You are on page 1of 14

C h a p t e r 7

Introduction to Aircraft
Stability and Control
Dava J. Newman and Amir R. Amir

7.1 | BASIC NOMENCLATURE


Airplane performance is governed by forces along and perpendicular to the flight
path. The translational motion of the airplane is a response to these forces. In con-
trast, airplane stability and control are governed by moments about the center of
gravity, and the rotational motion of the plane is a response to these moments.
Much of the information in this chapter is motivated by content in Anderson’s
Introduction to Flight [33], which is recommended for a more detailed discus-
sion of stability and control. This chapter introduces basic nomenclature, defines
airplane stability, derives the equations of motion for airplane stability, provides
an interactive animation of stability and control where use of the CD-ROM is
suggested, and concludes with an example to facilitate practice implementing
the governing equations.
Figure 7.1 shows a rectangular right-handed coordinate system attached to the
aircraft. The origin of the axes is at the aircraft’s center of gravity. The x axis is
along the fuselage, the y axis is along the wingspan, and the z axis points downward.
The aircraft’s translational motion is given by the velocity components U, V,
and W along the axes. Thus, the net velocity of the aircraft is the vector sum of
these three velocity components. The rotational motion is given by the angular
velocity components 1f , u, c 2 about the x, y, and z axes.
# # #
In summary, the nomenclature associated with rotational motion is as follows:
#
■ x axis: roll axis, L  rolling moment, f  rolling velocity.
#
■ y axis: pitch axis, M  pitching moment, u  pitching velocity.
#
■ z axis: yaw axis, N  yawing moment, c  yawing velocity.

147
148 CHAPTER 7 Introduction to Airplane Stability and Control

Figure 7.1 | Definition of the airplane’s axis system.


Pitch

y
V

CG

U
Roll W
Yaw

x N
L'
z

A classic airplane has three basic controls: ailerons, elevator, and rudder.
They are designed to change and control the moments about the roll, pitch, and
yaw axes. These control surfaces are flaplike surfaces that can be deflected back
and forth at the command of the pilot.
Figure 4.1 introduced and defined airplane control surfaces. The ailerons are
located at the trailing edge of the wing. Similarly, the elevator is located at the
trailing edge of the horizontal stabilizer, and the rudder is at the trailing edge of
the vertical stabilizer.
A downward deflection of a control surface will increase the lift, since this
makes the airfoil shape of the wing or tail “more bent downward” (in aeronauti-
cal jargon, it has a larger camber) and thus produces more lift. An increase or
decrease of the deflection will change the moment and thus will result in a rota-
tion about an axis.
■ Rolling. The ailerons control the roll or lateral motion and are therefore
often called the lateral controls.

Aileron down

Aileron up
Roll
L'

■ Pitching. The elevator controls pitch or the longitudinal motion and thus is
often called the longitudinal control.
SECTION 7.2 Airplane Stability 149

Elevator up
M

Pitch

■ Yawing. The rudder controls yaw or the directional motion and thus is
called the directional control.

Rudder deflected

Yaw N

7.2 | AIRPLANE STABILITY


There are two types of stability: static stability and dynamic stability.

7.2.1 Static Stability


Static stability can be visualized by a ball (or any object) on a surface. Initially
the ball is in equilibrium. The ball is then displaced from the equilibrium posi-
tion, and its initial behavior is observed.
■ Statically stable. If the forces and moments on the body caused by a
disturbance tend initially to return the body toward its equilibrium
position, the body is statically stable.

■ Statically unstable. If the forces and moments are such that the body
continues to move away from its equilibrium position after being
disturbed, the body is statically unstable.
150 CHAPTER 7 Introduction to Airplane Stability and Control

■ Neutrally stable. If the body is disturbed but the moments remain zero, the
body stays in equilibrium and is neutrally stable.

Of importance to us are only the first two cases; neutral stability occurs very
rarely. A very important point is that static stability deals only with the initial
tendency of a vehicle to return or diverge from equilibrium.

7.2.2 Dynamic Stability


Dynamic stability deals with the time history of the vehicle’s motion after it ini-
tially responds to its static stability.
Consider an airplane flying at an angle of attack (AOA) ae such that the
moments about the center of gravity (cg) are zero. The aircraft is therefore in equi-
librium at ae and is said to be trimmed, and ae is called the trim angle of attack.
Now imagine that a wind gust disturbs the airplane and changes its angle of
attack to some new value a. Hence, the plane was pitched through a displace-
ment a  ae. The plane’s behavior could be as shown in Figure 7.2.

Figure 7.2 | Dynamically stable behavior.

Initial Initial
Displacement

Displacement

disturbance disturbance

Time Time

In both situations the airplane eventually returns to its equilibrium position


after some time interval. In the first case the vehicle approaches monotonically
the equilibrium position (aperiodic behavior), while in the second case, it over-
shoots the equilibrium position (damped oscillation). A body is dynamically sta-
ble if, out of its own accord, it eventually returns to and remains at its equilib-
rium position over a period of time. It is important to note that static stability
does not imply dynamic stability, as Figure 7.3 shows. The plane is dynamically
unstable but still statically stable.
SECTION 7.3 Static Forces and Moments on an Aircraft 151

Figure 7.3 | Dynamically unstable behavior.

Displacement
Time

7.3 | STATIC FORCES AND MOMENTS


ON AN AIRCRAFT
7.3.1 Resulting Force on a Wing
There is an aerodynamic force created by the pressure (and shear stress1) distri-
bution over the wing surface. The resultant (net) force R can be resolved into two
components: the lift L (perpendicular to the relative wind v∞) and the drag D (in
the direction of the relative wind v∞).

7.3.2 Resulting Moment on a Wing


Consider just the pressure on the top surface of the wing. The net force due to
that pressure distribution, called F1, points downward and is acting through point
1 on the chord line. The pressure distribution on the bottom surface results in a
net force F2, pointing upward and acting through point 2 on the chord line. The
total aerodynamic force on the wing is of course a summation of F1 and F2. If
F2  F1, there is lift. Since the two forces do not act through the same point,
there will be a net moment on the wing. See Figure 7.4.

Figure 7.4 | The origin of the moment acting on an airfoil.


F1

1 2

F2

1. Pressure is the force perpendicular to the surface per unit area, while shear stress is the force
along the surface per unit area. Both have units of pascals (or newtons per meter squared).
152 CHAPTER 7 Introduction to Airplane Stability and Control

The magnitude of the moment depends upon the reference point about
which the moment is taken. If the moment is taken with respect to the leading
edge, it is denoted by MLE. For subsonic wings it is often customary to take the
moment about the quarter-chord point (i.e, the point that is a distance c/4 away
from the leading edge). This moment is denoted by Mc/4.
Both MLE and Mc/4 vary with the angle of attack. However, a special point
exists about which the moment essentially does not vary with a. This point is
called the aerodynamic center (ac). For that point,
Mac  constant 1independent of angle of attack2 [7.1]
The moment coefficient about the aerodynamic center is defined as
Mac
CM, ac  [7.2]
qq Sc
where q∞ is the dynamic pressure, S the wing area, and c the chord length. (Recall
the definition of the coefficient of lift and drag from Chapter 3, “Aerodynamics.”)
The value of CM, ac is zero for symmetric airfoils and varies from 0.02 to
0.3 or so for cambered airfoils.

7.3.3 Moment on an Aircraft


Having looked at a wing only, we can now consider a complete airplane, as
shown in Figure 7.5. In examining a whole aircraft, the pitching moment about
the center of gravity (center of mass) is of interest. The moment coefficient about
cg is defined analogous to the moment coefficient about the ac:
Mcg
CM,cg  [7.3]
qq Sc

Figure 7.5 | Contributions to the moment acting about the center of gravity.
Lwing
Ltail

Mac
T D

αe ac
W

Mcg = 0
SECTION 7.4 Attaining Airplane Longitudinal Static Stability 153

An airplane is in pitch equilibrium when the net moment about the center of
gravity is zero.
Mcg  CM, cg  0 airplane is trimmed [7.4]
Note that while drag plays an essential part in performance determination, its
role is small for stability and control. Its value is much less than that of the lift,
and its acts not too far from the center of gravity, so its effects are often neglected.

7.4 | ATTAINING AIRCRAFT


LONGITUDINAL STATIC STABILITY
Static stability and control about all three axes is necessary in the design of con-
ventional airplanes. However, a complete description of lateral, longitudinal, and
directional stability is difficult. We will focus on longitudinal motion (pitching
motion about the y axis), which is the most important.
Consider an airplane with fixed control surfaces. Wind tunnel testing may
reveal the following behavior (see Figure 7.6): The plot is almost linear and
shows the value of the CM, cg versus angle of attack a. The slope of the curve is
CM, cg/a, and is sometimes denoted with the letter “a.” (A partial derivative rather
than a total derivative is used since the coefficient does not depend on a alone.)
The value of CM, cg at an angle of attack equal to zero is denoted by CM, 0. The angle
at which the moment coefficient is zero is, of course, the trim angle of attack.

Figure 7.6 | The moment coefficient about the center of gravity as


a function of angle of attack for a longitudinally stable aircraft.

CM,CG
(+)
Trimmed
∂CM,cg
CM,0 Slope=
∂α

αe α
(–)

If the airplane is flying at its trim angle of attack ae and suddenly encoun-
ters a disturbance that causes it to pitch up or down (e.g., due to a wind gust),
the moment will be such that the plane will return to its equilibrium position. To
see that, imagine a wind gust pitching the plane up from ae to some larger a. By
looking at the plot in Fig. 7.6, you can see that the moment coefficient (and
hence the moment) will be negative, which makes the plane pitch down and return
to equilibrium.
154 CHAPTER 7 Introduction to Airplane Stability and Control

Suppose the curve of CM, cg versus a is as shown in Figure 7.7. The plane
would be unstable, as you can verify yourself. Thus, we can state the necessary
criteria for longitudinal static stability and balance as
0CM,cg
6 0 and CM,0 7 0 [7.5]
0a

Figure 7.7 | The moment coefficient about the center of gravity as


a function of angle of attack for a longitudinally unstable aircraft.
CM,cg

(+) Trimmed

αe α
(–)

CM,0

That is, the slope of the moment coefficient curve versus angle of attack has to be
negative, and the moment coefficient at zero angle of attack has to be positive.
An airplane can fly trough a range of angles of attack, but ae must be within
this range, or else the plane cannot be trimmed. If the aircraft can be trimmed, it
is said to be longitudinally balanced.
We can now answer the question, Why do airplanes have horizontal stabi-
lizers? If you have a wing by itself, it will usually have a negative CM, ac and thus
a negative CM, 0 (this is characteristic of all airfoils with positive camber). There-
fore a wing by itself is unbalanced. To correct the situation, a horizontal stabi-
lizer is mounted behind the wing. If the wing is inclined downward to produce
a negative lift, then a clockwise moment about the cg will be created. If this
clockwise moment is large enough, it will overcome the negative CM, 0 for the
wing-tail combination, making the aircraft as a whole longitudinally balanced.
The horizontal stabilizer does not have to be placed behind the wing. If it is
in front of the wing, it is called a canard configuration.

7.5 | USEFUL CALCULATIONS


AND AN EXAMPLE
First consider the tail (horizontal stabilizer) by itself, as shown in Fig. 7.8. Since
the tail is behind the wing, it feels two interference effects:
1. The airflow arriving at the tail does not have the same direction as that
arriving at the wing, since the wing deflects the airflow downward due to
SECTION 7.5 Useful Calculations and an Example 155

Figure 7.8 | Forces acting on the tail and relative wind seen by the tail [33].

the downwash effect (an effect due to the finite length of the wings).
Hence, the relative wind of the wing and tail makes up an angle .
2. Due to skin friction and pressure drag, the magnitude of the relative wind
seen by the tail is smaller than the magnitude of the relative wind seen by
the wing.
Now consider an idealized wing-tail configuration in steady, level flight
such as shown in Figure 7.9. The wing and the tail are set at incidence angles iw
and it, respectively, with respect to the longitudinal aircraft axis. The relative
wind V∞ comes in at an angle aw with respect to the wing. The relative wind V∞
comes in at an angle at with respect to the tail.

Figure 7.9 | Geometry of a wing-tail combination [39].


Lw

ε Lt
αt

Dw
Dt
iw cg it
αw αt
V'∞
V∞
xa

lt
156 CHAPTER 7 Introduction to Airplane Stability and Control

The tail angle of attack can be computed as follows:


a t  a w  e i t  iw [7.6]
where  is the downwash. Its value can be computed from the following equation:
e  e a aw [7.7]
where a  0.3 to 0.5. This allows us to rewrite the equation for at as
at  11  ea 2aw it  iw [7.8]
When the airplane is trimmed, the moments about the cg are zero. From this
we can find the trim condition. The coefficients of lift for the wing and for the
tail and the angle of attack can be defined as the product of the slope of the
moment coefficient.
CL,w  a wa w [7.9]

CL,t  a t 3a w 11  e a 2 i t  iw 4 [7.10]

After several steps and a few simplifying assumptions, the trim condition can be
CM,cg
 c  11  e a 2 d a w c 1i  i t 2 d  0
xa At l t a t CM,ac At lt a t

aw c Aw ca w a w A wca w w
[7.11]
The first term in brackets is the sensitivity to the angle of attack. Consider again
the situation where a wind gust disturbs a plane flying in trim and causes it to
pitch up. For the plane to be stable, the moment coefficient CM, cg (which was
0) has to be negative in order for the plane to pitch down. For stability we can
then write

11  ea 2
0CM,cg xa At lt at
6 0 or 6 [7.12]
0a c Aw caw
In designing your LTA vehicle, you can place the cg (and thus set the value
of xa) such that the vehicle is stable, using the above inequality. In the limiting
case, when the cg is as far back as possible,

a b  11  e a 2
xa At lt a t
[7.13]
c max A wca w

The cg is said to be at the neutral point.


The trim angle of attack can be written as
CM,ac>aw 3 A t lt a t > 1Aw ca w 2 4 1iw  it 2
1a w 2 trim 
1xa >c2 max  xa >c
[7.14]

For lift generation, aw  0, CM, ac  0, (usually), and iwit  0. Also, Lt may


have to be negative.
SECTION 7.5 Useful Calculations and an Example 157

An animation was created to help you visualize airplane pitch stability and
control. Run the animation to become familiar with the governing equations and
to see the results of moments on the airplane resulting in stability and instabil-
ity. Now that you have seen an airplane in motion, Example 7.1 provides addi- Pitch Stability and
tional practice in calculating airplane stability parameters. Control Movie.

EXAMPLE 7.1
This numeric example illustrates the use of the equations presented. Consider a light
aircraft with the following characteristics:
■ Aw  15 m2, c  1.6 m.
■ At  2.3 m2, lt  4.0 m.
■ m  1,050 kg, Iy  1,600 kg
m2.
■ aw  5 rad–1, at  rad1.
■ CM, ac  0.07, a  0.45.
■ V  50 m/s, r  1 kg/m3 (approximately 1,500 m altitude).
Question 1. Where is the neutral point of the aircraft?
The implied lift coefficient for level flight (neglecting the tail contribution) is
mg 11,0502 19.82
CL,w    0.549
10.5 2 112 1502 2 1152
[7.15]
1
rV 2Aw
2
The angle of attack of the wing is then
CL,w 0.546
aw    0.11 rad  6.3° [7.16]
aw 5
The neutral point can be calculated as follows:
12.3 2 14.0 2 142
a b 11  e a 2  11  0.452  0.264
xa At l t at

1152 11.62 152
[7.17]
c max Aw caw
1xa 2 max  0.42 m [7.18]
Question 2. Suppose that the cg is placed halfway between the ac and the neutral
point, that is, xa  0.21 m and xa/c  0.132. What is the angle of attack of the tail
and the lift produced by the tail?
First the difference between the wing and tail incidence angles needs to be
computed:
0.07>5 52.3142 142> 3 1152 11.62 152 4 61iw  i t 2
1aw 2 trim  0.11  [7.19]
0.264  0.132
1 iw  it  0.093 rad  5.3°
The tail angle of attack is then
a t  11  e a 2a w it  iw
 11  0.452 10.112  0.093 [7.20]
 0.0325 rad  1.9°
158 CHAPTER 7 Introduction to Airplane Stability and Control

The lift generated by the tail is


1
Lt  rV 2At at at
2
[7.21]
 a b 112 150 2 12.32 142 10.03252  374 N 1negative lift!2
1 2
2
Note that since this lift is negative (i.e, a downward force), it adds to the weight
of the vehicle as far as the required wing lift is concerned!

PROBLEMS
7.1 Design and make a paper airplane, focusing on performance and stability.
(a) Sketch a multiview drawing of your paper airplane. Necessary
viewpoints are the side view, top view, and frontal view.
(b) Describe your design.
(c) Make five trial flights and discuss your results. Bring your design to
the lecture ready to fly.
(d) Your in-class challenge will be to describe and then demonstrate your
chosen aspect of performance or stability (you are to demonstrate only
one stated objective). For example, use elevators to demonstrate pitch.
7.2 Sketch (in full scale) your rendition of an early Egyptian bird model
(legless, featherless, sycamore wood) that is actually an ancient model
airplane with an 18 cm wingspan and vertical stabilizer.
7.3 One definition of stability is the following:
Given a system in equilibrium [e.g., the position of a mechanical component is
not changing, or the current in an electric circuit is constant, or the temperature in a
furnace is constant, or the concentration of a species in a chemical reaction is con-
stant] if the system is subjected to a perturbation and returns to its initial equilib-
rium position after the perturbation has disappeared, the system is said to be stable.
Consider the pendulum represented here; it is in equilibrium in both positions
and can freely rotate around A.

Vertical

A
B
Problems 159

(a) Given the above definition of stability, which of the above systems is
in a stable equilibrium?
(b) Sketch the evolution of the angle u(t) between the segment AB and
the vertical as a function of time following a brief perturbation for
the two systems in the figures above. Justify your answer to part (a).
(c) Now consider the two pendulums immersed in a vat of honey!
Sketch u(t) for the two systems. How does it differ from your answer
to part (b)? Suggest a physical explanation for this difference.
7.4 A block diagram is a graphical representation of a system (a model or an
equation). The following conventions are used:

x K y

x is said to be the input, i.e., a quantity that is prescribed or imposed on


the system, and y the output. The above graph represents the equation y 
Kx (the input multiplies what is in the box).
(a) Consider a mass M on a frictionless surface. Apply a force F to the
mass (F is the input). Give a graphical representation of the
fundamental equation of dynamics.
(b) The following blocks are said to be in series or in cascade. What is
the relationship between x and z?

y
x z
K G

(c) The block here represents the equation t  x  y:

x t
+ –

y
160 CHAPTER 7 Introduction to Airplane Stability and Control

The following is the fundamental block diagram of a feedback control


system. Find the equation that relates x to y (the input to the output).

x y
K
+ –

You might also like