You are on page 1of 26

Nusa Flying International

AERODYNAMICS
(CPL and Instrument Rating)

Table of Contents

1. Load factors

▪ Definitions
▪ Types of loads
▪ Velocity/gust (VG) diagrams.

2. Aerodynamics of high-speed aircraft

▪ Mach Number
▪ Wide range of speed, vertical speed and pitch angle of jet aeroplanes.
▪ Swept wings

Page 1 of 26
1. LOAD FACTORS

LOAD FACTORS (PHAK chapter 4)

In aerodynamics, load factor is the ratio of the maximum load an aircraft can sustain to the gross
weight of the aircraft. The load factor is measured in Gs (acceleration of gravity), a unit of force equal
to the force exerted by gravity on a body at rest and indicates the force to which a body is subjected
when it is accelerated. Any force applied to an aircraft to deflect its flight from a straight line produces
a stress on its structure, and the amount of this force is the load factor. While a course in
aerodynamics is not a prerequisite for obtaining a pilot’s license, the competent pilot should have a
solid understanding of the forces that act on the aircraft, the advantageous use of these forces, and
the operating limitations of the aircraft being flown.

For example, a load factor of 3 means the total load on an aircraft’s structure is three times its gross
weight. Since load factors are expressed in terms of Gs, a load factor of 3 may be spoken of as 3 Gs,
or a load factor of 4 as 4 Gs.

If an aircraft is pulled up from a dive, subjecting the pilot to 3 Gs, he or she would be pressed down
into the seat with a force equal to three times his or her weight. Since modern aircraft operate at
significantly higher speeds than older aircraft, increasing the magnitude of the load factor, this effect
has become a primary consideration in the design of the structure of all aircraft.

With the structural design of aircraft planned to withstand only a certain amount of overload, a
knowledge of load factors has become essential for all pilots. Load factors are important for two
reasons:

1. It is possible for a pilot to impose a dangerous overload on the aircraft structures.

2. An increased load factor increases the stalling speed and makes stalls possible at seemingly safe
flight speeds.

Load Factors in Aircraft Design

The answer to the question “How strong should an aircraft be?” is determined largely by the use to
which the aircraft is subjected. This is a difficult problem because the maximum possible loads are
much too high for use in efficient design. It is true that any pilot can make a very hard landing or an
extremely sharp pull up from a dive, which would result in abnormal loads. However, such extremely
abnormal loads must be dismissed somewhat if aircraft are built that take off quickly, land slowly, and
carry worthwhile payloads.

The problem of load factors in aircraft design becomes how to determine the highest load factors that
can be expected in normal operation under various operational situations. These load factors are
called “limit load factors.” For reasons of safety, it is required that the aircraft be designed to
withstand these load factors without any structural damage. Although the Civil Aviation Safety
Regulations (CASR) requires the aircraft structure be capable of supporting one and one-half times
these limit load factors without failure, it is accepted that parts of the aircraft may bend or twist under
these loads and that some structural damage may occur.

Page 2 of 26
This 1.5 load limit factor is called the “factor of safety” and provides, to some extent, for loads higher
than those expected under normal and reasonable operation. This strength reserve is not something
which pilots should willfully abuse; rather, it is there for protection when encountering unexpected
conditions.

The above considerations apply to all loading conditions, whether they be due to gusts, maneuvers, or
landings. The gust load factor requirements now in effect are substantially the same as those that
have been in existence for years. Hundreds of thousands of operational hours have proven them
adequate for safety. Since the pilot has little control over gust load factors (except to reduce the
aircraft’s speed when rough air is encountered), the gust loading requirements are substantially the
same for most general aviation type aircraft regardless of their operational use. Generally, the gust
load factors control the design of aircraft which are intended for strictly non-acrobatic usage.

An entirely different situation exists in aircraft design with maneuvering load factors. It is necessary to
discuss this matter separately with respect to: (1) aircraft designed in accordance with the category
system (i.e., normal, utility, acrobatic); and (2) older designs built according to requirements which
did not provide for operational categories.

Aircraft designed under the category system are readily identified by a placard in the flight deck,
which states the operational category (or categories) in which the aircraft is certificated. The
maximum safe load factors (limit load factors) specified for aircraft in the various categories are:

¹For aircraft with gross weight of more than 4,000 pounds, the limit load factor is reduced.
To the limit loads given above, a safety factor of 50 percent is added.

There is an upward graduation in load factor with the increasing severity of maneuvers. The category
system provides for maximum utility of an aircraft. If normal operation alone is intended, the required
load factor (and consequently the weight of the aircraft) is less than if the aircraft is to be employed
in training or acrobatic maneuvers as they result in higher maneuvering loads.

Aircraft that do not have the category placard are designs that were constructed under earlier
engineering requirements in which no operational restrictions were specifically given to the pilots. For
aircraft of this type (up to weights of about 4,000 pounds), the required strength is comparable to
present-day utility category aircraft, and the same types of operation are permissible. For aircraft of
this type over 4,000 pounds, the load factors decrease with weight. These aircraft should be regarded
as being comparable to the normal category aircraft designed under the category system, and they
should be operated accordingly.

Load Factors in Steep Turns

In a constant altitude, coordinated turn in any aircraft, the load factor is the result of two forces:
centrifugal force and gravity. [Figure 4-44] For any given bank angle, the ROT varies with the
airspeed—the higher the speed, the slower the ROT. This compensates for added centrifugal force,
allowing the load factor to remain the same.

Page 3 of 26
Figure 4-44. Two forces cause load factor during turns

Figure 4-45 reveals an important fact about turns—the load factor increases at a terrific rate after a
bank has reached 45° or 50°. The load factor for any aircraft in a 60° bank is 2 Gs. The load factor in
an 80° bank is 5.76 Gs. The wing must produce lift equal to these load factors if altitude is to be
maintained.

1.
Figure 4-45. Angle of bank changes load factor

It should be noted how rapidly the line denoting load factor rises as it approaches the 90° bank line,
which it never quite reaches because a 90° banked, constant altitude turn is not mathematically
possible. An aircraft may be banked to 90°, but not in a coordinated turn. An aircraft which can be
held in a 90° banked slipping turn is capable of straight knife-edged flight. At slightly more than 80°,
the load factor exceeds the limit of 6 Gs, the limit load factor of an acrobatic aircraft.

For a coordinated, constant altitude turn, the approximate maximum bank for the average general
aviation aircraft is 60°. This bank and its resultant necessary power setting reach the limit of this type

Page 4 of 26
of aircraft. An additional 10° bank increases the load factor by approximately 1 G, bringing it close to
the yield point established for these aircraft. [Figure 4-46]

Figure 4-46. Load factor changes stall speed

Load Factors and Stalling Speeds

Any aircraft, within the limits of its structure, may be stalled at any airspeed. When a sufficiently high
AOA is imposed, the smooth flow of air over an airfoil breaks up and separates, producing an abrupt
change of flight characteristics and a sudden loss of lift, which results in a stall.

A study of this effect has revealed that the aircraft’s stalling speed increases in proportion to the
square root of the load factor. This means that an aircraft with a normal unaccelerated stalling speed
of 50 knots can be stalled at 100 knots by inducing a load factor of 4 Gs. If it were possible for this
aircraft to withstand a load factor of nine, it could be stalled at a speed of 150 knots. A pilot should be
aware:

▪ Of the danger of inadvertently stalling the aircraft by increasing the load factor, as in a steep
turn or spiral;
▪ When intentionally stalling an aircraft above its design maneuvering speed, a tremendous load
factor is imposed.

Figures 4-45 and 4-46 show that banking an aircraft greater than 72° in a steep turn produces a load
factor of 3, and the stalling speed is increased significantly. If this turn is made in an aircraft with a
normal unaccelerated stalling speed of 45 knots, the airspeed must be kept greater than 75 knots to
prevent inducing a stall. A similar effect is experienced in a quick pull up, or any maneuver producing
load factors above 1 G. This sudden, unexpected loss of control, particularly in a steep turn or abrupt
application of the back elevator control near the ground, has caused many accidents.

Since the load factor is squared as the stalling speed doubles, tremendous loads may be imposed on
structures by stalling an aircraft at relatively high airspeeds.

Page 5 of 26
The maximum speed at which an aircraft may be stalled safely is now determined for all new designs.
This speed is called the “design maneuvering speed” (VA) and must be entered in the DGCA-approved
Airplane Flight Manual/Pilot’s Operating Handbook (AFM/POH) of all recently designed aircraft. For
older general aviation aircraft, this speed is approximately 1.7 times the normal stalling speed. Thus,
an older aircraft which normally stalls at 60 knots must never be stalled at above 102 knots (60 knots
x 1.7 = 102 knots). An aircraft with a normal stalling speed of 60 knots stalled at 102 knots
undergoes a load factor equal to the square of the increase in speed, or 2.89 Gs (1.7 x 1.7 = 2.89
Gs). (The above figures are approximations to be considered as a guide, and are not the exact
answers to any set of problems. The design maneuvering speed should be determined from the
particular aircraft’s operating limitations provided by the manufacturer.)

Since the leverage in the control system varies with different aircraft (some types employ “balanced”
control surfaces while others do not), the pressure exerted by the pilot on the controls cannot be
accepted as an index of the load factors produced in different aircraft. In most cases, load factors can
be judged by the experienced pilot from the feel of seat pressure. Load factors can also be measured
by an instrument called an “accelerometer,” but this instrument is not common in general aviation
training aircraft. The development of the ability to judge load factors from the feel of their effect on
the body is important. A knowledge of these principles is essential to the development of the ability to
estimate load factors.

A thorough knowledge of load factors induced by varying degrees of bank and the VA aids in the
prevention of two of the most serious types of accidents:

1. Stalls from steep turns or excessive maneuvering near the ground


2. Structural failures during acrobatics or other violent maneuvers resulting from loss of control

Load Factors and Flight Maneuvers

Critical load factors apply to all flight maneuvers except unaccelerated straight flight where a load
factor of 1 G is always present. Certain maneuvers considered in this section are known to involve
relatively high load factors.

Turns

Increased load factors are a characteristic of all banked turns. As noted in the section on load factors
in steep turns, load factors become significant to both flight performance and load on wing structure
as the bank increases beyond approximately 45°.

The yield factor of the average light plane is reached at a bank of approximately 70° to 75°, and the
stalling speed is increased by approximately one-half at a bank of approximately 63°.

Stalls

The normal stall entered from straight-and-level flight, or an unaccelerated straight climb, does not
produce added load factors beyond the 1 G of straight-and-level flight. As the stall occurs, however,
this load factor may be reduced toward zero, the factor at which nothing seems to have weight. The
pilot experiences a sensation of “floating free in space.” If recovery is effected by snapping the
elevator control forward, negative load factors (or those that impose a down load on the wings and
raise the pilot from the seat) may be produced.

Page 6 of 26
During the pull up following stall recovery, significant load factors are sometimes induced. These may
be further increased inadvertently during excessive diving (and consequently high airspeed) and
abrupt pull ups to level flight. One usually leads to the other, thus increasing the load factor. Abrupt
pull ups at high diving speeds may impose critical loads on aircraft structures and may produce
recurrent or secondary stalls by increasing the AOA to that of stalling.

As a generalization, a recovery from a stall made by diving only to cruising or design maneuvering
airspeed, with a gradual pull up as soon as the airspeed is safely above stalling, can be effected with
a load factor not to exceed 2 or 2.5 Gs. A higher load factor should never be necessary unless
recovery has been effected with the aircraft’s nose near or beyond the vertical attitude, or at
extremely low altitudes to avoid diving into the ground.

Spins

A stabilized spin is not different from a stall in any element other than rotation and the same load
factor considerations apply to spin recovery as apply to stall recovery. Since spin recoveries are
usually effected with the nose much lower than is common in stall recoveries, higher airspeeds and
consequently higher load factors are to be expected. The load factor in a proper spin recovery usually
is found to be about 2.5 Gs.

The load factor during a spin varies with the spin characteristics of each aircraft, but is usually found
to be slightly above the 1 G of level flight. There are two reasons for this:

1. Airspeed in a spin is very low, usually within 2 knots of the unaccelerated stalling speeds.
2. Aircraft pivots, rather than turns, while it is in a spin.

High Speed Stalls

The average light plane is not built to withstand the repeated application of load factors common to
high speed stalls. The load factor necessary for these maneuvers produces a stress on the wings and
tail structure, which does not leave a reasonable margin of safety in most light aircraft.

The only way this stall can be induced at an airspeed above normal stalling involves the imposition of
an added load factor, which may be accomplished by a severe pull on the elevator control. A speed of
1.7 times stalling speed (about 102 knots in a light aircraft with a stalling speed of 60 knots) produces
a load factor of 3 Gs. Only a very narrow margin for error can be allowed for acrobatics in light
aircraft. To illustrate how rapidly the load factor increases with airspeed, a high-speed stall at 112
knots in the same aircraft would produce a load factor of 4 Gs.

Chandelles and Lazy Eights

A chandelle is a maximum performance climbing turn beginning from approximately straight-and-level


flight, and ending at the completion of a precise 180° of turn in a wings-level, nose-high attitude at
the minimum controllable airspeed. In this flight maneuver, the aircraft is in a steep climbing turn and
almost stalls to gain altitude while changing direction. A lazy eight derives its name from the manner
in which the extended longitudinal axis of the aircraft is made to trace a flight pattern in the form of a
figure “8” lying on its side. It would be difficult to make a definite statement concerning load factors in
these maneuvers as both involve smooth, shallow dives and pull ups. The load factors incurred
depend directly on the speed of the dives and the abruptness of the pull ups during these maneuvers.

Page 7 of 26
Generally, the better the maneuver is performed, the less extreme the load factor induced. A
chandelle or lazy eight in which the pull-up produces a load factor greater than 2 Gs will not result in
as great a gain in altitude, and in low-powered aircraft it may result in a net loss of altitude.

The smoothest pull up possible, with a moderate load factor, delivers the greatest gain in altitude in a
chandelle and results in a better overall performance in both chandelles and lazy eights. The
recommended entry speed for these maneuvers is generally near the manufacturer’s design
maneuvering speed which allows maximum development of load factors without exceeding the load
limits.

Rough Air

All standard certificated aircraft are designed to withstand loads imposed by gusts of considerable
intensity. Gust load factors increase with increasing airspeed, and the strength used for design
purposes usually corresponds to the highest level flight speed. In extremely rough air, as in
thunderstorms or frontal conditions, it is wise to reduce the speed to the design maneuvering speed.
Regardless of the speed held, there may be gusts that can produce loads which exceed the load
limits.

Each specific aircraft is designed with a specific G loading that can be imposed on the aircraft without
causing structural damage. There are two types of load factors factored into aircraft design, limit load
and ultimate load. The limit load is a force applied to an aircraft that causes a bending of the aircraft
structure that does not return to the original shape. The ultimate load is the load factor applied to the
aircraft beyond the limit load and at which point the aircraft material experiences structural failure
(breakage). Load factors lower than the limit load can be sustained without compromising the
integrity of the aircraft structure.

Speeds up to but not exceeding the maneuvering speed allows an aircraft to stall prior to experiencing
an increase in load factor that would exceed the limit load of the aircraft.

Most AFM/POH now include turbulent air penetration information, which help today’s pilots safely fly
aircraft capable of a wide range of speeds and altitudes. It is important for the pilot to remember that
the maximum “never-exceed” placard dive speeds are determined for smooth air only. High speed
dives or acrobatics involving speed above the known maneuvering speed should never be practiced in
rough or turbulent air.

Vg Diagram

The flight operating strength of an aircraft is presented on a graph whose vertical scale is based on
load factor. [Figure 4-47] The diagram is called a Vg diagram—velocity versus G loads or load factor.
Each aircraft has its own Vg diagram which is valid at a certain weight and altitude.

The lines of maximum lift capability (curved lines) are the first items of importance on the Vg
diagram. The aircraft in the Figure 4-47 is capable of developing no more than +1 G at 62 mph, the
wing level stall speed of the aircraft. Since the maximum load factor varies with the square of the
airspeed, the maximum positive lift capability of this aircraft is 2 G at 92 mph, 3 G at 112 mph, 4.4 G
at 137 mph, and so forth. Any load factor above this line is unavailable aerodynamically (i.e., the
aircraft cannot fly above the line of maximum lift capability because it stalls). The same situation
exists for negative lift flight with the exception that the speed necessary to produce a given negative
load factor is higher than that to produce the same positive load factor.

Page 8 of 26
If the aircraft is flown at a positive load factor greater than the positive limit load factor of 4.4,
structural damage is possible. When the aircraft is operated in this region, objectionable permanent
deformation of the primary structure may take place and a high rate of fatigue damage is incurred.
Operation above the limit load factor must be avoided in normal operation.

There are two other points of importance on the Vg diagram. One point is the intersection of the
positive limit load factor and the line of maximum positive lift capability. The airspeed at this point is
the minimum airspeed at which the limit load can be developed aerodynamically. Any airspeed greater
than this provides a positive lift capability sufficient to damage the aircraft. Conversely, any airspeed
less than this does not provide positive lift capability sufficient to cause damage from excessive flight
loads. The usual term given to this speed is “maneuvering speed,” since consideration of subsonic
aerodynamics would predict minimum usable turn radius or maneuverability to occur at this condition.
The maneuver speed is a valuable reference point, since an aircraft operating below this point cannot
produce a damaging positive flight load. Any combination of maneuver and gust cannot create
damage due to excess airload when the aircraft is below the maneuver speed.

Figure 4-47. Typical Vg diagram

The other point of importance on the Vg diagram is the intersection of the negative limit load factor
and line of maximum negative lift capability. Any airspeed greater than this provides a negative lift
capability sufficient to damage the aircraft; any airspeed less than this does not provide negative lift
capability sufficient to damage the aircraft from excessive flight loads.

The limit airspeed (or redline speed) is a design reference point for the aircraft—this aircraft is limited
to 225 mph. If flight is attempted beyond the limit airspeed, structural damage or structural failure
may result from a variety of phenomena.

Page 9 of 26
The aircraft in flight is limited to a regime of airspeeds and Gs which do not exceed the limit (or
redline) speed, do not exceed the limit load factor, and cannot exceed the maximum lift capability.
The aircraft must be operated within this “envelope” to prevent structural damage and ensure the
anticipated service lift of the aircraft is obtained. The pilot must appreciate the Vg diagram as
describing the allowable combination of airspeeds and load factors for safe operation. Any maneuver,
gust, or gust plus maneuver outside the structural envelope can cause structural damage and
effectively shorten the service life of the aircraft.

2. AERODYNAMICS OF HIGHT-SPEED AIRCRAFT

HIGHT SPEED FLIGHT (PJ Swatton, Part 6, chapter 15)

General Introduction

High-speed flight can be subdivided into subsonic, transonic, supersonic and hypersonic regimes. In
the transonic speed range both subsonic and supersonic speeds exist in the airflow around the
aeroplane but it is the Mach number that determines an aeroplane’s handling characteristics. The
flight regimes are subdivided by speed as shown in Table 15.1.

An aeroplane when moving through the air transmits a disturbance pressure wave in all directions.
That which travels ahead of the aeroplane is particularly important because it ‘warns’ the air in front
of the aeroplane of the approach of the aircraft and enables it to divide to allow its passage with the
least amount of disturbance.

Because air is compressible the changing air pressure created by the movement of the aeroplane is
accompanied by a change of air temperature and a change of air density. Owing to this, the speed of
movement generated by the pressure wave is equal to the speed of sound. Sound waves are audible
pressure waves travelling at the speed of sound.

As the forward speed of the aeroplane increases the distance the air ahead of the aeroplane that is
influenced by the approaching pressure wave decreases; there is also a change to the airflow and
pressure patterns surrounding the aeroplane. Eventually, this results in changes to the
manoeuvrability, stability and control characteristics of the aeroplane.

High-Speed Definitions
Page 10 of 26
The terms used throughout this chapter are defined as follows:

a. Free-Stream Mach number (MFS). The Mach number of the air at a point, unaffected by
the passage of the aeroplane but measured relative to the speed of the aeroplane, is the
Free Stream Mach number. (Mfs). If the speed of the aeroplane is above the critical Mach
number (Mcrit) a shockwave may well form, even if the Mfs is below Mach 1.

b. Local Mach number (ML). This is the ratio of the speed of the airflow at a point on the
aeroplane to the speed of sound at the same point.

c. Critical Mach number (MCRIT). As Mfs increases so do the local Mach numbers. Mcrit is
that Mfs at which any Ml has reached unity. It is the lowest speed in the transonic range.

d. Critical Drag Rise Mach number (MCDR). This is theMfs at which because of
shockwaves, the CD for a given angle of attack increases significantly.

e. Detachment Mach number (MDET). This is the speed at which the bow shockwave of an
accelerating aeroplane attaches to the leading edge of the wing or detaches if the aeroplane
is decelerating.

f. Indicated Mach number. TAS is the difference between pitot pressure and static pressure.
LSS is a function of static pressure and air density. Because air density is common to both
TAS and LSS, both can be expressed as pressure ratios; this is what the Machmeter
measures. Indicated Mach number is therefore the ratio of the dynamic pressure to the static
pressure.

g. Shock Stall. The airflow over an aeroplane’s wings is disturbed when flying at or near the
critical Mach number. This causes the separation of the boundary layer from the upper
surface of the wing behind the shockwave and is the shock stall, which is described in greater
detail later in this chapter.

h. The Speed of Sound (a). The speed of propagation of a very small pressure disturbance in
a fluid in specified conditions. The local speed of sound through the air is equal to 38.94
multiplied by the square root of the absolute temperature (A). Therefore, as air temperature
decreases so also does the local speed of sound (LSS).

LSS = 38.94 √A in kt

i. True Mach number (M). The ratio of the speed of an object to the value of the speed of
sound in the same environmental conditions is the Mach number; it has no units of
measurement because it is a ratio.

Mach number = TAS ÷ LSS in kt


Therefore, Mach number = TAS ÷ 38.94 √A

THE SPEED OF SOUND (Atlantic, chapter 14)

The Speed of Sound is defined as the rate at which small pressure disturbances are propagated
through the air, which is solely a function of air temperature.

Page 11 of 26
The speed of sound is therefore solely dependent on the ambient air temperature and varies with
altitude as illustrated in the following table.

The speed of sound at sea level is approximately 660 Knots and steadily reduces up to the base of the
tropopause, where after it remains constant.

Fully subsonic aircraft can be heard approaching because they send out pressure disturbances, or
waves in all directions, which travel at the Speed of Sound. This enables an approaching aircraft to be
heard, and more importantly for the aircraft, the air to be warned of its approach. The sound is
transmitted to ones ears by way of a series of molecular collisions. Conversely aircraft travelling
supersonically can not be heard because the air ahead gets no warning of their approach, and no
molecular collisions take place.

Mach Number

Mach number is named after Ernst Mach, an Austrian physicist, and is the ratio of the actual speed of
a body or flow, to the speed of sound in the surrounding atmosphere, so that:

For example, if an aircraft is travelling at half the speed of sound the Mach number will be 0.5. From
the basic Mach number the following definitions can be derived:

Free Stream Mach Number (MFS). This is the Mach number of the airflow sufficiently remote
from the aircraft so as not to be affected by it, so that:

Local Mach Number (ML). This is the actual speed of the flow over a surface. For example the
airflow accelerates as it passes over a wings surface, so that the local Mach number on top of the
wing is always greater than the free stream Mach number (Fig. 14.5).
Page 12 of 26
Figure 14.5

For a given aerofoil section and free stream Mach number the local Mach number also varies
directly with changes in the angle of attack.

Critical Mach Number (MCRIT). This is the value of the free stream Mach number when the
local Mach number first becomes sonic anywhere on an aircraft, which normally initially occurs on
the upper surface of an aircraft's wing near to the point of maximum thickness (Fig. 14.6).

Figure 14.6

Flight Speed Classifications

In flight the airspeed of an aircraft determines whether the airflow is travelling subsonically or
supersonically around the aircraft:

Subsonic. Aircraft speeds approximately Mach 0.75 or less, where the total airflow around an
aircraft is travelling at a speed less than the speed of sound.

Transonic. Aircraft speeds between Mach 0.75 and Mach 1.2, where the airflow around an
aircraft is partly subsonic, and partly supersonic.

Supersonic. Aircraft speeds between Mach 1.2 and Mach 5.0, where the total airflow around an
aircraft is travelling at a speed greater than the speed of sound.

THE SHOCKWAVE (PJ Swatton, Part 6, chapter 15)

Compressibility

An aeroplane flying at low, subsonic, speeds causes air-pressure changes as a result of its movement
but they are relatively small. It is convenient, therefore, to treat the air as though it could not be
compressed. However, the air-pressure changes caused by the movement of an aeroplane at high
speed, close to the speed of sound, are considerable and cannot be ignored. The air is compressible
and its effect on the aeroplane must be accounted for when it is travelling at high speed.

An aeroplane moving through the air creates a pressure wave around it, which is propagated away
from it in all directions simultaneously at the speed of sound. For an aeroplane flying at a relatively
low speed, less than the speed of sound, the air-pressure wave ahead of the aeroplane can move
Page 13 of 26
away from the aeroplane. Consequently, air-pressure, air-density, air-temperature and air-velocity
changes take place ahead of the leading edge of the wings and are a gradual process.

If, however, the aeroplane is travelling at the speed of sound then the air-pressure wave ahead of the
aeroplane is unable to move away from the aeroplane and a wave of compressed air builds up at the
leading edge of the wing. This means that the rise of static air pressure, air density, air temperature
and the local speed of sound together with the decrease of the Mach number and total pressure take
place rapidly almost instantaneously across the depth of the compression wave. This phenomenon is
often referred to as the normal shockwave or the shock front, which is positioned at ninety degrees to
the direction of travel of the aeroplane. The air behind the shockwave travels at subsonic speeds. The
compression of a normal shockwave is greater than that of an oblique shockwave at the same Mach
number.

Air-Pressure-Wave Patterns

The changing pressure-wave pattern with the increasing speed of an aeroplane is three-dimensional
and is shown in Figure 15.2. The speed of propagation in the diagram is ‘a’ and is equal to the local
speed of sound. The vector ‘v’ shows the speed of the aeroplane. There are three speeds depicted;
subsonic, sonic and supersonic.

Figure 15.1 Shockwave Formation

Page 14 of 26
Figure 15.2 Air-Pressure Waves

Page 15 of 26
Subsonic

The length of the vector ‘v’ in Figure 15.2(a) is less than the length of ‘a’, therefore, although the
pressure waves are compressed ahead of the aeroplane, they maintain their separation.

Sonic

In Figure 15.2(b) the length of vector ‘v’ is equal to the length of ‘a.’ This shows that the air-pressure
waves do not move ahead of the aeroplane but compress to form a straight-line Mach wave at right
angles to the direction of movement of the aeroplane and is referred to as a ‘normal shockwave.’ It is
a plane of discontinuity normal to the local airflow direction at which the compression of the air and
the loss of total pressure are greater than those of an oblique shockwave.

The loss of total pressure in the shockwave is due to the conversion of the flo kinetic energy into heat
energy. This effect can be minimized if the airflow just ahead of the shockwave is only just
supersonic. The total temperature of a normal shockwave is higher than that of the oblique
shockwave.

The air behind the shockwave is subsonic and, although there are large variations to the lift and drag,
the direction of the airflow is unchanged. This can cause variations to the pitching moment and to the
trim of the aeroplane; it may also affect the operation of the controls. Shock-induced separation can
occur behind a strong normal shockwave independent of the angle of attack.

Supersonic

The length of ‘v’ in Figure 15.2(c) is longer than the radius ‘a’, indicating that the aeroplane is
travelling faster than the speed of sound. This produces a boundary beyond which no pressure wave
passes; this is the ‘oblique Mach wave.’ It is a three-dimensional boundary and is therefore conic in
shape and within which all disturbances caused by the aeroplane’s movement are contained. The
aeroplane centre of pressure is further aft than at subsonic speeds and the speed of the airflow after

Page 16 of 26
passing through an oblique shockwave is greater than that of the speed of sound. The air
temperature increase is less than a normal shockwave.

The High-Speed CP

The position of the centre of pressure (CP) during transonic flight is determined by the angle of attack
and therefore is speed dependent. At Mach 0.75 the CP is located at approximately 20% of the length
of the chordline from the leading edge of the aerofoil. As the speed increases the CP moves
progressively aft, increasing the longitudinal static stability and requiring an increasing amount of
nose-up pitch input of the stabiliser, at Mach 1.0 the CP is at approximately 45% of the MAC from the
leading edge, until at Mach 1.4 the CP positioned at approximately 50% of the length of the chordline
from the leading edge of the aerofoil.

For swept-wing aeroplanes this rearward movement of the CP restricts their maximum speed.
Transferring fuel between tanks to move the CG can generate a correcting moment that counteracts
the rearward movement of the CP. However, this method is limited because as the flight progresses
and the fuel is used, although the aeroplane mass decreases enabling it to travel faster, there is less
fuel available for transference to move the CG and counteract the effect of the movement of the CP.

Critical Mach Number (MCRIT)

The critical Mach number (Mcrit) is the lower limit of a speed band, known as the transonic range, in
which the local Mach number may be either subsonic or supersonic. Its value varies with the angle of
attack and is the free-stream Mach number at which sonic speed is reached over the upper surface of
a wing. In other words, it is the speed at which the local Mach number is equal to the free-airstream
Mach number. It is also the highest speed possible without supersonic flow over the wing. For
example, if the air passing over a wing at some point reaches Mach 1, but the wing is only moving at
Mach 0.8 then the critical Mach number is 0.8.

Generally, a thicker wing with a large amount of camber has a lower critical Mach number than a thin
wing with little camber because the airflow over its upper surface accelerates to a higher speed than it
would over a thin wing. At the design stage of an aeroplane the inclusion of the following
characteristics either individually or in any combination will increase the value of the critical Mach
number but the largest increase will result from a combination of a and b:

a. A swept wing having the same wing area and loading.


b. A thin aerofoil cross-section (low thickness to chord ratio).
c. Area ruling. (Described later in this chapter).

The critical Mach number is inversely proportional to the angle of attack, i.e. Mcrit decreases with
increasing angle of attack and vice versa. This is because, as the angle of attack increases the local
peak velocity of the airflow over the wing also increases. It therefore reaches Mach 1 sooner than it
would have done at a lower angle of attack, thus the critical Mach number is less.

As a flight progresses and fuel is burnt the aeroplane mass decreases and it is able to accelerate to
a higher maximum speed. If the pilot maintains level flight and does not touch the throttles the
aeroplane will continually accelerate and the angle of attack must be continually decreased to allow
for the increased IAS. In such a case, the airflow peak velocity will not be attained until the aeroplane
is at a higher maximum speed. Thus, the critical Mach number increases continuously throughout a
flight assuming the aeroplane is permitted to accelerate in level flight and the angle of attack is
adjusted accordingly.

Page 17 of 26
If, in level flight, the pilot maintains a constant IAS as the aeroplane mass decreases then the angle of
attack decreases for that speed but not as much as it would have done had the aeroplane been
allowed to accelerate. As a result, as the flight progresses, the angle of attack diverges more from
that appropriate to the maximum speed for its mass, and although the critical Mach number will
increase it is still less than it could be.

At speeds in excess of the critical Mach number, a swept-wing aeroplane loses lift in the area of the
wing roots, which decreases stick-force stability and produces a sudden increase to the drag
coefficient. As a result, the aeroplane will experience buffeting, the first evidence of the formation of a
shockwave on the upper surface of the wing at the wing root, and a tendency to pitch nose-down.

During acceleration the pressure-distribution pattern on the upper surface of the wing becomes
irregular as the airflow becomes supersonic at Mcrit. However, with continued acceleration, when the
whole airflowbecomes supersonic the pressure pattern steadies and becomes a geometrical
parallelogram, some refer to as a rectangular pattern.

The Flying Controls

The deflection of a conventional hinged-flap type of control creates a force ahead of the control
surface that is normally greater than that which the control surface itself produces. When there is a
shockwave established on the lifting surface ahead of the control surface this is no longer the case
because pressure variations cannot be transmitted forward through the shockwave.

The shockwave moves slightly aft when an aileron is deflected downwards. This effectively diminishes
the result of deflecting the control surface because only the airflow passing over the control surface is
modified. This effect is exacerbated when the shockwave reaches the trailing edge of the aerofoil and
the control effectiveness is decreased even further. In transonic flight the ailerons are less effective
than in subsonic flight because aileron deflection only partially affects the pressure distribution around
the wing.

These characteristics have led to the substitution of the tailplane and elevator combination in
highspeed aeroplanes with a fully flying horizontal stabiliser. On these aeroplanes the conventional
wingtip ailerons are replaced with differential spoilers and/or inboard ailerons that are for use when
the aeroplane is flying at transonic speeds. They are fitted partially to combat the adverse effects just
described but primarily to avoid the adverse effects of wing twist that can negate or even reverse the
effect of conventional ailerons when deflected at high speed. The complete reversal of the aileron
controls occurs at high speeds generally outside of the flight envelope for most aeroplanes. However,
the effect of aeroelastic-distortion, the combination of decreased wing rigidity and high airspeed,
could manifest itself as a reduced rate of roll.

The movement of the shockwave rearward to the trailing edge of the aerofoil during transition to
supersonic speeds means that the shockwave must transit the control surface. Any deflection of the
control surface during this transition will move the shockwave, which results in significant changes in
the location and magnitude of the aerodynamic forces acting on the control. Consequently, the control
surface will experience a high frequency vibration referred to as ‘buzz.’

Because of the increased input forces required from the pilot during flight at transonic speeds and the
large variation of the stick forces encountered during the transition of the shockwave over the control
surface, most high-speed aeroplanes are fitted with fully power-operated controls.

Page 18 of 26
SWEPT WINGS

The primary reason that the swept-wing design is used for most jet transport aeroplanes is to
increase the value of the critical Mach number for that type of aeroplane. It is the lowest speed of the
free airflow that when passing over some part of the aeroplane becomes supersonic. Usually, it is the
upper surface of the wing, over which the airflow accelerates to a speed of Mach 1.

Therefore, a swept wing delays the onset of the effects of compressibility and delays the airflow from
becoming supersonic. It is best employed for aeroplanes that operate in the transonic regime of flight
because the sweepback necessary to delay the drag rise at extremely high Mach numbers or
continuous flight in the supersonic regime is too great to be practical.

The Effect of Sweepback

The critical Mach number of a wing of a given thickness/chord ratio and aspect ratio can be increased
by including a high-angled sweepback in the design. The angle of sweepback of such wings is limited
by the practicality of their construction. Although it is assumed that any sweepback is better than
none, to be of any significant value the sweepback should be at least 30◦. Nevertheless, the inclusion
of a relatively small angle of sweepback in the design of any wing increases the critical Mach number.

The Advantages of Sweepback

The advantages of an aeroplane having swept-back wings are:

a. Mcrit is increased in direct proportion to the sweep angle.


b. Cd is decreased in direct proportion to the angle of sweep.
c. Drag divergence is delayed to a higher speed.
d. Static directional stability is improved.
e. Static lateral stability is improved in a similar way to dihedral.
f. For a given aspect ratio and wing loading, the aeroplane is less sensitive to gusts than a straight
wing.

The Disadvantages of Sweepback

Despite their advantages, swept wings have the following disadvantages:

a. Trailing-edge controls, such as flaps and wing-tip ailerons, are less effective because they are not
at right angles to the airflow. Some flap systems only produce an increase of lift of 50% of that
which would have been produced by the same flap on a straight-winged aeroplane.
b. A swept wing of the same wing area and aspect ratio as that of a straight-winged aeroplane has
a greater wing span, which increases its mass. This causes greater bending and stress towards
the wing tip. It is also subject to the twisting effect of the wing in high-speed flight that
diminishes the effectiveness of wing-tip ailerons.
a. When combined with taper there is a strong tendency for the wing to tip stall first. This is
because, although taper produces a strong local lift coefficient towards the wing tip similar to
sweepback, there is a strong spanwise flow of the boundary layer towards the wing tip,
particularly at high angles of attack, that results in a low-energy pool at the wing tip which easily
separates from the wing surface.

Table 15.4 The Theoretical Approximate Effects of Sweepback. (Assuming a wing of moderate aspect
ratio in transonic flight.) (below)

Page 19 of 26
Figure 15.5 The Effect of Wing Sweep (below)

(a) STRAIGHT WING (b) SWEPT WING

REMEDIAL DESIGN FEATURES

To overcome some of these disadvantages some jet transport aeroplanes having a swept-wing design
are fitted with two sets of ailerons, outboard are the low-speed ailerons and inboard are the high-
speed ailerons. See Figure 15.6.

Low-Speed Ailerons

The low-speed ailerons, mounted just inboard of the wing tip, are provided for use at the speeds
experienced during take-off, take-off climb and landing. They are effective for such use because of
the large aerodynamic force couple they generate, due to their long moment arm, produces the
required rate of roll at these low speeds. These ailerons are de-activated when the slats and flaps are
retracted or the speed is above a predetermined value during high-speed flight.

Page 20 of 26
High-Speed Ailerons

The high-speed ailerons are used for the cruise portion of a flight because, being mounted inboard
towards the wing root where the wing is thicker they do not cause twisting of the wing. They also
have the advantage of generating a smaller aerodynamic force couple, which is essential to produce a
lower roll rate. High rates of roll are not required at high speeds. Inboard ailerons are less effective at
transonic speeds than at subsonic speeds because their deflection only partially affects the pressure
distribution around the wing. These ailerons are locked during low-speed flight.

Figure 15.6 Swept-Wing Ailerons

HIGH-SPEED-FLIGHT CHARACTERISTICS

High-Speed Buffet

High-speed buffet is the rapid movement of the control surfaces caused by turbulent airflow passing
over them. It can only occur when the Mach number is above the critical Mach number and is likely to
be accompanied by a tendency for the aeroplane to pitch nose-down. The maximum operating
altitude of an aeroplane can be restricted by the load factor. Above that altitude any turbulence may
induce Mach buffet, which could cause the aeroplane to exceed the maximum permitted load factor.

The airflow behind a shockwave is turbulent and can cause a rapid vibrating movement of the
horizontal stabiliser surfaces and is often referred to as the ‘shock stall’ and is the primary cause of
high-speed buffet. A secondary cause of the rapid movement of the control surfaces, especially with

Page 21 of 26
supercritical aerofoil sections, is when the wing reaches a speed ofMcrit the relatively weak
shockwave that forms moves rapidly backward and forward over the surface of the wing producing
similar indications to those of the shock stall.

Tuck Under

The marked nose-down pitching tendency often experienced by aeroplanes accelerating from high
subsonic to transonic speeds is known as ‘tuck under’ and usually occurs at a relatively high Mach
number just above Mcrit. It can only happen when the speed is greater than the critical Mach number,
and is caused by any one of the following three reasons:

a. The position of the shockwaves on the upper and lower surfaces of the wing do not coincide
when flying at a speed between Mcrit and Mach 1. The lower shockwave is forward of the
shockwave on the upper surface, which causes the CP to move aft. This creates a nose-down
pitching moment. See Figure 15.1(b).

b. For swept-wing aeroplanes the shockwave forms on the thick inboard wing roots first, thus
decreasing the lift on this part of the wing. This causes the CP to move aft because of the
sweepback and produces a nose-down pitching moment.

c. Due to the shockwave, the loss of the airflow downwash over the horizontal stabiliser decreases
its downward balancing force, which results in a nose-down pitching moment.

The Shock Stall

In the transonic range, the ‘shock stall’ is the separation of the boundary layer from the upper surface
of the wing behind a shockwave. This causes the total lift generated to decrease. Unlike the normal
high-incidence angle stall, the ‘shock stall’ occurs at low angles of attack and at an unexpectedly high
IAS. For swept-wing aeroplanes because the wing-tips stall first the aeroplane has a nose-up pitch
tendency.

The angle of attack of the high-speed shock stall decreases as the Mach number increases and CLmax
decreases in a similar manner. Because the turbulent airflow behind a shockwave results in an
increased loss of lift, any large rapid increase of speed can induce a premature stall.

Clmax at high subsonic speeds is less than that theoretically possible with an incompressible flow and
occurs at a lower speed. The maximum velocity of the airflow over the upper surface of the wing
increases with increased angle of attack; which decreases the value ofMcrit. This fact has great
significance when considering manoeuvring at high speed because a shockwave is induced at a lower
Mach number than in level flight.

The Buffet Boundary

The normal high-incidence angle stalling speed for an aeroplane is specified as an IAS. When
operating at high altitude, in the extremely low ambient temperatures present at those altitudes, the
TAS at which a high-incidence angle stall, for a normal, clean aeroplane will occur is very high. By the
same reasoning the local speed of sound, Mach 1, in the temperatures experienced at high altitudes in
terms of TAS is low; as a result the value of the Mcrit TAS is also low.

The cruising TAS of most jet transport aeroplanes operating at high altitude is just above Mcrit. The
difference between the normal high-incidence angle stalling speed and Mcrit can be as little as 20 kt.

Page 22 of 26
The ‘buffet boundary’ is the speed range between the high-speed buffet and the low-speed buffet,
which decreases with increased altitude. The separation of the cruising speed from the buffet-onset
speed is the ‘buffet margin.’ There is a low-speed margin and a high-speed margin.

Coffin Corner

For any given mass and ‘g’ loading there is one altitude, which is temperature dependent, at which
the high-incidence stalling speed and the critical Mach number are equal, this is colloquially known as
‘coffin corner’ or ‘Q corner.’ It refers to the apex of a triangular shape at the top of the flight envelope
chart where the stalling speed and critical Mach number lines converge to a point.

The colloquial name alludes to the fact that at this altitude the pilot is flying on a knife-edge that
demands great skill. If the speed is reduced the aeroplane will stall and lose altitude and if the pilot
increases the speed of the aeroplane it will lose lift, because of the shockwave and flow separation
that will cause the aeroplane to violently pitch nose-down and lose altitude. In other words, ‘tuck
under’ occurs.

When flying at the ‘Q’ corner altitude, if an aeroplane turns, the inner wing slows down and could fall
below the low-speed stalling speed, whilst the outer wing speeds up and could exceedMcrit. If such is
the case then the aeroplane will have exceeded both limits at the same time. Alternatively, turbulence
at high altitude could result in the aeroplane exceeding the ‘g’ limitations.

SPEED INSTABILITY

The centre of pressure moves aft initially because of the aft movement of the shockwave and creates
a nose-down tendency; this usually occurs from M0.80 to M0.98. The nose-down pitch causes the
aeroplane to accelerate and this creates a further nose-down pitch, which further exacerbates the
situation and continues the acceleration. Thus, the speed is unstable.

The rearward movement of the CP in the transonic range continues as the aeroplane accelerates,
causing a marked increase to the longitudinal stability. If corrective action is not taken the CP
continues to move rearward beyond the transonic envelope making the situation extremely
dangerous. Manual recovery from this situation is extremely difficult.

The effect a nose-down pitch angle with increased speed has on the handling characteristics of an
aeroplane is such that at a particular Mach number the aeroplane’s speed becomes unstable. The
large amount of up deflection of the elevator that is required to maintain straight and level flight
decreases the manoeuvring capability of the aeroplane.

To prevent such a situation arising, every aeroplane designed to operate in the high subsonic or in
the transonic speed range is fitted with a Mach trimming device. This system automatically moves the
elevators to counteract the pitching moments expected at any given Mach number by increasing the
nose-up pitch. It is therefore very sensitive to the Mach number. Such a system maintains the
aeroplane’s longitudinal stability and is known as the Mach trimmer.

The Mach Trimmer

The function of a Mach trimmer is to minimize the adverse effects of rearward movement of the CP at
high Mach speeds. It compensates for the nose-down pitching moment experienced by aeroplanes in
the transonic speed range. To do this it corrects the change to stick-force stability of swept-wing

Page 23 of 26
aeroplanes when flying above a specific Mach number by decreasing the elevator (stabilizer) incidence
by an amount greater than that required for the longitudinal trim change.

In so doing the Mach trimmer ensures positive stability and corrects insufficient stick-force stability
at high Mach numbers. It adjusts the elevator according to the Mach number and prevents ‘tuck
under.’ In other words, it maintains the stick-force gradient required to sustain level flight as the
speed increases, which for aeroplanes fitted with power-operated controls is approximately 1 pound
for every 6 kt change of speed in level flight.

A Mach trimmer should be fitted to aeroplanes that demonstrate unconventional elevator stick-force
characteristics at transonic Mach numbers. In the event of this equipment failing, the speed is limited
to specified Mach number.

Lateral Instability

The formation of a shockwave does not occur at precisely the same location on both wings or even at
the same Mach number. Consequently, when the aeroplane attains Mcrit, because of the asymmetry
of the shockwave formation one wing will drop because it generates less lift than the other. This
phenomenon causes lateral instability and will occur between Mcrit and the local speed of sound.

Lateral design features incorporated to increase the lift of the downgoing wing in a sideslip, such as
dihedral and/or sweepback, cause a greater local acceleration of the airflow that increases the Mach
number of the air passing over that wing. This results in the drag on that wing increasing and creates
a shockwave or intensifies the shockwave if one already exists and exacerbates the situation.

SPECIAL NOTE: Information for T-tail aeroplane.

SWEEP BACK (PHAK chapter 4)

Most of the difficulties of transonic flight are associated with shock wave induced flow separation.
Therefore, any means of delaying or alleviating the shock induced separation improves aerodynamic
performance. One method is wing sweepback. Sweepback theory is based upon the concept that it is
only the component of the airflow perpendicular to the leading edge of the wing that affects pressure
distribution and formation of shock waves.

On a straight wing aircraft, the airflow strikes the wing leading edge at 90°, and its full impact
produces pressure and lift. A wing with sweepback is struck by the same airflow at an angle smaller
than 90°. This airflow on the swept wing has the effect of persuading the wing into believing that it is
flying slower than it really is; thus the formation of shock waves is delayed. Advantages of wing
sweep include an increase in critical Mach number, force divergence Mach number, and the Mach
number at which drag rises peaks. In other words, sweep delays the onset of compressibility effects.

The Mach number, which produces a sharp change in drag coefficient, is termed the “force
divergence” Mach number and, for most airfoils, usually exceeds the critical Mach number by 5 to 10
percent. At this speed, the airflow separation induced by shock wave formation can create significant
variations in the drag, lift, or pitching moment coefficients. In addition to the delay of the onset of
compressibility effects, sweepback reduces the magnitude in the changes of drag, lift or moment
coefficients. In other words, the use of sweepback “softens” the force divergence.

Page 24 of 26
A disadvantage of swept wings is that they tend to stall at the wingtips rather than at the wing roots.
This is because the boundary layer tends to flow spanwise toward the tips and to separate near the
leading edges. Because the tips of a swept wing are on the aft part of the wing (behind the CL), a
wingtip stall causes the CL to move forward on the wing, forcing the nose to rise further. The
tendency for tip stall is greatest when wing sweep and taper are combined.

The stall situation can be aggravated by a T-tail configuration, which affords little or no pre-stall
warning in the form of tail control surface buffet. [Figure 4-62] The T-tail, being above the wing wake
remains effective even after the wing has begun to stall, allowing the pilot to inadvertently drive the
wing into a deeper stall at a much greater AOA. If the horizontal tail surfaces then become buried in
the wing’s wake, the elevator may lose all effectiveness, making it impossible to reduce pitch attitude
and break the stall. In the pre-stall and immediate post-stall regimes, the lift/drag qualities of a swept
wing aircraft (specifically the enormous increase in drag at low speeds) can cause an increasingly
descending flightpath with no change in pitch attitude, further increasing the AOA. In this situation,
without reliable AOA information, a nose-down pitch attitude with an increasing airspeed is no
guarantee that recovery has been effected, and up-elevator movement at this stage may merely keep
the aircraft stalled.

Figure 4-62. T-tail stall

It is a characteristic of T-tail aircraft to pitch up viciously when stalled in extreme nose-high attitudes,
making recovery difficult or violent. The stick pusher inhibits this type of stall. At approximately one
knot above stall speed, pre-programmed stick forces automatically move the stick forward, preventing
the stall from developing. A G-limiter may also be incorporated into the system to prevent the pitch
down generated by the stick pusher from imposing excessive loads on the aircraft. A “stick shaker,”
on the other hand provides stall warning when the airspeed is five to seven percent above stall speed.

Page 25 of 26
Page 26 of 26

You might also like