflected upward by


a
wing, rotor, rotor blade, and so on.

Upwash angle--a negative angle; that is, the acute angle, measured
downwash
between the
in a plane parallel to the plane of symmetry of an aircraft,
direction of upwash and the direction of the undisturbed airstream.
weight and maximum
Useful load--the difference, in pounds, between the empty
authorized gross weight of an aircraft.

Also a graphic
Vector--a quantity having both magnitude and direction.
illustration of such a quantity.
that includes both magnitude
Velocity--.!.: speed. I: a vector quantity
a given frame of reference. 3: time
(speed) and direction relative to
rate of motion in a given direction.
that increases the fluid
Venturi--a converging-diverging passage for fluid
a venturi tube.
velocity and lowers its static pressure;
Glossary-21
Vertical axis--an axis passing through an aircraft from top to bottom and
usually passing through the center of gravity. The axis about which yaw
occurs. Also called a Normal axis.

Vertical fin--(See Vertical stabilizer.)

Vertical stabilizer--a vertical fin mounted approximately parallel to the


longitudinal axis of an aircraft to which a rudder may be attached. The
vertical stabilizer aids in directional stability. Also called a Verti-
cal fin.

Vertical tail--a vertical, or substantially vertical, component of an


aircraft's stabilizing and controlling surfaces in most forms, compris-
ing both a fixed surface (vertical stabilizer) and a movable
( rudder) surface
.

Wake turbulence--(See Trailing vortex.)

Warp--to twist or to give twist to an airfoil, especially a wing. On certain


early aircraft, the wings were warped at will to perform the function
now performed by ailerons.

Wash--the air which has been disturbed by the passage of an


airfoil.
Washin--a wing twist design that has a
greater angle of incidence at the wing
tip than the wing root.
Washout--a permanent warp or twist given a wing such that some specified
or
understood angle of attack, usually the geometric angle of
attack, is
smaller at the tip than at the root.

Weathervane--the tendency of an aircraft on the ground to face into the wind.


Weight--a measure of the mass of an object unqer the acceleration of
gravity.
Wing--an airfoil thatprovides or is designed to provide for an lift
aircraft by extending on either side of the
aircraft and by being sepa-
rated from its mate by the fuselage or hull. (A monoplane so con-
structed is said to have wings.)
Wing area--the area within the outline of the projection of a wing on the
plane of its chords, including that area lying within the fuselage,
hull, or nacelles. (The wing in this case is considered as a unit,
extending on both sides of an aircraft.)

Wing drag--that part of the total drag on an aircraft arising from the
reaction of the air with its wing or wings, including profile drag and
the drag that is due to
lift. Where appropriate, the parasite drag of
components attached to the wing is also included.

Glossary-22
Wing root--the end of the wing that joins the fuselage or opposite wing.

Wing tip--the end of the wing farthest from the fuselage or cabin.

Work--a force exerted over a given distance.

Yaw--movement about the vertical axis.

other body
Zero angle of attack--the position of an airfoil, fuselage, or
when no angle of attack exists between two specified or understood
reference lines.
which no lift is
Zero-lift angle of attack--the geometric angle of attack at
created. Often called the angle of zero lift or the zero-lift angle.

climb--a brief, steep climb in or of an aircraft, with its


momentum
Zoom
being expended in the climb.

Section II
ABBREVIATIONS AND BREVITY CODES

AB advancing blade

AC aerodynamic center

AF angle of force

AH attack helicopter

A.1. induced .angle of attack

Ao section angle of attack


AGL above ground level

A/s airspeed

ATM aircrew training manual

avg average

BHP brake horsepower

BLC boundary-layer control

Glossary-23
BMEP
brake mean effective pressure

BSFC brake specific fuel consumption

BTS
blade-tip speed

C
Celsius
CA
circle of motion
CAT
carburetor air temperature
CD
coefficient of drag
CF
centrifugal force
CF coefficient of force
CG
center of gravity
CH
cargo helicopter

CI coefficient of the rolling moment

CL coefficient of lift
CM coefficient of the pitching moment

CN coefficient of the yawing moment

cos cosine
CP
center of pressure

D
drag

Di induced drag

Do profile drag

Dp parasite drag
Dt total drag

Glossary-24
ESHP equivalent shaft horsepower

ETL effective translational lift

f equivalent parasite area


F
force, Fahrenheit
FAA Federal Aviation Administration

FAR Federal Aviation Regulation

FF fuel flow

FLIP flight information publication


FM frequency modulated, field manual

FPM feet per minute


FPV flight-path velocity
FSRW
free-stream relative wind

ft foot

ft-lb foot-pound

G gravity

h height of a column of air


H
total pressure
HP horsepower

HPA horsepower available

HPR horsepower required

hr hour

Glossary-25
IAS indicated airspeed
ICAO
International Civil Aviation Organization
IF induced flow

IGE in-ground effect

IMC instrument meteorological conditions

KIAS knots indicated airspeed

kt knot

L
lift
lb pound

L/D lift/drag
LZ landing zone

MAC moment about the aerodynamic center


MAP manifold absolute pressure

max maximum

min minimum

mph mile(s) per hour

n
load factor

N
north
NOE
nap-of-the-earth
nW load factor times weight

Glossary-26
OGE out-of-ground effect
OR observation helicopter

p static pressure
P pressure

PR power required

q dynamic pressure

R resultant
RB retreating blade
RPM revolution(s) per minute
RW
relative wind

s surface area

sec second

sin sine
STANAG standardization agreement

T thrust

TA thrust available
TAJ total aerodynamic force

TAS true airspeed


Td total drag

TD touchdown

TH horizontal component of thrust

Glossary-27
TR thrust required
TV vertical component of thrust

DR utility helicopter

v
velocity

vel velocity
VH horizontal velocity

Vmax maximum level velocity


Vv vertical velocity

w
downwash vector, average specific weight of air
w weight

x times sign or multiplication sign

Section III
SYMBOLS

Ii angle of attack

T angle of climb, or angle of descent

ß beta
0

degree

negative (minus)
Ïo
percent
+
positive (plus)
p rho (air density)
Glossary-28
Section IV

V-SPEED ABBREVIATIONS

Va design maneuvering speed. The maximum speed for which the


aircraft is designed for full abrupt control deflection without
incurring structural damage.
Vb design speed for maximum gust intensity (turbulence penetration
speed).

Vc design cruising speed.

Vd design diving speed.

Vdf/Mcif demonstrated flight diving speed.

Vf design flap speed.

Vfc/Mfc maximum speed for stability characteristics.


Vfe maximum flap extended speed.
Vh maximum speed in level flight with maximum continuous power.

VIe maximum landing gear extended speed.

Vlo maximum landing gear extension/retraction speed.

Vlof lift-off speed (normally rotation speed +3 knots).


Vmc minimum control speed with the critical engine inoperative.
Vmo/Muto maximum operating limit speed.

Vmu minimum unstick speed. The minimum speed at which the aircraft
can be made to lift
off the ground without demonstrating
hazardous characteristics and be able to continue the takeoff.

Vne never exceed speed.

Vno maximum structural cruising speed.

Vr rotation speed.

Vref The indicated airspeed the aircraft should be at 50 feet above


the landing area in the landing configuration. It is the "land-
ing approach speed" shown in most aircraft operator's manuals.
If the operator's manual does not give Vref, 50-foot speed, or
landing approach speed, then 1.3 power-off stall speed in landing
configuration (Vso) is used as the Vref speed. Vref-plus speeds

Glossary-29
are used during visual and instrument approach procedures and
provide a simple method of computing the various speeds required
for traffic patterns and instrument approaches. The. following
are examples of Vref use:

Vref =
Landing-approach speed (desired speed at 50
feet above the landing area).
+ 10 =
Final-approach speed.
Vref

Vref + 20 =
Base-leg speed or instrument-approach speed.
+ 30 =
Speed after landing gear has been lowered.
Vref

Vref + 40 =
Pattern speed or instrument-approach speed
until lowering landing gear.
Vs power-off stalling speed. The stalling speed or the minimum
steady flight speed at which the aircraft is controllable.

Vso the calibrated power-off stalling speed or the minimum steady


flight speed at which an aircraft is controllable in the landing
configuration.

VsI the calibrated power-off stalling speed or the minimum steady


flight at which an aircraft is controllable in
speed a
specified
configuration.

Vsse the safe twin-engine operative speed selected to provide a rea-


sonable margin against the occurrence of an unintentional stall
when making intentional engine cuts. In flight, simulated engine
failures below this speed are prohibited. If Vsse is not listed
in the aircraft operator's manual, simulated engine failures at
or below Va. will not be performed.

Vx the calibrated airspeed at which an aircraft will obtain the


highest altitude in a given horizontal distance. This best-
angle-of-climb speed normally increases slightly with altitude.

Vxse speed for the best single-engine angle of climb.


Vy the calibrated airspeed at which an aircraft will obtain the
maximum increase in altitude per unit of time. This best-rate-
of-climb speed normally decreases slightly with altitude.

Vyse speed for the best single-engine rate of climb.

VI takeoff decision speed (formally denoted as critical engine fail-


ure speed).

Glossary-30
V2
takeoff safety speed.

V2min minimum takeoff safety speed.

SECTION V

AERODYNAMIC EQUATIONS

Chapter 1 Equation

static pressure p =
wh 1.1

dynamic pressure q =
1/2 P V2 1.2

Bernoulli's equation H =
p + 1/2 P V 2; or H =
p + q 1.3
(total pressure)

Chapter 2

free stream pressure (See Equation 1.3)


(total pressure)

aerodynamic force F =
CF 1/2 P S V2; or F =
Cf q S 2.1

F
coefficient of force CF V2; or CF -..1...- 2.2
= -

1/2 P S q S

L 1/2 V2; or
lift =
CL P S L =
CL q S 2.3

drag D =
CD 1/2 P S V 2; or D =
CD q S 2.4

L
coefficient of lift CL V2; or CL =-1... 2.5
=

1/2 P S q S

D D
coefficient of drag CD =

1/2 P S V2
; or CD =

q S
2.6

CL L CL 1/2 P S V2
lift/drag ratio -1... 2.7
or D
- . =
-

D CD
'
CD 1/2 P S V2

Glossary-31
Dp 1/2pf V2
parasite drag =
2.8

power Power
Work (in ft-lb)
-

Time (in see) 2.9

(in ft-lb)
horsepower Power (in see)
Horsepower =

550 2.10

power required PR = DV
2.11

DV (V in ft/sec)
horsepower required HPR =

550
-
-
TR -Y-
550

(or) 2.12

HPR -
-
-
DV (V in knots)
325

HPA TV
horsepower available =

325 2.13

horsepower for maximum HPR


max
=

~ -
CL 3/2) 2.14
endurance CD max

maximum range (FF). (HPR)'


~ mln ~ mln
=

2.15

Chapter 4

brake specific fuel BSFC engine fuel flow C lb/hr


=

BHP
; or =

BHP 4.1
consumption
(BSFC C)
=

Chapter 6

blade tip speed BTS (in ft/sec) = rotor radius (in it) x RPM
6.1
9.55

Chapter 8

To compute the development of stall speed, see Equations 8.1 through 8.5.

Glossary-32
stall speed L =
Constant 1/2 p V2 S CL
8.1
-

lift equation in flight L = W =


1/2 p V
2 S CL
8.2

solve for V
V-v~ -

CL P S 8.3

substitute W
for
(then)
L
V-V~ -

CL P S
8.4

power-off stall speed V


s
=
VC~x p S 8.5

load factor n =
L/W
8.6

lift L = nW
8.7

stall with
speed
acceleration
V
S
-
-

V 2nW

CLmax P S
8.8

vertical summing forces L + T


sin Q -
nW = 0
8.9

vertical summing forces L = nW -


T s
in a
8.10
continued

stall speed with thrust Vs =V2(nW -


T
P S
sin a )
8.11
CLmax

lift equation in L = W cos T 8.12


climbing flight

T D + W
=
sin T

Glossary-33
REFERENCES

RELATED PUBLICATIONS

Related publications are sources of additional information. The user


does not have to read them to understand this publication.

ARMY REGULATIONS

95 Series (aviation regulations)

310-25 Dictionary of United States Army Terms

310-50 Authorized Abbreviations, Brevity Codes, and


Acronyms

DEPARTMENT OF THE ARMY PAMPHLET

25-30 Consolidated Index of Army Publications and Blank


Forms

FIELD MANUAL

1-240 Instrument Flying and Navigation for Army Aviators

TECHNICAL MANUALS

55-1510-XXX-10 Series (operator's manual for the appropriate helicopter)

55-1520-XXX-10 Series (operator's manual for the appropriate airplane)

TRAINING CIRCULARS

1-209 Aircrew Training Manual, OH-58D Observation Heli-


copter, Aviator/Aeroscout Observer

1-210 Aircrew Training Program, Commander's Guide

1-216 Aircrew Training Manual, Cargo Helicopter

DA FORM

2028 Recommended Changes to Publications and Blank


Forms

References-l
COMMAND RELATED PUBLICATIONS

Command publications cannot be obtained through Armywide


resupply channels. Determine availability by contacting
the address shown. Field circulars expire three years
from the date of publication unless sooner rescinded.

FIELD CIRCULARS

1-214 Aircrew Training Manual, Attack Helicopter, AH-64.


May 1986. Commander, US Army Aviation Center,
ATTN: ATZQ-ATB-O, Fort Rucker, AL 36362-5218.

1-215-1 Aircrew Training Manual, Aeroscout Observer.


March 1986. Commander, US Army Aviation Center,
ATTN: ATZQ-ATB-O, Fort Rucker, AL 36362-5218.

1-222 Rotary-Wing Instructor Pilot Handbook. October


1985. Commander, US Army Aviation Center, ATTN:
ATZQ-ATB-O, Fort Rucker, AL 36362-5218.

OTHER PUBLICATIONS

FEDERAL AVIATION ADMINISTRATION

Federal Aviation Regulation

NOTE: To obtain the FAR, see your unit FLIP manager.

PROJECTED RELATED PUBLICATIONS

Projected publications are sources of additional information


that are scheduled for printing but are not yet available.

TRAINING CIRCULARS

1-211 Aircrew Training Manual, Utility Helicopter, UH-1


(projected date: August 1988)

1-212 Aircrew Training Manual, Utility Helicopter, UH-60


(projected date: September 1988)

1-213 Aircrew Training Manual, Attack Helicopter, AH-1

References-2
1-215 Aircrew Training Manual, Observation Helicopter,
OH-58

1-216-1 Aircrew Training Manual Supplement, Flight


Engineer

1-217 Aircrew Training Manual, Surveillance Airplane,


OV-1

1-218 Aircrew Training Manual, Utility Airplane


1-221 Fixed-Wing Instructor Pilot Handbook

References-3
INDEX

This index is organized alphabetically by topics and subtopics.


Topics and subtopics are identified by page numbers.

Acceleration, 1-1, 6-79, 6-80, 8-19


Action-reaction, 1-1
Adverse yaw, 3-20, 3-21, 8-49
Aerodynamic
balance, 8-52 through 8-54
center, 2-4, 2-20
force, 2-13
limit of performance, 6-83, 8-35
pitching moment, 2-20, 2-21
properties, 2-4, 2-5
stall, 3-15, 8-12 through 8-14
Aerodynamics, 1-1
Aeroelastic effects, 4-5
Ailerons, 3-21, 8-45, 9-2
Aircraft
classification, 8-1
control surfaces, 8-3
design, 4-1, 4-4
fuselage, 8-2
landing gear, 8-4
light-twin, 10-1
single-engine, 10-1
structure, 4-1, 4-4, ~-2
wings, 8-3
Airflow
Bernoulli's principle of, 2-7, 2-9
development on rotor system, 6-4
during hovering, 6-23
incompressible, 1-2
in forward flight, 6-10
patterns in forward flight, 6-28
Airfoil
airflow, 2-7
cambered, 2-1, 2-2, 2-18, 2-20
nonsymmetrical. See Cambered airfoils
positive-cambered, 3-9
section, 2-3
section theory, 2-7
shape, 2-3, 2-4
symmetrical, 2-1, 2-2, 2-18, 2-20, 8-45
terminology, 2-3
Airspe.ed
indicator, 1-6
maximum-range, 6-77 through 6-79
measurement, 1-1

Index-1
Air work
climbs, 9-8
descents, 9-9
glides, 9-9
straight and level flight, 9-7
turns, 9-11 through 9-13
Angle
of attack, 2-4, 2-19, 2-20, 2-28, 2-37, 8-25, 8-26
of climb, 6-75 through 6-77, 8-23 through 8-25
of incidence, 2-4
Anhedral angle, 3-18, 3-19
Antitorque
pedals, 5-22, 7-6, 7-7
rotor, 5-1, 5-22
Approach, 7-21, 9-22, 9-23, 10-8, 10-13
confined area, 7-21
engine out, 10-13
Asymmetric
control, 10-4
thrust, 8-50, 10-2 through 10-5, 10-11, 10-13
Attitude
control and airspeed, 7-5
control and coordinated turns, 7-3
flying, 7-2, 7-3, 9-14
pitch, 7-6
rotation, 6-68
shift, 5-33
takeoff, 8-48
Autorotation, 9-19, 6-53 through 6-70

Balance, aerodynamic, 8-52 through 8-54


board, 8-53
tabs, 8-54
Balance of forces, 6-1, 6-2
Bank angle, 10-4, 10-5
Banking, 10-4 through 10-7, 8-32, 8-33
Banks
with turns, 9-13
without turns, 9-14
Barometer, aneroid, 1-3, 1-5
Bernoulli
equation, 1-5, 1-6
principle, 1-1, 1-6, 2-9
Best-angle-of-climb speed
Best-rate-of-climb speed
Blade
drag, 5-18
element theory, 2-7, 2-8
flapping, 6-14 through 6-16
grip feather bearing, 5-3
lead and lag, 5-8, 5-18 through 5-20
pitch, 6-19
Index-2
tip speed, 6-6
twist, 5-3
Blowback, 6-17 through 6-19, 6-30
Boundary layer, 2-9, 2-31, 6-36
control, 8-6, 8-7
separation, 8-5, 8-13, 8-14
Bound vortex, 2-34
Bow wave, 6-36
Brake
horsepower, 4-18
mean effective pressure, 4-19
specific fuel consumption, 4-23
Buffet, 4-5
Cabin pressurization system
components of, 10-15
operating principles of, 10-14
Cambered airfoils, 2-1, 2-2, 2-18, 2-20
Cathedral angle, 3-18
Center-of-gravity limits, 5-31, 5-32
Center of pressure, 2-4
Centrifugal force, 2-6, 2-7, 5-9, 5-10, 5-25, 7-3, 7-4, 8-32, 8-33
Centripetal force, 2-6, 2-7, 5-9, 6-80, 6-82, 8-33
Chord, 2-3, 5-2
Chord line, 2-3
Circular motion, 5-5
Climb, 10-2, 10-7
angle of, 8-23 through 8-25
rate of, 8-23, 8-26
steady-state, 8-23
Climbing
flight, 8-22, 8-24
performance, 8-22, 8-23
stall speed, 8-26
Coefficient
of drag, 2-24
of force, 2-13
of lift, 3-6, 3-7, 3-9, 8-4
of the pitching moment, 3-5 through 3-7
of the rolling moment, 3-17
of the yawing moment, 3-13, 3-16
Collective
blade angles, 7-6
control, 7-13
inputs, 7-14
Combustion chamber, 4-11
Compressibility, 4-5, 6-30
Compressible flow, 6-30, 6-33
Compression wave, 6-31
Compressor
axial-centrifugal, 4-10
axial-flow, 4-9, 4-10
Index-3
centrifugal, 4-8
stall, 4-15
Confined area
approach to, 7-21
ground operations in, 7-22
operations, 7-18, 7-21, 7-22
takeoff from, 7-22
Coning, 5-9
Conservation of energy, 2-9
Control
aileron, 9-2
asymmetric, 10-4
collective, 7-13
directional, 8-49, 10-4
effectiveness, 8-46
elevator, 9-2
fOTces, 8-51
lateral, 8-51
longitudinal, 8-47
moments, 3-23
problem, 10-5
response, 8-47
speed, 10-4
surface operation theory, 8-44
systems
conventional, 8-55
flap, 8-56
full-power, 8-56
landing gear, 8-57
power-boosted, 8-55
spoiler, 8-56
stability augrnenter, 8-55
Coriolis force, 5-18
Critical
engine, 10-3, 10-4
rollover angle, 7-25
Cross-controlling, 7-7 through 7-9
Crosswind
landing, 8-50, 8-51, 9-24, 9-25
takeoff, 8-50, 8-51, 9-5
Curves
horsepower-required, 2-42, 2-43
performance, 2-39, 6-71
power-available, 2-44
power-required, 2-42
Cyclic
control response, 5-32, 5-33
feathering, 6-14, 6-18, 6-20, 8-59, 10-5, 10-8
pitch, 5-14, 5-16, 5-17

Damping, 3-5, 3-29, 3-30, 3-35


Deceleration, 6-58, 6-60, 6-68

Index-4
Decompression, 10-16, 10-17
Delta hinge, 6-20
Density, 2-17
Descents, 6-58, 9-10, 9-11, 10-7
Development of aerodynamic properties, 2-4, 2-5
Differential
airflow in rotor system, 6-14
collective pitch, 7-6
thrust, 7-6
Diffuser, 4-8
Dihedral, 3-18, 3-19, 8-51
Directional
instability, 3-14
stability, 3-13 through 3-22
Dissymmetry of lift, 6-14, 6-15, 6-20, 6-30
Distance
Accelerate-after-lift-off, 10-10
Accelerate-stop, 10-9, 10-10
Divergence
directional, 3-22
spiral, 3-22
Dorsal fin, 3-15
Downwash, 2-35, 6-4
Drag
blade, 5-18
brace, 5-20
coefficient of, 2-24
equation, 2-14, 2-15
form, 2-30
hinge, 5-3, 5-8, 5-18
induced, 2-22, 2-23, 2-34, 2-37, 2-38, 8-37
interference, 2-32
parasite, 2-22, 2-30, 2-40, 2-47, 3-23
profile, 2-22, 2-30
skin-friction, 2-31
total, 2-23
Driven region, 6-54
Driving region, 6-54
Droop stop, 5-9
Dutch roll, 3-22
Dynamic
pressure, 1-3, 1-4, 2-8, 2-9, 4-5
rollover, 7-25, 7-26
stability
during forward flight, 3-34
during hovering flight, 3-28, 3-29
fixed-wing, 3-1
nonoscillatory motion, 3-2, 3-3
oscillatory motion, 3-3, 3-4
rotary-wing, 3-22

Index-5
Effect, Venturi, 1-2, 1-3, 1-5
Effective translational lift, 6-29, 6-30
Elastic limit, 4-2
Elevator, 8-32, 8-45
Emergency situations, 6-39
Engine
failure, 6-51, 10-10 through iO-14
gas turbine, 4-7 through 4-17
reciprocating, 4-17 through 4-26
Entry, 6-57
Equation
Bernoulli, 1-5, 1-6
blade tip speed, 6-6
brake specific fuel consumption, 4-23
coefficient
of drag, 2-16
of force, 2-13
of lift, 2-16
drag, 2-14, 2-15
dynamic pressure, 1-4
force, 2-13
free-stream pressure, 2-10, 2-11
horsepower
.

available, 2-43
required, 2-41
required for maximum endurance, 2-44
lift, 2-14, 2-15
lift/drag, 2-24
maximum range, 2-44
parasite drag, 2-32, 2-33
power, 2-41
power-available, 2-43
power-required, 2-41
pressure differential on an airfoil, 2-11
s~all-speed, 8-18
static pressure, 1-3
total drag, 2-23
Equilibrium, 3-2, 3-11, 3-13, 6-55, 8-18
Equivalent shaft horsepower, 4-14
Error
instrument, 1-7
position, 1-7, 1-8
Exhaust nozzle, 4-13
Expansion space, 4-8

Fatigue, 4-3
Feather bearing blade grip, 5-3
Feathering, 8-59, 10-5, 10-8
Federal Aviation Regulation, 10-1
Flame-front propagation, 4-22
Flameout, 4-16
Flapping, 5-6, 5-14, 6-14, 6-15, 6-20, 6-21

Index-6
Flaps, 8-44
Flight
control systems, 8-44
envelope, 4-3, 4-6
Flight-path velocity, 2-4
Flow, 6-30, 6-32
compressible, 6-30, 6-33
fluid, 1-1
incompressible, 6-30, 6-33
patterns, 6-32, 6-34
states, 6-43, 6-44, 6-47
Force
aerodynamic, 2-13
centrifugal, 2-6, 2-7, 5-9, 5-10, 5-25, 7-3, 7-4, 8-32, 8-33
centripetal, 2-6, 2-7, 5-9, 6-80, 6-82, 8-33
coefficient of, 2-13
Coriolis, 5-18
equation, 2-13
friction, 8-39
total aerodynamic, 2-12, 2-36
Forward flight, 6-73, 6-80
Free-stream relative wind, 3-20
Frictional torque, 5-23
Friction force, 8-39
Frise ailerons, 3-21
Fuel-air ratio, 4-16, 4-20, 4-21
stoichiometric, 4-20
Fuselage, 5-1, 5-34
add-on, 5-33, 5-34
hovering attitude
single-rotor helicopters, 5-25
tandem-rotor helicopters, 5-28
pitch attitude, 5-33, 5-34

Gas engines
turbine, 4-7 through 4-17
turbine-propeller combination, 4-13
G-force, 4-4
Glide, 6-61, 8-29 through 8-32
Go-around, 9-26
Ground effect, 6-24, 6-27, 8-39, 8-48
Gyroscopic precession, 5-6
Gyro stability, 5-4, 5-6
Heading control, 5-23, 5-24, 7-6, 7-10, 7-11
single-rotor helicopters, 5-23
tandem-rotor helicopters, 5-23
Height-velocity diagram, 6-50 through 6-53, 6-64, 6-70, 6-76
High~lift devices, 8-4, 8-7
leading-edge, 8-10
trailing-edge, 8-8

Index-7
Binge
delta, 6-20
flapping, 6-20, 6-21
moments, 8-51, 8-52
offset, 6-21
pin, 5-3
Horizontal
component
of lift, 8-32
of thrust, 3-23
hinge pin, 5-3
Horn, 8-52
Horsepower-required curves, 2-42, 2-43
Hovering, 6-23 through 6-25
Hydroplaning, 8-42
Hypoxia, 10-14, 10-16, 10-17

lCAO standard atmosphere chart, 1-3, 1-4


Incompressible flow, 1-2, 6-30, 6-33
Induced
drag, 2-22, 2-23, 2-34, 2-37, 2-38, 8-37
flow, 6-4, 6-7, 6-8
Inertia, 1-1
In-ground effect, 6-24, 6-25, 6-27, 6-28
Interference drag, 2-32

Jet thrust, 4-13

Kinetic energy, 4-8

Landing, 8-38, 8-41, 9-23, 10-8


cross-slope, 7-25
crosswind, 8-50, 8-51, 9-24, 9-25
downhill, 7-25
power-off, 6-67, 6-68
uphill, 7-25
upslope, 7-25
velocity, 8-44
Lateral
control, 7-4
stability, 3-17 through 3-20
Law of conservation of angular momentum, 5-18
Lead and lag, 5-8, 5-18 through 5-20
blade drag, 5-18
Coriolis force, 5-18
Leading-edge radius, 2-4
Legs, 9-22
Level flight, 10-7
Lift, 2-5, 2-12, 2-17, 3-8, 5-3, 5-10, 9-11
coefficient of, 3-6, 3-7, 3-9, 8-4
differential, 5-3
dissymmetry of, 6-14, 6-15, 6-20, 6-30

Index-8
negative, 6-12
positive, 6-12
translational, 6-28 through 6-30
vector, 5-10
Lift-drag ratio, 2-24, 2-25
Limitations
aerodynamic, 4-5
operating, 4-3
structural, 4-4
weight and balance, 5-28
Linear motion, 5-5
Load factor, 6-85, 7-5, 8-20, 8-35
Longitudinal
control, 7-6
stability, 3-7 through 3-12
Mach numbers, 6-32, 6-34, 6-36 through 6-38
Main rotor, 5-1
Maneuvering flight, 6-79, 8-22
Mass
balancing, 8-54
conservation of, 1-2
flow rate, 1-2
Maximum
endurance, 2-44
range, 2-44
Mean
aerodynamic chord, 8-47
camber line, 2-3
Metal
design stress limit of, 4-3
elastic limit of, 4-2
fatigue, 4-3
static strength of, 4-1
stress and strain on, 4-2
ultimate stress on, 4-3
Minimum safe single-engine speed, 10-9
Moment
about the aerodynamic center, 3-8
arm, 3-11
pitching, 2-20, 2-21, 3-5 through 3-9, 3-26 through 3-28
Momentum theory, 2-7, 2-9, 6-7, 6-9
Motion sign conventions, 3-1

Newton's laws of motion, 1-1, 6-1


No-lift areas, 6-12
Nonoscillatory motion, 3-2, 3-3
Nonsymmetrical airfoils, 2-1, 2-2, 2-18, 2-20

Index-9
Offset hinge, 6-21
Operations
confined-area, 7-18, 7-21, 7-22
field, 7-18
ground, 7-22
pinnacle, 7-22, 7~23
ridgeline, 7-22
slope, 7-24 through 7-29
takeoff, 7-22, 7-23
terrain-flight, 7-23, 7-24
Oscillatory motion, 3-3, 3-4
Out-of-ground effect, 6-24 through 6-26
Overcontrolling, 5-32

Parallelogram method. See Vector solutions


Parasite drag, 2-22, 2-30, 2-40, 2-47, 3-23
Pedal-control linkage, 7-12
Pendular action, 5-32, 5-34
Performance
curves, 2-39, 6-71
effect
of altitude on, 2-47
of configuration on, 2-46
of weight on, 2-46
of wind on, 2-48
fixed-wing, 8-1
forward flight, 6-73, 6-80
rotary-wing, 6-1
speeds, 6-73 through 6-78, 10-1 through 10-14
P-factor, 8-49, 10-3. See also Asymmetric thrust
Phase
angle, 5-17
lag, 5-4, 5-6, 5-14 through 5-16
Pinnacle operations, 7-22, 7-23
Pitch, 7-2
angle, 5-17
attitude, 7-6
change, 5-16, 5-17
cyclic, 5-14, 5-16, 5-17
stability, 3-5. See also Longitudinal stability
Pitching moment, 2-20, 2-21, 3-5 through 3-9, 3-26 through 3-28
coefficient of the, 3-5 through 3-7
development of a, 3-23, 3-24
fuselage, 3-27, 3-28
Pitot-static .

system, 1-7
tube, 1-5
Plenum chamber, 4-8
Polar diagram, 8-26 through 8-28
Polygon method. See Vector solutions

Index-10
Positive
lift, 6-12
stall, 6-12
Power
available, 2-44, 6-73
charts, 10-7
control, 7-13
equations, 2-41
limit of performance, 6-86, 8-37
required curves, 2-42, 2-46
requirements, 2-40
settings, 7-13
Precession, 5-6
Pressure
center of, 2-4
differential, 2-9, 2-10, 2-37
dynamic, 1-3, 1-4, 2-8, 2-9, 4-5
fluid flow, 1-3
incompressible, 1-2
ratios, 4-8, 4-9
static, 1-3, 2-8, 2-9, 4-8
total, 1-2, 1-5, 2-8, 2-9
waves, 2-8
Pressurized aircraft, 10-14 through 10-17
Profile drag, 2-22, 2-30
Propeller
constant speed, 8-58
fixed pitch, 8-58
operating limitations, 8-60
Propulsion, 4-6, 4-7
Proverse roll, 3-21

Radius of turn, 6-83, 6-86, 6-87, 8-34


Rate
of climb, 8-23, 8-26
of descent, 6-61
of turn, 6-87, 8-38
Ratio
fuel-air, 4-16, 4-20, 4-21
lift-drag, 2-24
Reciprocating engines, 4-17 through 4-26
Reconnaissance, 7-20
Redline speed, 4-5
Relative wind, 2-4, 2-5, 3-20, 6-4
Resonance, 6-48
Resultant relative wind, 6-4, 6-10
Retreating blade stall, 6-39 through 6-42, 6-73
Reverse flow, 6-12
Ridgeline operations, 7-22
Roll
Dutch, 3-22
proverse, 3-21

Index-ll
rate, 8-51
stability. See Lateral stability
Rolling
friction, 8-39
moment, 3-17
motion, 7-25 through 7-27
Rotary-wing performance, 6-1
Rotational
inertia, 5-4
relative wind, 6-4, 6-6
Rotor
anti torque, 5-22
blade movements
feathering, 5-6, 5-8
flapping, 5-6 through 5-8, 5-17
hunting (lead and lag), 5-8, 5-19
comparison with gyro, 5-4
head control system, 5-11 through 5-16
system design, 3-23, 3-24, 5-1 through 5-34
Roundout, 9-23, 9-24
Rudder deflection, 8-32, 8-45, 10-5
Runway, 8-42

Scalars, 1-9
Service ceiling, single-engine, 10-12
Settling with power, 6-43, 6-47, 6-48
Shock wave, 6-34 through 6-37
Sideslip, 7-11, 8-49, 8-50, 10-6, 10-7, 10-11
Single-engine
absolute ceiling, 10-7
go-around, 10-14
operation, 10-2, 10-3
service ceiling, 10-12
Skidding, 10-5, 10-6
Skin-friction drag, 2-31
Slips, 9-23
Slipstream rotation, 8-49
Slope operations, 7-24 through 7-29
Slow flight, 9-15
Span, 5-3
Spark-ignition timing, 4-21
Spark-plug fouling, 4-25
Speed
angle-of-climb, 10-7
blade tip, 6-6
climbing stall, 8-26
maneuvering speed, 6-85
minimum safe single-engine, 10-9
of sound, 6-30, 6-32, 6-36
performance and operating, 10-1
rate-of-climb, 10-7
restrictions, 6-73
Index-12
Spins, 8-49, 9-19, 9-20
Spoilers, 3-21
Stability
cross effects and, 3-20
directional divergent, 3-22
dynamic, 3-3, 3-28
fixed-wing, 3-1
gyro, 5-4, 5-6
pitch, 3-5
roll, 3-17
rotary-wing, 3-22
rotor, 5-4
rotor angle of attack, 3-26, 3-30, 3-32
speed, 3-25, 3-31, 3-33
static, 3-2, 3-28
Stagnation point, 1-6, 1-7, 2-8
Stall
accidental, 9-21
aerodynamic, 8-12 through 8-14
attitudes, 9-18
characteristics, 8-11
negative, 6-12
positive, 6-12
power-on, 9-17
recovery, 8-17, 9-17
region, 6-55
speed, 8-4, 8-35, 8-36
strip, 8-15
warning, 8-14, 8-17
STANAG 3554 (Edition Two),
7-1, 9-1
State
autorotative, 6-45
normal thrusting, 6-44
vortex ring, 6-44 through 6-48
windmill brake, 6-45, 6-58
Static
pressure, 1-3, 2-8, 2-9, 4-8
stability, 3-2, 3-28, 3-30
strength, 4-1, 4-2
Steady-state
climb, 8-23
descent, 6-58 through 6-67
Stinger, 2-14
Strain, 4-1
Stress
metal, 4-1
ultimate, 4-3
Structural
failure, 6-37
limit of performance, 6-85, 8-35

Index-13
Subsonic
compressible flow, 6-34
incompressible flow, 6-33
Supercharging, 4-23, 4-24
Supersonic flow, 6-31, 6-32, 6-34, 6-36
Switch, flapper, 8-17
Symmetrical airfoils, 2-1, 2-2, 2-18, 2-20, 8-45

Tabs
balance, 8-54
servo, 8-53
trim, 8-54
Tail rotors, 5-1, 5-22, 6-20, 6-21
Takeoff, 8-38, 8-39, 9-4
crosswind, 8-50, 8-51, 9-5
distance, 8-40
minimum run, 9-6
normal, 9-4
operations, 7-22, 7-23
with obstacle clearance climb, 9-6
Taxiing, 9-1
Techniques
fixed-wing flight, 9-1
rotary-wing flight, 7-1
Terrain-flight operations, 7~23, 7-24
Thrust, 2-6, 2-40
asymmetric, 8-50, 10-2 through 10-5, 10-11, 10-13
axis, 3-12
differential, 7-6
jet, 4-13
reverse, 8-59
vertical and horizontal components of, 3-22
Tip-path plane, 3-23 through 3-29
Torque, 5-21, 5-23
Total
aerodynamic force, 2-12, 2-36
drag curve, 2-27
pressure. See Pressure
Traffic pattern, 7-14 through 7-17, 9-21, 9-22
Trailing-edge vortex, 2-34
Translating tendency, 5-24 through 5-26
Translational
flight, 6-28
lift, 6-28, 6-29
Transonic flow patterns, 6-34, 6-35
Transverse flow effect, 6-21, 6-22, 6-30
Triangle method. See Vector solutions
Trim tabs, 9-4
Trunnion, 5-3, 5-6, 5-26
Turbine
gas-generator, 4-12
power, 4-12

Index-14
Turbojets. See Engine, gas turbine
Turboprop power plant, 4-13
Turbulence, 7-18, 7-19
Turbulent flow, 8-14
Turning flight, 6-80 through 6-84, 8-33
Turns, 6-86, 7-3 through 7-5, 8-32, 8-33, 8-37
Twist
aerodynamic, 8-15
geometric, 8-15

Underslung, 5-20

Vector
downwash-velocity, 2-36
quantities, 1-9
relative-wind, 2-35
solutions, 1-9 through 1-12
vertical-velocity, 2-35
Velocity, 1-1
Venturi effect, 1-2, 1-3, 1-5
Vertical
component of lift,
8-33
component of
thrust, 3-23
force vector, 5-10
hinge pin, 5-3, 5-8, 5-18
lift, 7-3
stabilizer, 3-15
Vg diagram, 4-6
Vortex
bound, 2-34
generators, 8-7
ring state, 6-44 through 6-48
trailing-edge, 2-34
V See Performance speeds
speeds.

Weather-vane effect, 5-34


Weight, 2-6, 8-19
Weight and balance limitations, 5-28
Wind effect, 6-77
Windmill brake state, 6-45, 6-58
Windmilling, 10-5, 10-8
Wing flutter, 4-5

Yaw, 3-20, 3-21, 7-2, 7-12, 8-49


Yawing moments, 3-13, 3-14, 3-16
Yoke, 5-3

Index-15
4.'Z
~
IJ'i1
FM 1-203
3 OCTOBER 1988

By Order of the Secretary of the Army:

CARL E. VUONO
General, United States Army
Chief of Staff

Official:

WilLIAM J. MEEHAN II
Brigadier General, United States Army
The Adjutant General

DISTRIBUTION:

Active Army, USAR. and ARNG: To be distributed in accordance with DA Form 12-11A.
Requirements for Fundamentals of Flight (Qty rqr block no. 720).

~ U.S.GOYERNMENT PRINTING OFFICE'1988-526-027/80019


FM 1-203

FUNDAMENTALS
OF FLIGHT

HEADQUARTERS. DEPARTMENT OF THE ARMY


"

OCTOBER 1988
-..- ,----~_.-

You might also like