You are on page 1of 28

Learning Material

Immersed Model
Physics on: How to fly an Airplane

A. Introduction
Airplanes are constructed such that the airflow pattern around them generates lift, thereby
enabling them to fly. The airflow in turn is produced by the forward motion of the plane
relative to the air. This forward motion is produced by engine thrust, delivered by way of
propeller engines or air-breathing engines (turbines). Airplane engines produce thrust by
accelerating the airflow in the rearward direction. This backwards acceleration of the
airflow exerts a "push" force on the airplane in the opposite direction, by Newton's third
law, causing the airplane to move forward.

All the forces (and moments) acting on an airplane are a result of pressure forces (normal
to the airplane surfaces) and shear forces (along the airplane surfaces), both created by
the airflow pattern over the airplane body. The moments (torques) are a result of the
forces and, in addition to forces, are part of rigid body motion analysis. Airplanes can
generally be treated as rigid bodies when analyzing the dynamics of their motion.

The lift force "holding" a plane up is generated by airflow over the wings. This airflow is
only possible if the plane moves relative to the air, hence lift is only possible if the plane
moves relative to the air. And the relative air speed must be large enough for sufficient
lift to be generated.

B. Airfoils Shape
In aerodynamics, airplane wings are called airfoils. They have a cambered shape which
enables them to produce lift, even for angles of attack (α) equal to zero. The figure below
shows a cross-sectional view of an airfoil, with nomenclature shown.

The orientation of the airfoil relative to the airplane body is shown below. The angle of
incidence is defined as the angle between the chord line and the longitudinal axis of the
plane. For general aviation designs, an angle of incidence commonly used is about 6
degrees.
The presence of air friction (viscosity) is what allows an airfoil to generate lift. The
reasoning behind this is complicated and involves rather complex mathematics. But
basically, the air flow pattern around an airfoil results in the lower half of the airfoil
experiencing greater pressure force than the top half of the airfoil. As a result, a lift force
is generated. This is perhaps the least understood, and most asked, aspect of how
airplanes fly.
Figure of airfoil Shape:
Examples of airfoils in nature and within
various vehicles. Though not strictly an
airfoil, the dolphin flipper obeys the same
principles in a different fluid medium.
Streamlines around a NACA 0012 airfoil at
moderate angle of attack

Airfoil design is a major facet of aerodynamics. Various airfoils serve different flight regimes.
Asymmetric airfoils can generate lift at zero angle of attack, while a symmetric airfoil may better
suit frequent inverted flight as in an aerobatic airplane. In the region of the ailerons and near a
wingtip a symmetric airfoil can be used to increase the range of angles of attack to avoid spin–
stall. Thus a large range of angles can be used without boundary layer separation. Subsonic
airfoils have a round leading edge, which is naturally insensitive to the angle of attack. The cross
section is not strictly circular, however: the radius of curvature is increased before the wing
achieves maximum thickness to minimize the chance of boundary layer separation. This
elongates the wing and moves the point of maximum thickness back from the leading edge.
Lift and Drag curves for a typical airfoil
Airfoil Terminology:
The various terms related to airfoils are defined below:[6]

The suction surface (a.k.a. upper surface) is generally associated with higher velocity and lower
static pressure.
The pressure surface (a.k.a. lower surface) has a comparatively higher static pressure than the
suction surface. The pressure gradient between these two surfaces contributes to the lift force
generated for a given airfoil.
The geometry of the airfoil is described with a variety of terms:

The leading edge is the point at the front of the airfoil that has maximum curvature (minimum
radius).[7]
The trailing edge is defined similarly as the point of maximum curvature at the rear of the airfoil.
The chord line is the straight line connecting leading and trailing edges. The chord length, or
simply chord, c, is the length of the chord line. That is the reference dimension of the airfoil
section.
Different definitions of airfoil thickness

An airfoil designed for winglets (PSU 90-125WL)


The shape of the airfoil is defined using the following geometrical parameters:

The mean camber line or mean line is the locus of points midway between the upper and lower
surfaces. Its shape depends on the thickness distribution along the chord;
The thickness of an airfoil varies along the chord. It may be measured in either of two ways:
Thickness measured perpendicular to the camber line.[8][9] This is sometimes described as the
"American convention";[8]
Thickness measured perpendicular to the chord line.[10] This is sometimes described as the
"British convention".
Some important parameters to describe an airfoil's shape are its camber and its thickness. For
example, an airfoil of the NACA 4-digit series such as the NACA 2415 (to be read as 2 - 4 - 15)
describes an airfoil with a camber of 0.02 chord located at 0.40 chord, with 0.15 chord of
maximum thickness.

Finally, important concepts used to describe the airfoil's behavior when moving through a fluid
are:

The aerodynamic center, which is the chord-wise length about which the pitching moment is
independent of the lift coefficient and the angle of attack.
The center of pressure, which is the chord-wise location about which the pitching moment is
zero.
Thin airfoil theory:

Thin airfoil theory is a simple theory of airfoils that relates angle of attack to lift for
incompressible, inviscid flows. It was devised by German-American mathematician Max
Munk and further refined by British aerodynamicist Hermann Glauert and others[11] in the 1920s.
The theory idealizes the flow around an airfoil as two-dimensional flow around a thin airfoil. It
can be imagined as addressing an airfoil of zero thickness and infinite wingspan.
Thin airfoil theory was particularly notable in its day because it provided a sound theoretical
basis for the following important properties of airfoils in two-dimensional flow:[12][13]
(1) on a symmetric airfoil, the center of pressure and aerodynamic center lies exactly one quarter
of the chord behind the leading edge
(2) on a cambered airfoil, the aerodynamic center lies exactly one quarter of the chord behind the
leading edge
(3) the slope of the lift coefficient versus angle of attack line is   units per radian
As a consequence of (3), the section lift coefficient of a symmetric airfoil of infinite wingspan is:

where   is the section lift coefficient,


 is the angle of attack in radians, measured relative to the chord line.
(The above expression is also applicable to a cambered airfoil where   is the angle
of attack measured relative to the zero-lift line instead of the chord line.)
Also as a consequence of (3), the section lift coefficient of a cambered airfoil of
infinite wingspan is:

where   is the section lift coefficient when the angle of attack is zero.
Thin airfoil theory does not account for the stall of the airfoil, which usually
occurs at an angle of attack between 10° and 15° for typical airfoils.[14]

Derivation of thin airfoil theory:


The airfoil is modeled as a thin lifting mean-line (camber line). The mean-line, y(x), is
considered to produce a distribution of vorticity   along the line, s. By the Kutta condition,
the vorticity is zero at the trailing edge. Since the airfoil is thin, x (chord position) can be used
instead of s, and all angles can be approximated as small.
From the Biot–Savart law, this vorticity produces a flow field   where

 is the location where induced velocity is produced,   is the location of the vortex element
producing the velocity and   is the chord length of the airfoil.

Since there is no flow normal to the curved surface of the airfoil,   balances that from the
component of main flow  , which is locally normal to the plate – the main flow is locally
inclined to the plate by an angle  . That is:

This integral equation can by solved for  , after replacing x by

 ,

as a Fourier series in   with a modified lead term 


That is

(These terms are known as the Glauert integral).


The coefficients are given by

and

By the Kutta–Joukowski theorem, the total lift force F is proportional to

and its moment M about the leading edge to


The calculated Lift coefficient depends only on the first two terms of the Fourier series, as

The moment M about the leading edge depends only on   and   , as

The moment about the 1/4 chord point will thus be,

.
From this it follows that the center of pressure is aft of the 'quarter-chord' point 0.25 c, by

The aerodynamic center, AC, is at the quarter-chord point. The AC is where the pitching moment
M' does not vary with angle of attack, i.e.,

C. Forces Acting On An Airplane 

The figure below shows the resultant forces


acting on an airplane in level flight, moving at
constant velocity. 
Since the airplane is moving at constant
velocity it is experiencing zero acceleration, and the forces must balance. This means that the lift
force (L) generated by the airplane wings must equal the airplane weight (W), and the thrust
force (T) generated by the airplane engines must equal the drag force (D) caused by air
resistance.
An airplane undergoing takeoff, or landing, experiences similar forces acting on it. The figure
below shows the typical forces acting on an airplane during takeoff. Note that the lift force (L) is
defined as perpendicular to the velocity (V) of the plane relative to the air. The drag force (D) is
defined as parallel to the velocity (V). As one would expect, the thrust force (T) is in the same
direction as (V). The weight (W) of the plane points straight down in the direction of gravity.
Now, W = mg , where m is the mass of the plane and g is the acceleration due to gravity, where

g = 9.8 m/s2.

If the plane is moving at constant velocity with respect to ground then all the forces acting on the
plane must be balanced. This means that in the vertical direction the sum of the forces is equal to
zero, and in the horizontal direction the sum of the forces is equal to zero. Mathematically, in the
vertical direction: 
Lcosθ + Tsinθ - Dsinθ - W = 0.
In the horizontal direction: 
Tcosθ - Dcosθ - Lsinθ = 0.
If the plane is experiencing acceleration one can account for this in the force equations, by
including acceleration terms in the force equations, using Newton's second law. 
When an airplane flies, the wing is designed to provide enough Lift to overcome the airplane’s
Weight, while the engine provides enough Thrust to overcome Drag and move the airplane
forward. And the Thrust of a rocket engine overcomes the Weight of the object to move the
rocket forward.

Increasing the weight of an aircraft affects the amount of lift needed. In turn, a larger wing would
provide more lift, but that would increase the amount of drag and therefore increase the amount
of thrust needed. The forces of flight are interconnected, and a change in one affects the others.
Four forces affect things that fly:
The proportion between weight and thrust is determined by the airplane designer depending on
the anticipated missions. For example, if by design an airplane must be able to accelerate
vertically upwards then the thrust must be greater than the weight and drag combined. In small
aircraft the weight/thrust ratio is about. 10:1.
Maintaining a steady flight requires a balance, often described as an equilibrium of all the forces
acting upon an airplane. Weight, lift, thrust and drag are the acting forces on an airplane.
Assuming a straight and level flight, lift must be equal to weight and drag must be equal to
thrust. This is what happens if this equilibrium is violated:
 If lift becomes greater than weight, then the plane will accelerate upward.
 If the weight is greater than the lift, then the plane will accelerate downward.
 When the thrust becomes greater than the drag, the plane will accelerate forward.
 If drag becomes greater than the thrust a deceleration will occur
 Acceleration is best explained by using Newton's Second Law of Motion.
Here are 4 force on an airplane flights:
1. Thrust opposes drag. The engine creates thrust and moves the plane forward. (Gravity
provides the thrust for a glider.) The engines push air back with the same force that the
air moves the plane forward; this thrust force-pair is always equal and opposite according
to Newton's 3rd Law. When thrust is greater than drag, the plane accelerates according
to Newton's 2nd Law. When the plane flies level at constant velocity, thrust equals drag.
When the plane flies level at constant velocity, all opposite forces of flight are equal: drag
= thrust and weight = lift. How the 4 forces of flight interact

 
2. Drag opposes thrust. Imagine sticking your hand out the window of a moving car and
flying your hand. The force that pushes your hand back is called "drag". As your hand
pushes on the wind, the wind also pushes against your hand. Isaac Newton would say that
force of your hand pushing on the air is always equal to the force of the air pushing on
your hand; this is his third law. When the plane flies level at constant velocity, weight =
lift! When the engines of a plane quit, drag slows the plane down according to Newton's
2nd Law. How the 4 forces of flight interact
  Legs of birds and wheels of planes are tucked in to reduce drag. Drag is unwanted  
because it makes the plane or bird inefficient. Planes with more drag require more
thrust to fly successfully. To reduce drag and increase efficiency, planes are
streamlined. Planes also use camber and high aspect ratios to reduce drag.

3. Lift opposes weight. Newton's Laws and Bernoulli's Principle generate lift. A plane that
sits on a runway doesn't have any lift, but it does have weight. Lift is proportional to the
square of the velocity of an airplane and as a plane goes faster, its lift increases. As a
plane moves forward, its lift force increases until it equals its weight. When lift equals
weight, the plane can fly. In level flight, lift equals weight as the plane flies at constant
velocity. When a plane accelerates upward, lift is greater than weight. See Newton's 2nd
Law. In contrast, lift = weight when the plane flies level at constant speed. See Newton's
3rd Law.
If the lift coefficient for a wing at a specified angle of attack is known (or estimated using
a method such as thin airfoil theory), then the lift produced for specific flow conditions
can be determined using the following equation:

where
 L is lift force,
 ρ is air density,
 v is true airspeed,
 A is the wing area, and
 is the lift coefficient at the desired angle of attack, Mach number, and Reynolds number
4. Weight opposes lift. Weight and lift are equal when a plane flies level at constant
velocity. Because excess weight requires more lift, and therefore more thrust, heavy
planes are more difficult to get off the ground as compared to lighter planes. Planes with
less weight require less thrust. Thus, planes are designed to be as light as possible. 
A plane that sits on a runway has weight. The force from the ground balances the weight of the
plane and there is no change in motion.
When the total force on the plane is in one direction, the force is called “unbalanced”. An
unbalanced force changes the motion of the plane. For instance, when thrust is greater than drag,
it is the unbalanced force that causes the plane to speed up, or accelerate. In addition, as the
velocity of the plane increases, the lift force increases and becomes the unbalanced force that
causes the plane to fly.
If the plane exerts 3 tons of thrust, it will speed up until it experiences 3 tons of drag. During the
acceleration, passengers will feel as though they are pushed back into their seats. Eventually,
drag and thrust will be balanced and the velocity of the plane will be constant. In other words,
when these forces are balanced, there will be no change in motion.

In fact, when a plane flies level at constant velocity, the opposite forces of flight are equal. For
example, if the plane weighs 50 tons, the lift force also equals 50 tons. (Typically, lift and weight
are 10 to 40 times greater than drag and thrust.)
D. Maneuvering And Navigation
Airplanes control their navigation path and attitude (orientation relative to the direction of
air flow) by adjusting physical elements on the outside of the airplane, elements which
modify the airflow pattern around the plane, causing the plane to adjust its attitude and
flight path. These physical elements are called control surfaces and consist of ailerons,
elevators, rudders, spoilers, flaps, and slats. Adjusting a plane's flight path always
involves either pitching, rolling, or yawing, or a combination of these. The figure below
illustrates what these are. 

Source: Wikipedia via ZeroOne 

For example, let's say a plane is to go around a horizontal turn. For a plane to make a turn
its body orientation must be tilted (roll) such that the resulting aerodynamic forces enable
the plane to go around a turn. Lateral forces enabling a turn are only possible by tilting
the airplane such that the lift force (L) has a lateral component needed to balance the
centripetal acceleration produced during the turn. The figure below illustrates this. 
 

We can analyze this as follows. 

By Newton's second law, the force balance for the centripetal acceleration, in the lateral
direction, is given by 

where θ is the bank angle, m is the mass of the plane, V is the velocity of the plane
(normal to the page) with respect to ground, and R is the radius of the turn. 

The force balance in the vertical direction is given by 

Combine the above two equations to give the radius of the turn. We have 
 

Another example of a situation where flight path needs to be adjusted occurs if there is a
wind blowing in a direction different from the direction the pilot want to go, with respect
to ground. For example, let's say there is a lateral wind blowing, which results in the
plane being "carried" in the direction of the wind. And let's further say that the pilot want
to fly straight ahead. To compensate for the wind, the pilot must fly at a certain angle
relative to the air such that the plane's flight with respect to ground is in the desired
direction. The figure below illustrates this. 

where Φ is the angle with respect to the desired direction of travel, Vw is the wind
velocity with respect to ground, Vp is the plane velocity with respect to the air, and Vr is
the resultant plane velocity with respect to ground. 

Using vector addition we can construct the following vector diagram. The angle Φ and
the velocity Vp must be selected such that the resultant velocity vector Vr is in the desired
direction. 
 

On a related note, planes will often fly in high altitude jet streams. These are fast flowing
air currents. A plane can be "carried" by a jet stream and as a result dramatically reduce
flight time if the jet stream is flowing in the intended direction of travel. Of further
benefit, high altitude flight also reduces the drag force since air has much lower density at
high altitude. To understand this consider the drag force equation: 
D = (1/2)CρAv2,
where Cis the drag coefficient, ρ is the density of the air through which the airplane is
moving, A is the projected cross-sectional area of the airplane body perpendicular to the
flow direction (that is, perpendicular to v), and v is the speed of the airplane relative to
the air. As you can see, drag force is proportional to density, and lower drag force results
in greater fuel efficiency since less engine thrust is needed to overcome the drag force. 

E. Bernouli’s Principle

In some textbooks and general references, Bernoulli's Principle is the sole explanation of lift.
However, this explanation is incomplete. Newton's Laws must also be considered. For instance,
the effect of camber creates lower pressure and it forces air down. According to Newton's
3rd Law, the reaction of the air being pushed down is that the wing is pushed up.

Bernoulli's Principle, although never considered by the Wright brothers, helps to explain the
camber and efficiency.

  "Bernoulli's Principle" states that when the velocity of a fluid is  


faster, the pressure of the fluid is lower. In contrast, this means that
slower moving fluids have higher pressure.

Bernoulli's principle applies to flight in the following manner: Air passes faster over the top of a
cambered wing and therefore results in lower pressure. See the diagram below. As you can see,
the top of the wing is curved. The air that passes over the top of the wing moves faster because it
travels a greater distance in the same amount of time as compared to the air that passes under the
wing. Lift is created because the air under the wing is slower and exerts higher air pressure. The
differences in pressure create lift.

 
Here is the secret of lift: For an airplane to fly
there must be greater/stronger/higher air
aster Air = lower pressure pressure under a wing and
smaller/weaker/lower air pressure on top.
 
Bernoulli’s Principle, states that pressure
decreases when air moves faster. Air moves
faster over the top of a wing, which results in an
area of lower pressure. Meanwhile, the bottom
Slower air = higher pressure of the wing experiences higher
pressure. Camber increases the difference in
air pressure between the top and bottom
surfaces.

F. Newton’s Law Principle


It can be seen in the picture to the right that as the angle of attack is increased, as in the
bottom picture, the resultant force from the deflection of the air both above and below the
wing is also a major component to lift. As the air is deflected downward, as in the top
picture, it pushes on the wing in an equal and opposite direction. Newton’s Third Law in
conjunction with Bernoulli’s principle can be used to explain the physics behind lift that
allows an airplane to fly. Lift combined with drag, weight, and thrust; provide the 4
forces which need to be controlled to allow an airplane to maintain flight.

according to Newton’s 3rd Law, wings push on the air and the air pushes back with an
equal and opposite force. The result of tons of air being pushed down is that the plane is
pushed up! The camber of a wing and its angle of attack deflect air downward.
Those who advocate an approach to lift by Newton's laws appeal to the clear existance of a
strong downwash behind the wing of an aircraft in flight. The fact that the air is forced
downward clearly implies that there will be an upward force on the airfoil as a Newton's 3rd law
reaction force. From the conservation of momentum viewpoint, the air is given a downward
component of momentum behind the airfoil, and to conserve momentum, something must be
given an equal upward momentum. Those who prefer to discuss lift in these terms often invoke
the Kutta-Joukowski theorem for lift on a rotating cylinder. The lift on a spinning cylinder has
been clearly demonstrated, and its discussion includes a vortex in the circulating air. Many
discussions of airfoil lift invoke such a vortex in the effective circulation of air around the
moving airfoil. Conservation of angular momentum in the fluid requires an opposite circulation
in the air shed from the trailing edge of the wing, and such vortex motion has been observed.

Part of the fascination of an aerobatics display is that with loops and upside-down flight. If the
greater curvature on top of the wing and the Bernoulli effect are evoked to explain lift, how is
this possible? The illustrations below attempt to show that an increase in airstream velocity over
the top of the wing can be achieved with airfoil surface in the upright or inverted position. It
requires adjustment of the angle of attack, but as clearly demonstrated in almost every air show,
it can be done.
G. Air Viscousity

The natural question is "how does the wing divert the air down?" When a moving
fluid, such as air or water, comes into contact with a curved surface it will try to
follow that surface. To demonstrate this effect, hold a water glass horizontally under a
faucet such that a small stream of water just touches the side of the glass. Instead of
flowing straight down, the presence of the glass causes the water to wrap around the
glass as is shown in figure 8. This tendency of fluids to follow a curved surface is
known as the Coanda effect. From Newton’s first law we know that for the fluid to
bend there must be a force acting on it. From Newton’s third law we know that the
fluid must put an equal and opposite force on the object which caused the fluid to
bend.

Fig 8 Coanda effect.

Why should a fluid follow a curved surface? The answer is viscosity; the resistance to
flow which also gives the air a kind of "stickiness". Viscosity in air is very small but it
is enough for the air molecules to want to stick to the surface. At the surface the
relative velocity between the surface and the nearest air molecules is exactly zero.
(That is why one cannot hose the dust off of a car and why there is dust on the
backside of the fans in a wind tunnel.) Just above the surface the fluid has some small
velocity. The farther one goes from the surface the faster the fluid is moving until the
external velocity is reached (note that this occurs in less than an inch). Because the
fluid near the surface has a change in velocity, the fluid flow is bent towards the
surface. Unless the bend is too tight, the fluid will follow the surface. This volume of
air around the wing that appears to be partially stuck to the wing is called the
"boundary layer".
H. Lift as a function of angle of attack

There are many types of wing: conventional, symmetric, conventional in inverted


flight, the early biplane wings that looked like warped boards, and even the proverbial
"barn door". In all cases, the wing is forcing the air down, or more accurately pulling
air down from above. What each of these wings have in common is an angle of attack
with respect to the oncoming air. It is this angle of attack that is the primary parameter
in determining lift. The inverted wing can be explained by its angle of attack, despite
the apparent contradiction with the popular explanation involving the Bernoulli
principle. A pilot adjusts the angle of attack to adjust the lift for the speed and load.
The popular explanation of lift which focuses on the shape of the wing gives the pilot
only the speed to adjust.

To better understand the role of the angle of attack it is useful to introduce an


"effective" angle of attack, defined such that the angle of the wing to the oncoming air
that gives zero lift is defined to be zero degrees. If one then changes the angle of
attack both up and down one finds that the lift is proportional to the angle. Figure 9
shows the coefficient of lift (lift normalized for the size of the wing) for a typical wing
as a function of the effective angle of attack. A similar lift versus angle of attack
relationship is found for all wings, independent of their design. This is true for the
wing of a 747 or a barn door. The role of the angle of attack is more important than
the details of the airfoil’s shape in understanding lift.

Fig 9 Coefficient of lift versus the effective angle of attack.


Typically, the lift begins to decrease at an angle of attack of about 15 degrees. The
forces necessary to bend the air to such a steep angle are greater than the viscosity of
the air will support, and the air begins to separate from the wing. This separation of
the airflow from the top of the wing is a stall.

I. The wing as air "scoop"

We now would like to introduce a new mental image of a wing. One is used to
thinking of a wing as a thin blade that slices though the air and develops lift somewhat
by magic. The new image that we would like you to adopt is that of the wing as a
scoop diverting a certain amount of air from the horizontal to roughly the angle of
attack, as depicted in figure 10. The scoop can be pictured as an invisible structure put
on the wing at the factory. The length of the scoop is equal to the length of the wing
and the height is somewhat related to the chord length (distance from the leading edge
of the wing to the trailing edge). The amount of air intercepted by this scoop is
proportional to the speed of the plane and the density of the air, and nothing else.

Fig 10 The wing as a scoop.

As stated before, the lift of a wing is proportional to the amount of air diverted down
times the vertical velocity of that air. As a plane increases speed, the scoop diverted
more air. Since the load on the wing, which is the weight of the plane, does not
increase the vertical speed of the diverted air must be decreased proportionately. Thus,
the angle of attack is reduced to maintain a constant lift. When the plane goes higher,
the air becomes less dense so the scoop diverts less air for the same speed. Thus, to
compensate the angle of attack must be increased. The concepts of this section will be
used to understand lift in a way not possible with the popular explanation.

J. Lift power

When a plane passes overhead the formerly still air ends up with a downward
velocity. Thus, the air is left in motion after the plane leaves. The air has been given
energy. Power is energy, or work, per time. So, lift must require power. This power is
supplied by the airplane’s engine (or by gravity and thermals for a sailplane).
How much power will we need to fly? The power needed for lift is the work (energy)
per unit time and so is proportional to the amount of air diverted down times the
velocity squared of that diverted air. We have already stated that the lift of a wing is
proportional to the amount of air diverted down times the downward velocity of that
air. Thus, the power needed to lift the airplane is proportional to the load (or weight)
times the vertical velocity of the air. If the speed of the plane is doubled the amount of
air diverted down doubles. Thus the angle of attack must be reduced to give a vertical
velocity that is half the original to give the same lift. The power required for lift has
been cut in half. This shows that the power required for lift becomes less as the
airplane's speed increases. In fact, we have shown that this power to create lift is
proportional to one over the speed of the plane.

But, we all know that to go faster (in cruise) we must apply more power. So there
must be more to power than the power required for lift. The power associated with
lift, described above, is often called the "induced" power. Power is also needed to
overcome what is called "parasitic" drag, which is the drag associated with moving
the wheels, struts, antenna, etc. through the air. The energy the airplane imparts to an
air molecule on impact is proportional to the speed squared. The number of molecules
struck per time is proportional to the speed. Thus the parasitic power required to
overcome parasitic drag increases as the speed cubed.

Figure 11 shows the power curves for induced power, parasitic power, and total power
which is the sum of induced power and parasitic power. Again, the induced power
goes as one over the speed and the parasitic power goes as the speed cubed. At low
speed the power requirements of flight are dominated by the induced power. The
slower one flies the less air is diverted and thus the angle of attack must be increased
to maintain lift. Pilots practice flying on the "backside of the power curve" so that
they recognizes that the angle of attack and the power required to stay in the air at
very low speeds are considerable.
Fig 11 Power requirements versus speed.

At cruise, the power requirement is dominated by parasitic power. Since this goes as
the speed cubed an increase in engine size gives one a faster rate of climb but does
little to improve the cruise speed of the plane.

Since we now know how the power requirements vary with speed, we can understand
drag, which is a force. Drag is simply power divided by speed. Figure 12 shows the
induced, parasitic, and total drag as a function of speed. Here the induced drag varies
as one over speed squared and parasitic drag varies as the speed squared. Taking a
look at these curves one can deduce a few things about how airplanes are designed.
Slower airplanes, such as gliders, are designed to minimize induced drag (or induced
power), which dominates at lower speeds. Faster airplanes are more concerned with
parasite drag (or parasitic power).
Fig 12 Drag versus speed.

K. Wing efficiency

At cruise, a non-negligible amount of the drag of a modern wing is induced drag.


Parasitic drag, which dominates at cruise, of a Boeing 747 wing is only equivalent to
that of a 1/2-inch cable of the same length. One might ask what effects the efficiency
of a wing. We saw that the induced power of a wing is proportional to the vertical
velocity of the air. If the length of a wing were to be doubled, the size of our scoop
would also double, diverting twice as much air. So, for the same lift the vertical
velocity (and thus the angle of attack) would have to be halved. Since the induced
power is proportional to the vertical velocity of the air, it too is reduced by half. Thus,
the lifting efficiency of a wing is proportional to one over the length of the wing. The
longer the wing the less induced power required to produce the same lift, though this
is achieved with an increase in parasitic drag. Low speed airplanes are affected more
by induced drag than fast airplanes and so have longer wings. That is why sailplanes,
which fly at low speeds, have such long wings. High-speed fighters, on the other
hand, feel the effects of parasite drag more than our low speed trainers. Therefore, fast
airplanes have shorter wings to lower parasite drag.

There is a misconception by some that lift does not require power. This comes from
aeronautics in the study of the idealized theory of wing sections (airfoils). When
dealing with an airfoil, the picture is actually that of a wing with infinite span. Since
we have seen that the power necessary for lift is proportional to one over the length of
the wing, a wing of infinite span does not require power for lift. If lift did not require
power airplanes would have the same range full as they do empty, and helicopters
could hover at any altitude and load. Best of all, propellers (which are rotating wings)
would not require power to produce thrust. Unfortunately, we live in the real world
where both lift and propulsion require power.

L. Power and wing loading

Let us now consider the relationship between wing loading and power. Does it take
more power to fly more passengers and cargo? And, does loading affect stall speed?
At a constant speed, if the wing loading is increased the vertical velocity must be
increased to compensate. This is done by increasing the angle of attack. If the total
weight of the airplane were doubled (say, in a 2g turn) the vertical velocity of the air
is doubled to compensate for the increased wing loading. The induced power is
proportional to the load times the vertical velocity of the diverted air, which have both
doubled. Thus the induced power requirement has increased by a factor of four! The
same thing would be true if the airplane’s weight were doubled by adding more fuel,
etc.

One way to measure the total power is to look at the rate of fuel consumption. Figure
13 shows the fuel consumption versus gross weight for a large transport airplane
traveling at a constant speed (obtained from actual data). Since the speed is constant
the change in fuel consumption is due to the change in induced power. The data are
fitted by a constant (parasitic power) and a term that goes as the load squared. This
second term is just what was predicted in our Newtonian discussion of the effect of
load on induced power.

Fig 13 Fuel consumption versus load for a large transport airplane 


traveling at a constant speed.
The increase in the angle of attack with increased load has a downside other than just
the need for more power. As shown in figure 9 a wing will eventually stall when the
air can no longer follow the upper surface. That is, when the critical angle is reached.
Figure 14 shows the angle of attack as a function of airspeed for a fixed load and for a
2-g turn. The angle of attack at which the plane stalls is constant and is not a function
of wing loading. The stall speed increases as the square root of the load. Thus,
increasing the load in a 2-g turn increases the speed at which the wing will stall by
40%. An increase in altitude will further increase the angle of attack in a 2-g turn.
This is why pilots practice "accelerated stalls" which illustrates that an airplane can
stall at any speed. For any speed there is a load that will induce a stall.

Fig 14 Angle of attack versus speed 


for straight and level flight and for a 2-g turn.

M.Wing vortices

One might ask what the downwash from a wing looks like. The downwash comes off
the wing as a sheet and is related to the details on the load distribution on the wing.
Figure 15 shows, through condensation, the distribution of lift on an airplane during a
high-g maneuver. From the figure one can see that the distribution of load changes
from the root of the wing to the tip. Thus, the amount of air in the downwash must
also change along the wing. The wing near the root is "scooping" up much more air
than the tip. Since the root is diverting so much air the net effect is that the downwash
sheet will begin to curl outward around itself, just as the air bends around the top of
the wing because of the change in the velocity of the air. This is the wing vortex. The
tightness of the curling of the wing vortex is proportional to the rate of change in lift
along the wing. At the wing tip the lift must rapidly become zero causing the tightest
curl. This is the wing tip vortex and is just a small (though often most visible) part of
the wing vortex. Returning to figure 6 one can clearly see the development of the
wing vortices in the downwash as well as the wing tip vortices.

Fig 15 Condensation showing the distribution of lift along a wing. 


The wingtip vortices are also seen. 
(from Patterns in the Sky, J.F. Campbell and J.R. Chambers, NASA SP-514.)

Winglets (those small vertical extensions on the tips of some wings) are used to
improve the efficiency of the wing by increasing the effective length of the wing. The
lift of a normal wing must go to zero at the tip because the bottom and the top
communicate around the end. The winglets blocks this communication so the lift can
extend farther out on the wing. Since the efficiency of a wing increases with length,
this gives increased efficiency. One caveat is that winglet design is tricky and winglets
can actually be detrimental if not properly designed.

N. Ground effect

Another common phenomenon that is misunderstood is that of ground effect. That is


the increased efficiency of a wing when flying within a wing length of the ground. A
low-wing airplane will experience a reduction in drag by 50% just before it touches
down. There is a great deal of confusion about ground effect. Many pilots (and the
FAA VFR Exam-O-Gram No. 47) mistakenly believe that ground effect is the result
of air being compressed between the wing and the ground.

To understand ground effect it is necessary to have an understanding of upwash. For


the pressures involved in low speed flight, air is considered to be non-compressible.
When the air is accelerated over the top of the wing and down, it must be replaced. So
some air must shift around the wing (below and forward, and then up) to compensate,
similar to the flow of water around a canoe paddle when rowing. This is the cause of
upwash.
As stated earlier, upwash is accelerating air in the wrong direction for lift. Thus a
greater amount of downwash is necessary to compensate for the upwash as well as to
provide the necessary lift. Thus more work is done and more power required. Near the
ground the upwash is reduced because the ground inhibits the circulation of the air
under the wing. So less downwash is necessary to provide the lift. The angle of attack
is reduced and so is the induced power, making the wing more efficient.

Earlier, we estimated that a Cessna 172 flying at 110 knots must divert about 2.5
ton/sec to provide lift. In our calculations we neglected the upwash. From the
magnitude of ground effect, it is clear that the amount of air diverted is probably more
like 5 ton/sec.

Dari bapak Nanangs…

You might also like