Download as pdf or txt
Download as pdf or txt
You are on page 1of 74

© Aditya A.

Paranjape
Chapter 1
Introduction to Flight Mechanics
Review of Aerodynamics; Thrust
Aditya Paranjape

1 Perspective on Flight Mechanics

Figure 1: The big picture of aeronautics. Flight mechanics sits squarely at the intersection of aerodynamics
and dynamical systems theory.

Figure 1 gives a broad division of aeronautics. Notice the areas that fall in the intersection zones: flight
mechanics, aeroelasticity and vibrations. As Sinha and Ananthkrishnan note in their textbook 1 , flight
mechanics “is the point of confluence of other disciplines with aerospace engineering and the gateway to
aircraft design.” This becomes apparent when we consider that this is the only sub-discipline which deals
with the aircraft as a whole! 2

2 Review of Aerodynamics
We assume the reader has had a basic course on subsonic aerodynamics at the very least. We quickly review
the major concepts on aerodynamics required in the coming chapters.

2.1 Anatomy of a Fixed Wing Aircraft


Figure 2 shows the standard components of an aircraft. The functions of each component are as follows:
• Fuselage: carry passengers, payload and fuel
1 Sinhaand Ananthkrishnan, Elementary Flight Dynamics with an Introduction to Bifurcation and Continuation Methods,
CRC Press, 2013.
2 A useful allegory is that of an elephant and blind men. Flight mechanics is like a blind man sitting atop the elephant - he

can feel how the elephant moves but can’t tell just what makes it move.

1
Figure 2: The major components that make up a generic aircraft. Image source:
http://dduino.blogspot.ca/2012/08/blu-baby-42-rc-plane-build-log.html

• Wings: produce lift, carry fuel and provide a mounting for engines

• Empennage: provide stability (more on this in the latter chapters), house the primary power unit
The aircraft is a highly integrated and connected system, and it is getting more so. For example, although
it may appear that the engines can be designed independently of the airframe as long as they provide the
required value of thrust, this is not so any more. The aerodynamic properties of the wing around the engine
are affected by the location and the geometry of the engine inlet to the point where they have to be designed
together. The reader is encouraged to read the Wikipedia entry on Boeing 737-MAX and follow it up with
technical papers for an illustration. The wings, despite simplistic exteriors, are incredibly complex: they
contain fuel tanks, actuators for the flaps and spoilers, and a whole slew of sensors for detecting the load
distribution and structural deformation. All of these sensors provide real-time feedback to the flight control
system which enables it to maintain the aircraft to the highest possible degree of efficiency.
Figure 3 shows some unconventional configurations. The X-31, shown on the left, never made it past
the experimental stage. Notice two major differences: that it has canards instead of a horizontal tail, and it
has vanes at the engine nozzle for thrust-vectoring. These design changes were made specifically to achieve
improvements in the agility of the airframe. The B-2 bomber, on the other hand, is actively deployed by the
US Air Force. It is a flying wing configuration with no empennage and integrated engines. The configuration
was designed specifically to ensure that it is stealthy, but also achieves a high degree of aerodynamic efficiency.
Incidentally, both these configurations present significant challenges for control design.
Figure 4 shows the standard notation used for writing the equations of motion and for defining the
attitude of the aircraft. Of particular importance here is the orientation of the aircraft with respect to
the wind velocity vector. The orientation is defined via two angles: the sideslip β which measures how far
the velocity vector is oriented from the plane of symmetry of the aircraft, and the angle of attack α which
measures how far the nose has pitched up from the projection of the velocity vector on the plane of symmetry.
The flight speed is denoted conventionally by V∞ , but we drop the subscript ‘∞’ unless it is required. Other
symbols will be introduced in later chapters when they are required.

2.2 Forces on an Aircraft


Figure 5 shows the forces acting on aircraft. The aerodynamic forces are resolved along and perpendicular
to the wind velocity vector. The lift acts in the plane of symmetry perpendicular to the wind velocity vector.
It is produced primarily by the wings, while the horizontal tail typically reduces the net lift for reasons

2
(a) X-31 (b) B-2

Figure 3: Unconventional aircraft configurations: an aircraft where the horizontal stabilizer is moved ahead
of the wing and referred to as a canard, and an aircraft without a tail. Source: Wikipedia

δ
Longitudinal plane
q

α
yB 90 - β
p xB
Pitch r
β Roll
zB
Yaw
V

Figure 4: The standard flight mechanics notation.

which will become clear later in this book. The fuselage also produces a small amount of lift which can be,
however, safely ignored in a basic setting.
The drag acts along the velocity vector and is contributed mainly by the fuselage (skin friction) and the
wings (skin friction and induced drag). In addition, aircraft flying at transonic speeds experience wave drag
due to shockwaves produced locally over the aircraft.
The angle γ in Fig. 5 is called the flight path angle. Thrust typically acts at a fixed angle to the body
x axis. For our analysis, we will assume that α is small enough so that thrust can be assumed to act along
the velocity vector. This assumption, however, needs to be used with caution.
The lift and drag are written as
1 2
L= ρ∞ V∞ SCL
2
1 2
D = ρ∞ V ∞ SCD
2
where S is the wing reference area, and CL and CD are functions of
• Angle of attack α and sideslip β, with α being the most important of all factors. We will ignore the
sideslip.
ρ∞ V∞ c
• Reynolds number Re = , where c is the mean wing chord length and µ is the coefficient of
µ
viscosity of air. The Reynolds number affects the individual terms CL and CD , as shown in Fig. 6.

3
Figure 5: Forces on an aircraft.

Clearly, low Reynolds number regime is highly nonlinear and in general not suitable for fixed-wing
flight. This regime is highly conducive, however, for efficient flapping flight.
V∞
• Mach number M a = . We will not delve into the effects of Mach number in this course.
a
We will assume that the flow is at a sufficiently large Reynolds number and subsonic so that across the
entire range of flight speeds, CL and CD are functions of just α. Furthermore, we will assume that α is
sufficiently small so that the lift-α relationship is linear:

CL = CL0 + CLα α

where CL0 depends on the wing camber and aspect ratio AR , while CLα depends on the aspect ratio:

CLα = 1+2/A R
. Note that α0 , the angle of attack at which lift is zero, does not change and depends only on
the aerofoil cross-section. The coefficient of drag in the low-α regime can be expressed in terms of CL :
1
CD = CD0 + C2
πeAR L
where CD0 depends largely on the airfoil geometry (aside from Re), while e is the Oswald efficiency factor
(0 < e < 1).
The lift-to-drag ratio L/D = CL /CD is an important design assessment metric for an aircraft. Since both
CL and CD are functions of α, we will try to locate an α such that CL /CD is maximized. Let F = CL /CD .
Since CL is a linear function of α, maximizing F with respect to α is equivalent to maximizing F with
respect to CL
CL CL
F = =
CD CD0 + kCL2
∂F 1 2kCL2
= 0 =⇒ − 2 =0
∂CL CD CD
=⇒ CD = 2kCL2 or CD0 = kCL2

The L/D ratio is maximized when the angle of attack equals


p
CD0 /k − CL0
α(L/D)max =
CLα

Finally, we note that at high angles of attack, the CL − α relation is nonlinear so that we can no longer
write CL = CL0 + CLα α. A more accurate representation is of the form CL = k1 sin(kα α + k0 ), which is
consistent with Fig. 7.

4
(a) Lift-to-drag ratio (b) Lift

(c) Drag

Figure 6: Effect of Reynolds number on lift and drag. Source: Mueller, 1999 (first plot) and Sandia National
Labs Report 802114 (last two plots)

Figure 7: Lift curve across a ±180 degree angles of attack range. Source: Sandia National Labs Report
802114

3 Standard Atmosphere
The standard atmosphere is a set of relations between the pressure p, density ρ, temperature T and altitude
hG . In particular, it gives mean values of p, ρ and T as functions of hG (‘G’ denotes geometric; above the

5
mean sea level). The standard atmosphere is used for
• Air speed calculation: determining true air speed from measured air speed (recall how a Pitot tube
works).
• Altitude determination using static pressure
• Design: provide baseline values of temperature and density for designing air frames and propulsion
The atmosphere is split into layers of constant temperature lapse rate, as shown in Fig. 8. Pressure and
density are calculated from the governing equations.
Recall the universal gas law
p = ρRT
and the ydrostatic equation
dp = −ρg dhG
where hG = 0 denotes the mean sea level. These equations yield
dp g dhG
=−
p RT
These equations help derive p and ρ as functions of temperature and altitude. There is one caveat though: g
depends on altitude hG . To get around the difficulties posed by this dependence, we define the geopotential
altitude h via the equation
dp g0 dh
g0 dh = g dhG =⇒ =−
p RT
g Re2
where g0 is the gravitational constant at mean sea level (hG = 0). Since = , we get
g0 (Re + hG )2

dhG Re2 Re hG
dh = Re2 2
=⇒ h = R e − =
(Re + hG ) Re + hG Re + hG
The relationship between h and hG is nonlinear, and furthermore hG = 0 =⇒ h = 0 and h → Re as
hG → ∞.

Figure 8: Atmospheric properties as functions of altitude. Image source: noaa.gov

The temperature lapse rates are defined with respect to h, so that dT = aT dh. If aT 6= 0 and if the
temperature at the base of the layer, altitude h1 , is known as T1 , then

T (h) = T1 + aT (h − h1 )

6
This gives the following equation for the layer with base at h = h1 :
dp g0 dh
=−
p R(T1 − aT h1 ) + RaT h

whose solution is given by


 −g0 /RaT  −g0 /RaT
T1 + aT (h − h1 ) T
p(h) = p(h1 ) = p(h1 )
T1 T1

These formulae are useful in the lowest layer of the atmosphere, namely the toposphere, where the temper-
ature lapse rate is −6.5 deg/km (degrees Celcius or Kelvin). If aT = 0, then dp/p = −g0 dh/RT1 and we
get  
p = p(h1 )exp −g0 (h − h1 )/RT1

The universal gas law can now be used to calculate the density ρ(h) from p(h) and T (h). These are tabulated
in the International Standard Atmosphere (ISA) tables which are available quite readily.
The standard atmosphere finds application in two important areas: determining the altitude of an aircraft
from pressure measurements, and determining the air speed accurately from Pitot tube measurements.

3.1 Determination of altitude


The standard atmosphere gives a staight-forward correspondence between p and h, while hG can be found
using the following expression:
Re hG Re h
h= =⇒ hG =
Re + hG Re − h
To be precise, the altitude calculated this way is called the pressure altitude which equals the altitude only
under ISA conditions: 1 atm pressure and 298 K temperature at sea level. Under non-standard conditions,
the ground values of pressure and temperature are adjusted to compute the altitude.

3.2 Air Speed Measurement


The pitot tube measures the air speed of an aircraft directly. This is called the Indicated Air Speed (IAS).
The IAS is calculated using Bernoulli’s equation:
s
2(p0 − p)
V∞ =
ρ

where p0 is the total pressure and p is the static air pressure. In traditional aircraft instruments, p and
ρ, which are altitude-dependent, are set to their static (M a = 0) sea-level values. Therefore, the ISA
is generally far from the actual speed of the the aircraft which can be calculated by making systematic,
step-wise corrections. Each step yields a refined value of the air-speed (see Fig. 9:

• The calibrated air speed (CAS) is obtained from the IAS by correcting for position errors, instrument
errors, etc., using charts supplied by the manufacturer.
• The density ρ is corrected by accounting for compressibility effects at Mach numbers outside the
low-subsonic range. This gives the equivalent air speed (EAS):
s 
2(p0 − p)
r
ρsea,Ma=0
VEAS = = VCAS
ρsea ρsea

7
• The altitude is corrected to obtain the true air speed
s 
2(p0 − p)
r
ρsea
VTAS = = VEAS
ρ ρ

• The ground speed is found by subtracting the speed of the head-wind Vground = VTAS − Vhead wind

Figure 9: The four speeds.

4 Propulsion
There are four typical propulsion mechanisms used in aircraft, depending on the flight regime:
• Jet engines: most commonly used device, up to mid-supersonic speeds (M a < 3)
• Propeller and engine: moderate subsonic speeds and low altitudes
• Ramjets and Scramjets: high supersonic and hypersonic speeds; used in missiles (BrahMos)
• Rockets: used in a few cases for take off (Me262c), and in some others for braking assistance while
landing (YMC-130 Hercules)
Figure 10 is a schematic diagram of a turbojet. The thrust produced by a turbojet is given by

Figure 10: Schematic diagram of a turbojet. Source:Wikipedia

T = (ṁair + ṁf )Ve − ṁair V∞ + (pe − p∞ )Ae

where e denotes the exit of the nozzle; Ve is the exhaust velocity of the air, pe is the exit pressure, and Ae is
the nozzle exit area. The mass flow rate is given by ṁair = ρ∞ Ain V∞ . The thrust is generally approximated
by
T ≈ ṁair (Ve − V∞ ) ≈ ρ∞ Ain V∞ Ve

8
Clearly, thrust reduces with increasing altitude since ρ∞ decreases with increasing altitude. Likewise, thrust
decreases when the ambient temperature increases. However, jet engine thrust is unaffected by flight speed
in the subsonic regime, i.e., until M a = 1. The latter property will be used in the next chapter when we
analyse level flight.
The derivation of an expression for the thrust produced by a propeller is beyond the scope of this course.
Figure 11 shows the forces on a propeller cross-section: the net force is found by integrating these across the
length of all blades.

Figure 11: Forces on a spinning propeller, at an arbitrarily chosen cross-section. Source:McCormick

The thrust produced by a propeller is strongly dependent on the advance ratio J = VnD ∞
, where n is
the propeller rotational rate, and D is the diameter. This has been illustrated in Fig. 12. Clearly, propeller
thrust drops as a function of the flight speed due to the increasing advance ratio, unlike a turbojet. Propeller
thrust also drops with decreasing air density (increasing altitude).

Figure 12: Nondimensional thrust as a function of the advance ratio for a typical propeller. Source: Mc-
Cormick

9
Chapter 2
Straight and Level Flight
Aditya Paranjape

Topics covered: lift and drag in level flight; flight speed and altitude envelope;
range and endurance.

1 Basics

Figure 1: Forces on an aircraft in wings-level flight.

Figure 1 shows the forces on an aircraft: the thrust acts along the longitudinal axis of the body;
the lift acts perpendicular to the velocity vector in the plane of symmetry, while the drag acts along
the velocity vector. The equations of motion can be written in terms of the speed V and the flight
path angle γ as:
mV̇ = T cos α − D − mg sin γ (Linear)
mV γ̇ = T sin α + L − mg cos γ (Centripetal)
We are interested in equilibrium conditions (also called trims or trimmed flight conditions) found
by setting V̇ = γ̇ = 0:
T cos α = D + mg sin γ (Linear)
T sin α + L = mg cos γ (Centripetal) (1)
We assume that α is small enough so that T cos α ≈ T and T sin α << L. This gives us the
equations of motion of straight and wings-level flight:

T = D + mg sin γ = D + W sin γ
L = mg cos γ = W cos γ

1
In this chapter, we focus on the equilibrium equations for straight and level flight which are obtained
by setting γ = 0 and correspond to flight at constant speed and altitude:

1
T = D = ρV 2 SCD (α)
2
1
L = W = ρV 2 SCL (α)
2

Straight and level flight is the most common mode of flight and is employed during

• Regular cruising flight

• Holding pattern and loiter

• Step-descent while landing in bad weather

2 Analysis of the Lift Equation


Consider the lift equation
1
L = W = ρV 2 SCL (α)
2
For a given aircraft, W and S are constant, while ρ is constant at a given altitude. Isolating the
constant terms on one side, we get
2W 2W
V 2 CL = = = constant (2)
ρS ρ S

The ratio W/S is called the wing loading, and it is one of the most important design parameters
for an aircraft. This relation helps us determine

• The value of α required to fly straight and level at a given speed.

• Alternately, the flight speed which will be achieved for a chosen value of α.

Figure 2 plots the flight speed as a function of α for several values of wing-loading. For a given
aircraft, there is a maximum attainable value of CL which is represented by CLmax . Recall that
the corresponding angle of attack is referred to as the stalling angle of attack. The stall speed of
an aircraft is given by s
2 W
Vstall =
ρCLmax S
This is an important value of the flight speed since, for all practical reasons, controlled flight at
sub-stall speeds is impossible. A slew of dangerous flight conditions arise out of stall, such as spin,
auto-rotation and deep-stall. Consequently, aircraft generally fly at speeds greater than or equal
to Vmin ≈ 1.2 Vstall .
From Figure 2, notice that the stall speed increases as the wing-loading W/S increases. Aircraft
use flaps to increase the wing area during take-off and landing; this allows them to achieve a higher

2
Figure 2: The relationship between flight speed and CL during straight and level flight for an Airbus
A330 at sea level.

wing loading for cruise flight at higher speeds. The cruise speed increases further at high altitudes
due to a reduction in the air density ρ (see Eq. (2)). As altitude increases, the stall speed increases

as well. Notice that Vstall ∝ 1/ ρ and has been illustrated in Fig. 3. Interestingly, the ratio between
the stall speed at a given altitude and the stall speed at sea-level is independent of the aircraft
geometry and weight, since r
Vstall (ρ) ρsea
=
Vstall (sea) ρ
The stall-speed doubles around an altitude of 12000 m, which is close to the cruising altitude of
most commerical jet aircraft.

Figure 3: Increase in the stall speed with altitude. These numbers are obtained for the Airbus
A330.

3
3 Analysis of the Drag Equation
Consider the equation for thrust-drag balance:
1
T = D = ρV 2 SCD , CD = CD0 + kCL2
2
2W
First, we observe that since CL = , the coefficient of drag during level flight is given by
ρV 2 S
 2
2W 1
CD = CD0 + kCL2 = CD0 + k
ρS V4

The thrust required for level flight at speed V is equal to the drag and is given by

2W 2
   
1 2 1
TR = D = ρSCD0 V + k (3)
2 ρS V2
| {z } | {z }
skin friction induced drag

The above equation tells us that while skin friction drag increases with flight speed as expected,
the induced drag reduces with increasing flight speed due to the reduction in CL . The drag CD

Figure 4: The drag as a function of the flight speed for the Airbus A330.

has been plotted in Fig. 4 (again, for the Airbus A330) as a function of the flight speed. The drag
is minimum at a speed where the skin friction drag is exactly equal to the induced drag. This
observation can be confirmed theoretically from Eq. (3) by using ∂D/∂V = 0.
Since
2W 2 1
 
CD = CD0 + kCL2 = CD0 + k
ρS V4
The condition for minimum drag yields
 2  1/4 r
4 k 2W 4k W 1
VDmin = =⇒ VDmin = 2
, k= (4)
CD 0 ρS ρ CD0 S πeAR

4
The minimum-drag CL is given by
r
Dmin CD0
= πeAR CD0 =⇒ αDmin = CL−1 ( πeAR CD0 )
p p
CL = (5)
k
where αDmin is the angle of attack at which the aircraft should fly in order to ensure that the
level-flight drag is minimized.
We make the following observations for the minimum drag speed.
1. The angle of attack for minimum drag depends only on the shape of the wing planform, the
aspect ratio of the wing and on CD0 . It is notably independent of the weight of the aircraft
and the wing area.

2. The minimum drag speed increases with increasing wing loading. This is a consequence of
CLDmin being independent of the wing loading.

3. The minimum-drag speed increases with altitude. This is, yet again, a consequence of Eq. (5).

4. The speed for minimum drag reduces with increasing CD0 . Therefore, the “smoother” the
aircraft configuration, the faster it can fly with given engines (and fuel, but that will be taken
up later in this chapter).
The minimum drag is given by
r
CDmin p CD 0
Dmin = W × = 2W kCD0 = 2W (6)
CLDmin πeAR

We make the following observations


1. The value of minimum drag is independent of altitude

2. The minimum drag reduces as the aspect ratio AR increases and as the wings become more
elliptical (e → 1).

4 Thrust and Power


The thrust required for sustained level flight equals drag, so that

1 CD CD0 + kCL2
TR = ρV 2 SCD = W × =W ×
2 CL CL
Let TA denote the maximum available thrust. There are two level flight conditions at which all
available thrust is required. The value of CL at these two conditions is found by solving the
quadratic equation  
2 TA
kCL + CL + CD0 = 0
W
The above equation has two solutions; let us denote them as CLS ≤ CLB . We make the following
observations.
q
1. The fastest speed at which an aircraft can fly is given by 2W/ρSCLS .

5
q
2. The slowest speed at which the aircraft can fly is given by max(Vstall , 2W/ρSCLB ).

3. When CLS = CLB , we infer that the aircraft is flying at the maximum flyable altitude.

For a jet engine, the maximum thrust TA depends on the altitude and temperature, but is inde-
pendent of the flight speed. However, the minimum drag is independent of altitude. Consequently,
as altitude increases, the maximum available thrust reduces, but the drag does not decrease. The
situation is illustrated in Fig. 5. The maximum flyable altitude can thus also be found by solving
TA (ρ) = Dmin . In Chapter 1, we pointed out that
q the maximum available thrust of a jet engine at
ρ
a given altitude obeys the law TA (ρ) = TA (sea) ρsea . Thus, the maximum flyable altitude of a jet
aircraft is found from the ISA tables for the density
 2
Dmin
ρ = ρsea
TA (sea)

Everything else being equal, notice that ρ(max altitude) ∝ W 2 ; this forces a heavy aircraft to fly
low and slow. We will show in the last part of this chapter how the range and endurance of an
aircraft are affected because of this.

Figure 5: Computing the maximum flight speed for jet and propeller-powered aircraft.

For a propeller-powered aircraft, thrust reduces with altitude as well as speed. However, the
shaft power PA does not change with flight speed, but reduces monotonically with altitude. An
analytical calculation of the maximum flight speed and altitude of a propeller-powered aircraft
require that we work with power rather than thrust.
The power required for straight and level flight at speed V is defined as PR = TR V , where TR
is the thrust required to fly straight q
and level at V . To calculate the minimum power required for
2W
level flight, we start by writing V = ρSC L
and note that TR = W (CD /CL ). This gives
s
2W 3 CD
PR = × 3/2
ρS CL
 3/2

CL
Clearly, the power required is minimized at the speed corresponding to CD .
max

6
A direct way to calculate the flight speed for minimum power, VP min , is as follows. Recall that

2W 2
   
1 2 1
TR = ρSCD0 V + k
2 ρS V2

Hence
2W 2 1
   
1 3
PR = ρSCD0 V + k
2 ρS V
Condition for minimum power:

2W 2
   
∂PR 1 2 1
= 0 =⇒ 3 ρSCD0 VP min = k 2 (7)
∂V 2 ρS VP min

Thus,
 1/4  1/4
1 4kW 2 VDmin VDmin
VP min = 2 2
= 1/4 = = 0.76VDmin
3 ρ S CD0 3 1.3161
Recall that the power available in a propeller aircraft (PA (ρ)) is constant at a given altitude. The
maximum flight speed is found by solving

2W 2 1
   
1
PA (ρ) = ρSCD0 V 3 + k
2 ρS V

This equation has two real positive roots; the larger real root yields the maximum flight speed at
that given altitude.
A commonly employed quantity in flight mechanic studies is the so-called specific excess power
(SEP) which is defined as
PA − DV
SEP =
m
The SEP depends on the flight speed and altitude, and measures the ability of the aircraft to
change its energy (more on this in Lecture 3). The maximum flight speed at a given altitude and
the maximum flyable altitude are both found by solving for SEP = 0. This is true for any aircraft,
but is particularly convenient for propeller-powered aircraft.

5 Interregnum
Before proceeding, we need to find the value of CL and the corresponding level flight at which the
1/2
ratio CL /CD is maximized
1/2 1/2
CL CL
Let F = CD = 2
CD0 +kCL
Condition for maximum F :
−1/2 3/2
∂F C 2kCL
= 0 =⇒ L − 2 =0
∂CL 2CD CD
=⇒ CD = 4kCL2
=⇒ CD0 = 3kCL2

7
Figure 6: Plot showing the three speeds: minimum power, drag, and (T /V ).

Thus, skin friction drag is thrice as large as the induced drag. Corresponding speed:

12kW 2
VR4 = 4
= 3VDmin
ρ 2 S 2 CD 0

=⇒ VR = 31/4 VDmin ≈ 1.316VDmin

TR
Consider the ratio F = V . Then,

2kW 2
 
TR 1 1
F = = ρSV CD0 +
V 2 ρS V3

Minimize F :
∂F 1 6kW 2
= 0 =⇒ ρSCD0 =
∂V 2 ρSVR4
which yields the expression we obtained on the previous slide:

12kW 2
VR4 =
ρ2 S 2 CD0
 
T
Thus, the speed VR yields
V min

5.1 The Three Speeds


Figure 6 shows the three speeds graphically in a drag versus speed plot. The speed for minimum
power does not have any associated straight line; rather, it occurs where a hyperbola of the form
DV = constant is tangential to the drag curve. The three speeds have been tabulated in Table 5.1.

8
Table 1: The three speeds

Condition on CL , CD Flight Speed Physical significance

3/2
CL
maxCL CD VP min = 0.76 VDmin Minimum power

CL
maxCL CD VDmin Minimum drag (thrust required)

1/2
CL
maxCL CD VR = 1.316 VDmin Minimum (T /V ) ratio

6 Range and Endurance


• Definition: the range of an aircraft is the maximum distance that it can fly with the given
amount of fuel.

• Definition: the endurance of an aircraft is the maximum time for which the aircraft can fly
with the given amount of fuel. The endurance is optimized during surveillance missions and
while holding in traffic prior to landing.

In order to calculate the range and endurance, we must quantify the fuel consumption of the
engines. This is done using two quantities:
|ṁf |g
• Thrust-specific fuel consumption: TSFC = T
|ṁf |g
• Power-specific fuel consumption: PSFC = P

TSFC and PSFC may additionally depend on the thrust/power settings, but we will assume these
to be constant. TSFC/PSFC are fundamentally engine parameters: among other things, they
include the propulsive efficiency of the engine. The reason why jet engines are characterised by
TSFC, and turboprops by PSFC is apparent enough.

6.1 Endurance of a Jet Aircraft


The endurance is given by Z tf
tF = dt
0
The rate of fuel consumption can be linked to the rate of change of weight of the aircraft:
1 dW
ṁf = −
g dt

9
so that
−dW
dt =
ṁf g
Therefore, Z tF Z W0
dW
tF = dt =
0 Wf ṁf g
where W0 is the initial weight of the aircraft, and Wf is the final weight, after fuel has been burnt.
This formula for endurance is identical for all aircraft; the difference between jet and propeller
aircraft arises in how we model ṁf .
Let cT denote the TSFC of the jet engine, so that

ṁf g = cT T
W
Also, since T = CL /CD ,
Z W0
1 CL dW
tF =
Wf cT CD W
If we choose α so that CL /CD ratio is maximum, then
   
1 CL W0
tF = ln
cT CD max Wf

Note that the endurance is independent of the altitude and the geometric parameters of the aircraft!

6.2 Endurance of Propeller-Powered Aircraft


Recall that the endurance is given by
Z tF Z W0
dW
tF = dt =
0 Wf ṁf g

Let cP denote the PSFC of the engine, so that

ṁf g = cP P

W 3/2
2
Also, note that P = DV = 3/2 √ .
C /CD ρS
L
3/2
We will choose an angle of attack to maximize the ratio CL /CD , so that

3/2
!
1 CL p 1 1
tF = 2ρS p −√
cP CD Wf W0

Clearly, the endurance is higher (everything else being equal) when the aircraft flies at a lower
altitude and has a large wing area (or a low wing loading).

10
6.3 Range of a Propeller-Powered Aircraft
The range is given by Z τ Z Wf
dW
R= V dt = V
0 W0 ṁf g
Yet again, this formula is the same for all aircraft; we will specialize it for jet aircraft and propeller
aircraft.
For a propeller-powered aircraft, since ṁf g = cP P = cP DV , we get
Z Wf Z Wf
dW 1 CL dW
R= =
W0 cP D W0 cP CD W

This gives
 
1 CL W0
R= ln
cP CD Wf

This formula is very much like that obtained for endurance of a jet aircraft: it is independent of the
altitude and the geometric parameters of the aircraft. Note, however, that the maximum altitude
is limited by thrust availability.

6.4 Range of a Jet Aircraft


Recall that Z Wf
dW
R= V
W0 ṁf g
Since ṁf g = cT T , we get
Z Wf
dW
R=
W0 cT (T /V )
r
CD 2W
Again, T = W and V = , so that
CL ρSCL
Z Wf
√ 1/2
2 CL dW
R= √
W0 cT ρS CD W 1/2
√ 1/2
2 2 C L p p 
=⇒ R = √ W0 − Wf
cT ρS CD

We observe that range improves at higher altitudes (lower ρ) and when S is small (i.e., the wing
loading is high).
Table 6.4 summarizes the observations made for the range and endurance of aircraft.

11
Table 2: Summary of Range and Endurance

Parameter Jet Aircraft Propeller Aircraft

3/2
Optimum α (Endurance) Max CL /CD Max CL /CD
1/2
Optimum α (Range) Max CL /CD Max CL /CD

Optimum V (Endurance) VDmin 0.76 VDmin


Optimum V (Range) 1.316 VDmin VDmin

Altitude dependence (Endurance) Insignificant Decreases


Altitude dependence (Range) Increases Insignificant

Wing loading dependence (Endurance) Insignificant Low wing loading


Wing loading dependence (Range) High wing loading Insignificant

6.5 Example
Consider a modern jet aircraft such as the Boeing 777 or the Airbus A330. The fuel to wright ratio
is typically 0.5, so that W0 /Wf ≈ 2. The maximum value of CL /CD is around 20. The aircraft
seek to fly at speeds 30% higher than those for minimum drag in order to maximise their range.
However, the stall speed of the aircraft and the available thrust require that they start their cruise
at relatively low altitudes under 30, 000 ft when fully loaded. As the aircraft burns fuel, it climbs
up in steps of 2000 ft to comply with the flight rules. Towards the end of their cruise, they are
typically around 38, 000 − 40, 000 ft.

12
Chapter 3
Climbing Flight, Level Turns, Take off and Landing
Aditya Paranjape

1 Gliding Flight
In the absence of thrust, the equilibrium equations of motion are given by

L = W cos γ, D + W sin γ = 0

This gives the following equation for the flight path angle:
CD 1 1
tan γ = − or tan(−γ) = =⇒ tan(|γ|) =
CL CL /CD CL /CD

Note that γ < 0, since the aircraft descends; furthermore,


loss in altitude
tan(|γ|) =
forward distance covered
What does this mean: for the shallowest glide, choose an angle of attack to maximize the lift-to-drag ratio
Recall that CL /CD is maximum when
1
CD0 = kCL2 , k =
πeAR
Thus, the flight speed is given by
s
2 2W 2W k
V = =⇒ V =
ρSCL ρ S CD0

The lowest flight path angle is given by


1 p
tan(|γ|min ) = = 2 kCD0
(CL /CD )max

We observe that for a long and shallow glide, the aircraft should
• Use a large aspect ratio, so that k is small
• Keep CD0 small by minimizing skin friction and frontal area

Example: Let CD0 = 0.01, AR = 16, and e ≈ 1. This gives


1 1
k= =
πeAR 16π
p 1 1
tan(|γ|min ) = 2 kCD0 = √ =
20 π 35.45

1
Thus, if this glider is left from an altitude of 3000 m (about 10000 ft), then it will cover nearly 106 km on
the ground.
To calculate the optimum flight speed, assume W/S ≈ 330:
s
2 2W k 10 W
V = = √ ≈ 775 =⇒ V = 28 m/s = 101 km/h
ρ S CD0 2.4 π S

• Gliding flight can directly yield the lift and drag coefficients of an airframe
• The flight path angle γ yields the CL /CD ratio
• CL can be obtained from V and γ:
2W cos γ
V2 =
ρS

• Therefore, data from several gliding flights will be of the form CD (CL ): obtain CD0 and k by fitting a
curve of the form CD0 + kCL2

2 Climbing Flight
The objectives of this section are to:
• Determine the maximum rate of climb

• Determine the corresponding flight speed


Yet again, we start with the equilibrium equations of motion which are given by
1 2
L = W cos γ = ρV SCL
2
1
T = D + W sin γ = ρV 2 SCD + W sin γ
2
The student should verify these by drawing a free-body diagram.
The motion in the vertical plane is described by

ẋ = V cos γ, ḣ = V sin γ

so that the rate of climb is given by RC = ḣ = V sin γ.


We start our analysis with the drag equation,

(T − D)V
RC = V sin γ =
W
At a given flight speed, the rate of climb is maximized by setting the thrust or the power available (for jet
and propeller-powered aircraft, respectively) to the maximum value:

(TA V − DV ) PA − DV SEP
RCmax (V ) = = =:
W W g
where TA and PA are the maximum available values of thrust and power, respectively.

2
2.1 Maximum RC in a Propeller Aircraft
In a propeller powered aircraft, PA is independent of the flight speed, and depends only on the altitude.
Thus, the maximum climb rate is given by
PA − DV
RCmax (V ) =
W
which means that the global maximum climb rate is
PA DV
RCmax = max RCmax (V ) = − min V
V W W
Notice, however, that
1 2 2W cos γ
D= ρV S(CD0 + kCL2 ), CL =
2 ρV 2 S
The presence of cos γ is a complicating factor. Fortunately, most aircraft climb at angles where it is reasonable
to approximate cos γ = 1. Once this approximation is made, notice the drag term is identical to that for
straight and level flight. Therefore, the maximum rate of climb is achieved by selecting
!
3/2
r
CL 3CD0
α : , i.e., CL =
CD k
max
1/4
4kW 2
 
VP min = = 0.76VDmin
3ρ2 S 2 CD0
Note that the speed at which the maximum RC is achieved depends strongly on the altitude. The drag at
minimum power is given by
2
3/4 √ s
W3

4kW 4 2 1/4 3/4
Pmin = 2ρVP3min SCD0 = 2ρSCD0 = (C )k
3ρ2 S 2 CD0 31/4 D0 ρS

Hence, the maximum rate of climb is given by


√ s
PA 4 2 1/4 3/4 W
RCmax = − 1/4 (CD0 )k
W 3 ρS

Note that the maximum cruising altitude can be found by obtaining the ρ at which RCmax (ρ) = 0. The
maximum rate of climb reduces with increasing altitude and increasing weight. Furthermore, the maximum
flyable altitude reduces with increasing wing loading.

2.2 Maximum RC of a Jet Aircraft


In a jet aircraft, PA is a linear function of the flight speed:
(TA − D)V
RCmax (V ) =
W
We find the speed V for RCmax by maximizing TA V − DV :
∂(TA V − DV )
=0
∂V
3 2kW 2
=⇒ TA = ρV 2 SCD0 −
2 ρSV 2
p
2 TA + TA2 + 12kCD0 W 2
=⇒ VRCmax =
3ρSCD0

3
Fastest
Climb
RC
Thrust constraint
Steepest
Climb

Flight
Envelope
Stall
constraint

Steepest Fastest V
Climb Climb

Figure 1: A sketch of the maximum rate of climb as a function of the flight speed. Notice that climbs are
possible at sub-stall speeds.

The speed for the maximum rate of climb increases with increasing altitude. The rate of climb, though,
reduces with altitude. The maximum rate of climb can be found by substituting for VRCmax in the equation
for RCmax above. The resulting expression is cumbersome to write (and look at), but there is much to be
gained by plotting the resulting RCmax (V ) as shown in Fig. 1.
We make some quick observations about the speed for maximum RC:
• Since TA ≥ Dmin , we have that VRCmax > VDmin .
• CL in climbing flight:
2W cos γ
CL (climb) = = CL (level) cos(γ)
ρV 2 S
Thus, it is possible for the aircraft to climb at stall speed and at lower speeds provided sufficient thrust
is available:
W
CD0 + kCL2 max

TA > Tstall =
CL max
• This gives a flight path angle envelope:
– V > Vstall : [0, γmax ]; constraint on TA dictates γmax
ρV 2 SCL max
– V < Vstall : [γmin , γmax ], where cos γmin = 2W

• Minimum flyable speed: γmax = γmin > 0

2.3 Steepest Climb versus Fastest Climb


The flight path angle can be found from
T −D
sin γ =
W
Yet again, the maximum climb angle at a given V is achieved by setting T = TA , i.e., by ensuring that the
excess thrust (unlike excess power for rate of climb) is maximized.
Assuming that cos γ ≈ 1, so that D(V ) is the same as that under level flight conditions, it is evident that
a jet aircraft climbs steepest at V = VDmin . That the steepest climb speed should be slower than that for
maximum rate of climb is apparent from Fig. 1. For a jet aircraft, this is validated by the observations in
the previous section.
The speed at which a propeller-powered aircraft achieves the steepest climb is found by formally differ-
entiating the expression for sin γ:
∂(TA (V ) − D)
γmax =⇒ =0
∂V

4
This exercise requires a model for TA (V ). Since TA (V ) is usually quite nonlinear, the exercise is cumbersome
and is left to the reader’s enthusiasm.

3 Level Turn Performance

R O

Source: Russell

Figure 2: Free-body diagram of a turning aircraft

So far, we have looked at flight in a straight line. Next, we move to turning manoeuvres which are
performed to bring about a prescribed change in aircraft heading. We would like to determine the maximum
turn rate that can be achieved at a given flight speed and, by extension, the maximum achievable turn rate
and the minimum attainable turn radius.
We make some assumptions regarding the turn:
• The aircraft is banked through a constant angle φ (see Fig. 2
• The sideslip angle is zero; i.e., β = 0
This is shown in Fig. 2. The free-body diagram can be used to deduce two equations of motion which arise
from lift:
Weight balance: L cos φ = W
W ωV
Centripetal force: L sin φ = (1)
g
Depending on whether or not a third condition γ = 0 is met, we get two types of turns (both of which satisfy
β = 0 and φ = constant):
• Sustained turn: γ = 0 is satisfied. This entails that T = D.
• Instantaneous turn: γ need not be zero. The condition T = D is not imposed on the problem.

3.1 Instantaneous Turn


We start by analyzing Eq. (1). By dividing the weight balance equation by that for centripetal force, we get
ωV
tan φ =
g
We define the load factor n = L/W . With this new terminology, we see that
g gp 2
ω = tan φ = n −1 (2)
V V
This is the fundamental equation which describes the turn rate in terms of the flight speed and the load
factor. We make some important observations:

5
• As V increases, ω decreases for fixed load factor
• For a given V , the turn rate increases as n increases
Evidently, in order to increase the turn rate at a given V , it is necessary to maximize the load factor n.
There are two fundamental constraints on the load factor:
ρV 2 SCLmax
• CLmax : the maximum achievable value nlift (V ) = . Load factors higher than nlift (V ) are
2W
simply not attainable.
• Structural and human limits: regardless of the flight speed, the aircraft structure and the human pilot
impose a load factor bound denoted by nsafe . At high flight speeds, this bound is attainable, but unsafe
for the aircraft.

Thus, the maximum turn rate at a given flight speed is found using
gp 2
ωmax (V ) = nmax (V ) − 1, nmax = min(nlift (V ), nsafe ) (3)
V
With this, we claim (and it is quite easy to prove) that the turn rate is maximized globally at a speed Vc
which satisfies s
2W nsafe √
nlift (Vc ) = nsafe =⇒ Vc = = Vstall nsafe
ρSCLmax
This speed is called the corner speed. When V < Vc , the constraint due to CLmax determines the maximum
achievable turn rate, while for V > Vc , the maximum achievable turn rate is determined by the safe load
factor. Note also that Vc is directly proportional to Vstall . Thus, for a given value of nsafe , a low corner speed
can be achieved (for a high turn rate, from Eq. (2)) by designing for as small a Vstall as possible.
ng
When the load factor n2 >> 1, we have that ω ≈ . We can thus write the global maximum instanta-
V
neous turn rate as r r
gnsafe g √ ρnsafe CL max S
ωmax ≈ = nsafe = g
Vc Vstall 2 W
Thus, to maximize the instantaneous turn rate, we need the following design pointers:
• Keep the wing loading as low as possible. This is in sharp contrast to the wing loading required to
maximize range and flight speed. Thus, the most manoeuvrable aircraft seldom have a large range.
Their flight speed may be quite high, but that requires powerful engines.
• The values of CL max and nsafe should be as high as possible.
Notice also that the turn rate reduces as altitude increases.

3.2 Sustained Turn


A sustained turn differs from an instantaneous turn only in that the constraint T = D is obeyed, which
allows the turn to be performed at a constant altitude and maintained that way for a long time. Note that
Eq. (1) does not change and consequently, the analysis described above holds for sustained turns as well,
except for an additional constraint on CL (aside from CLmax and the one due to nsafe .
Consider an aircraft turning at a speed V and coefficient of lift CL . Then, the thrust required for level
flight, given by
1
T = D = ρV 2 S(CD0 + kCL2 ),
2

6
increases monotonically with CL . Thus, the maximum sustainable value of CL at a given V , denoted by
CL,T (V ), is given by
  s  
2 1 2TA 1 2TA
CL,T (V ) = − CD0 =⇒ CL,T (V ) = − CD0
k ρV 2 S k ρV 2 S

Hence, the maximum sustainable load factor is given by


s 
ρV 2 S 1

2TA
nT (V ) = − CD0
2W k ρV 2 S

Note that an additional constraint nT ≥ 1 is required for level flight. It is evident that nT first increases
with V up to a critical speed Vc,T , and then decreases as speed is increased. The speed Vc,T plays the role
of the corner speed for the thrust constraint alone. It can be checked that Vc,T is given by
2
ρVc,T SCD0 = TA

The corresponding interpretation for Vc,T is that it is the speed at which the skin friction drag equals the
induced drag, and the net drag equals the maximum thrust available at that altitude. The straight-and-level
flight analogue is that Vc,T is VDmin and Vmax rolled into one.
Maximum load factor is now given by
s s
2
ρVc,T S TA TA 1 TA
nc,T = 2 S = 2W =
2W kρVc,T kCD0 Dmin

If n2T >> 1, so that the maximum turn rate is approximately


r ! √  r ! r !
gnT ρS TA g ρS TA
ωc,T ≈ =g =
VT k 2W 2 kW W

Therefore, the sustained turn rate can be maximized by designing the aircraft with low wing loading, high
thrust-to-weight ratio and large aspect ratio.
It is important to note that the sustained turn constraint operates alongside the bounds imposed by
CLmax and nsafe . Therefore, the maximum permissible load factor for sustained turns at any given speed is

nmax (V ) = min(nsafe , nT (V ), nlift )

so that the maximum sustainable turn rate is


gp
ω(V ) = nmax (V )2 − 1
V
Figure 3 shows the maximum turn rate of the F/A-18 HARV as a function of the flight speed for each
individual constraint. We have considered two cases of nsafe : 5 and 9. This plot is quite typical for most
aircraft. The following observations are made from the plot, which has been split conveniently into several
regions:
• The two curves DKL CP-G correspond to nsafe = 5 and 9, respectively. The CLmax constraint is
captured by the curve A-C-D-CP, while the thrust is maximum on M-B-K-Q.
• Region 1: all constraints are satisfied, and represents sustained turns. The maximum sustained turn
rate occurs at point T.
• Regions 2 and 5: in these regions, only the thrust constraint is violated. These regions correspond to
instantaneous turns. The maximum instantaneous turn rate occurs at points D or CP, depending on
whether nsafe = 5 or 9.

7
Figure 3: Maximum turn rate as a function of flight speed for the F/A-18 HARV. Source: Paranjape and
Ananthkrishnan, AFM 2005.

• Region 3: the constraint on CL is violated. Therefore, turn rates in region 3 are simply not attainable.
• Region 4: If nsafe = 5, then only the safe load factor is exceeded. Sustained turns in this region are
possible, but not safe if nsafe = 5.
Note that such plots are altitude-specific. The overlap between the various regions may change as the altitude
changes.

3.3 Turn Radius


An important metric for studying the turning performance is the turn radius defined by
V
R=
ω
Just as for steepest climbs, the minimum turn radius can be found by drawing a straight line from the
origin to an ω − V plot such as Fig. 3 and looking for the line that is tangential to the plot. Evidently, the
minimum instantaneous turn radius is obtained along the CL max branch. For sustained turns, it may occur
on the branch for the nT constraint.
There is simple way to calculate the smallest instantaneous turn radius. It is to be expected that the
sharpest turns are obtained for a large value of n >> 1. Thus, we can write

V2 2 W
Rmin = =
gnlift ρgCLmax S

4 Take-Off and Landing


Having seen all the major phases of flight, we turn our attention to the ground roll that accompanies take-off
and landing. The objective here is to minimize the ground roll and enable aircraft to use shorter runways.

4.1 Take-Off
During take-off, the aircraft accelerates to a critical speed called VR , at which it rotates its nose upwards
and climbs. The speed VR is called the rotation speed and is typically 1.25 Vstall . The ground roll distance,
denoted by sT O , is found using
Z τT O Z VR
V dV
sT O = V dt =
0 0 V̇
The acceleration V̇ is decided by a combination of thrust, drag and friction:
TA − D − µN
V̇ = = (TA − µW ) − (D − µL)
m

8
where µ is the coefficient of friction and N = W − L is the normal force. Substituting for V̇ in the equation
for sT O gives
Z τT O Z VR Z VR
V dV mV dV
sT O = V dt = =
0 0 V̇ 0 (TA − µW ) − (D − µL)
2
Since the aircraft is at a near-zero angle of attack, CL = CL0 and CD = CD0 + kCL0 . Therefore,
1 2 2

D − µL = ρV S CD0 + kCL0 − µCL0
2 | {z }
CDe

Thus,
VR  
T − µW
Z
mV dV W
sT O = 1 = ln
0
2
(TA − µW ) − 2 ρV SCDe ρgSCDe T − µW − 12 ρVR2 SCDe
To minimize the ground roll distance, the aircraft needs:
• Large T /W ratio: ensures that ln(·) is small
• Smallest possible CDe , ensured by designing CL0 = µ/k ≈ 1.
• Small wing loading W/S. Flaps can be used so that CL,max is increased over the normal wings, W/S
is reduced, and CL0 is close to the optimum value.
Notice that the take-off length increases when ρ reduces; i.e., at high altitude and under hot and dry
conditions.

4.2 Landing
The ground roll required to come to a complete halt during landing is given by
Z 0
V dV
sL =
VT D V̇
where VT D is the touch-down speed. The net acceleration:
D + TR + µN D − µL + TR + µW
|V̇ | = =
m m
where TR is the thrust from thrust-reversers.
Following an approach similar to that used for calculating the ground roll for take-off, we get

TR + µW + 21 ρVT2D SCDe
 
W
sL = ln
gρSCDe TR + µW
 
W (TR /W ) + µ + 1.5CDe /CL,max
≈ ln
gρSCDe (TR /W ) + µ
2
CDe = CD0 + kCL0 − µCL0

To minimize the landing distance, the aircraft needs:


• Large CDe : flaps and wing spoilers are deployed for this purpose.
• As large a value of µ as possible. This is done using brakes, but only at low speeds to prevent excessive
mechanical wear.
• As large TR as possible: this is accomplished using thrust reversers.

9
Chapter 4
Longitudinal Static Stability
Aditya A. Paranjape

1 Introduction
Thus far, we have looked at how aircraft can be designed to meet the desired performance specifications. We
also know the angle of attack and thrust that have to be set in order to achieve the desired flight speed and
climb rate. The performance analysis, however, dealt with a point-mass model of the aircraft. Aside from
macro-scale metrics such as W/S and the aspect ratio, the geometry of the aircraft was all but ignored.
The fact is that the geometry is an essential element of what makes an aircraft fly. Consider the following
questions, neither of which our prior analysis would enable us to answer but both of which are essential for
safe flight:
1. The performance analysis essential prescribes the angle of attack α at which the aircraft must be flown.
Can this value even be attained by the aircraft? What does it take to attain a specific value of α?
2. If the aircraft is disturbed by an external gust and its angle of attack changes in the process, can the
aircraft restore its angle of attack by itself? If not, what can be done to ensure that it does?
These questions fall within the purview of stability and control - two topics which will occupy us for the
remainder of this course.
We start by reviewing a narrow but useful notion of stability, called static stability. Next, by deriving
expressions for rotational equilibrium and by analysing small perturbations about an equilibrium, we answer
the questions about stability and control.

2 Static Stability
2.1 Broad Definition
Broadly speaking, an aircraft is said to be stable if it is able to restore its flight speed, angle of attack, etc., to
the set values after encountering a disturbance. It is evident that stability refers to the long-term behaviour
of aircraft.
Here is a different perspective on stability: suppose an aircraft is disturbed from its operating condition.
Then, it may be said to be stable if the forces and moments instantaneously try to restore the operating
condition. They may or may not succeed in doing so in the long-run. If the instantaneous response is in the
correct direction, we say that the aircraft is statically stable.
An analogy is shown in Fig. 1. One may conjecture that the ball is stable inside the valley and unstable
at the peak. Strictly speaking, the ball is only statically stable inside the valley. In the absence of friction,
the ball would simply travel back and forth like the bob of a pendulum and is clearly not stable in the
long-run.
Formally, we say that a system is statically stable if, upon a disturbance ∆x, the force F produced by
the system is such that F · ∆x < 0. This is considered to be a necessary condition for long-term stability, but
is not sufficient (as in the case of a ball in a valley). In fact, it is not difficult to produce an example where
a statically stable system is, in fact, unstable in the long run. This simple exercise is left to the reader.

1
Figure 1: Classic example of a ball in a valley or on a ridge.

2.2 Longitudinal Static Stability


In this chapter, we are interested in the pitching motion of an aircraft. As a matter of convention, the
pitching moment on aircraft is denoted by M . We may write M (α) since M depends on α.
Definition: We define an equilibrium angle of attack α0 as one at which M (α0 ) = 0. This is the
conventional definition of equilibrium: a state in which the net forces and moments on a body add up to
zero. Suppose that ∆α denotes the perturbation about α0 . This perturbation produces a pitching moment
∆M (∆α). The aircraft is said to possess longitudinal static stability if ∆α · ∆M = 0.
Since M (α0 ) = 0, we can use Taylor series expansion to write
 
∂M
∆M = M (α0 + ∆α) = ∆α , Mα (α0 )∆α
∂α α0

It follows that the aircraft is statically stable at α0 only if Mα (α0 ) = 0. Notice that we have defined stability
as being a property not just of the aircraft as a whole, but also a pointwise property at each α0 .
If we assume that the pitching moment M is a linear function of α, then we may write

M (α) = M0 + Mα α

In this case, the partial derivative Mα does not depend on the trim value α0 .
As a matter of convention, we define the non-dimensional coefficient of pitching moment Cm such that
1 2
M= ρV ScCm
2
and we can further write
1 2 1
M0 = ρV ScCm0 , Mα = ρV 2 ScCmα
2 2
Notice that
∂Cm
Cmα =
∂α
We now ask the question: under what conditions is it possible to trim and statically stabilize the aircraft
at some α0 > 0. From Cm = Cm0 + Cmα α, we readily conclude that

1. In order to trim at some positive angle of attack, we require that Cm0 > 0. The corresponding trim
Cm0
angle of attack is given by α0 = −
Cmα
2. In order for the trimmed condition to be stable, we require that Cmα < 0.
These are fundamental conditions for static trim and stability, and the rest of this chapter builds upon these
equations systematically.

2
Figure 2: Wing-only model for trim and stability analysis.

3 Longitudinal Static Stability and Aircraft Geometry


3.1 Wing-Body Only
Consider an aircraft with a wing and a fuselage, but no horizontal tail. Suppose the wing AC is at a distance
xAC from the CG of the aircraft, as shown in Fig. 2. The pitching moment is derived only from lift, so that
1 2
M= ρV S (cCmac + xAC CL (α)) , CL = CL0 + CLα α
2
We first determine the trim angle of attack, α0 , by setting M (α0 ) = 0:
1 2
ρV S cCmAC + xAC (CL0 + CLα α0 ) = 0

2
cCmAC + xAC CL0
=⇒ α0 = −
xAC CLα
The trimmed value of lift (with CmAC < 0) is given by

1 2

cCmAC > 0, xAC > 0

L0 = ρV S −
2 xAC 
< 0, xAC < 0

We conclude that the trimmed value of lift is positive only if xAC > 0, i.e., if the CG is behind the AC, as
shown in Fig. 2. However, is this configuration stable? To determine the stability of this trim, suppose the
AoA is perturbed by ∆α. The instantaneous pitching moment is given by
1 2
∆M = xAC Lα ∆α = ρV SCLα xAC ∆α
2
To ensure static stability, we need ∆α · ∆M < 0, i.e., xAC < 0, which runs contrary to the requirement for
positive lift. Therefore, a wing-body combination that is flight-worthy (capable of producing positive lift)
will be statically unstable. Since static stability is a necessary condition for stability, the aircraft will be
unstable in the long-term as well. There are two ways to tackle this problem:
• Use a flap and active control: a flap changes lift as well as Cmac and offers a direct way to mitigate a
disturbance.
• Use an airfoil cross-section with a negative camber at the trailing edge. This configuration ensures
that Cmac > 0. Such airfoils, however, are not particularly suitable at high angles of attack.
• Use a combination of wing sweep and washout (reducing wing twist at outboard sections). This shifts
the CG of the aircraft behind the aerodynamic centre of the inboard wing sections, so that stability
characteristics improve substantially.

3.2 Longitudinal Control: Trimming with a Movable Horizontal Tail


Let α0 be the trim angle of attack of the aircraft. We will assume that the wing inclination angle iw = 0
and that the tail is symmetric. The lift on the tail is given by
1 2
Lt = ρV St CLα (α + it )
2

3
Figure 3: Wing and tail schematic for trim and stability analysis.

Note: if the tail has an elevator instead of variable it as the control input, then
1 2
Lt = ρV St (CLα (α + it ) + CLδe δe )
2
At trim, M = M wing + M tail = 0 about the CG; i.e.,

M wing + Lwing xAC = (lt − xAC )Ltail


 
lt St xAC St
Thus, xAC (CL0 + CLα α) + cCmac = − CLα (α + it )
S S
 
xAC St xAC
i.e., (CL0 + CLα α) + Cmac = VH − CLα (α + it )
c S c

The trim AoA is given by

VH − (St /S) xAC


 
1
α0 = ((xac /c)CL0 + Cmac ) − c
it (1)
VH − xAC VH − xAC

c (1 + St /S) CLα c (1 + St /S)

xAC
As we will see shortly, VH > (1 + St /S). Therefore, as it increases (i.e., deflects downwards), the trim
c
α0 reduces, and vice-versa. Therefore, the tail-based control surface allows the aircraft to trim across a wide
range of values of lift (and flight speeds). The same principle applies to elevator-based control where it is
held constant, but the elevator is deflected to achieve the desired moment.
Suppose that the angle of attack is perturbed by a small ∆α. We note the following change in the
aerodynamic forces and pitching moment:
• Change in lift on the wing: ∆Lw = 21 ρV 2 SCLα ∆α
• Change in lift on HT: ∆Lt = 21 ρV 2 St CLα ∆α

• Net change in pitching moment


   
1 2 xAC St
∆M = ∆Lw xAC − ∆Lt (lt − xAC ) = ρV ScCLα ∆α 1+ − VH
2 c S

We deduce that    
1 xAC St
Mα = ρV 2 ScCLα 1+ − VH
2 c S
We infer that the aircraft is statically stable only if xAC < VH /(1 + St /S) = VH0 . For a given aircraft, VH0 is
thus a constant, and so is c. Thus, the static stability condition imposes the requirement on how the aircraft
is loaded: the CG location should satisfy xAC < cVH0 . Since xAC > 0 when the CG is behind the AC, the
expression xAC > cVH0 tells us how far behind the wing AC the CG is allowed to lie.
Definition: We define the neutral point as the CG location at which xAC = cVH0 , i.e., Mα = 0. When
the CG is located ahead of the neutral point, the aircraft is statically stable, and vice-versa. It is important
to note that the location of the neutral point depends only on the geometry and aerodynamic coefficients.

4
It is instructive to consider the non-dimensional CG location xAC /c. For static stability,
!
xAC St lt 1
>
c S c 1 + SSt

Since St /S is usually between 0.25 and 0.3, we infer that the CG can lie at most a quarter of the way from
the wing to the tail.
Definition: We define the static margin:
xN P − xAC
SM ,
c
An aircraft is loaded on the ground with payload and fuel so that the static margin never decreases below a
threshold. The static margin is used largely for operational purposes.

3.3 Effect of Downwash

Figure 4: Trailing edge vortices. Source: aerospaceweb.org.

Recall that trailing edge vortices are generated by finite wings (Fig. 4 as a consequence of lift production.
The resulting downwash reduces the angle of attack of the horizontal tail:

αt = α + it − 
|{z}
| {z }
geometric downwash

The downwash can be written as  := (α) = 0 + α α, where 0 > 0 and α > 0. Thus,

αt = α(1 − α ) + it − 0

The presence of α changes Mα :


 
xAC St xAC
CMα = CLα − VH − CLα (1 − α )
c S c

Clearly, downwash reduces the longitudinal-stability


 of the aircraft, and the neutral point shifts forward to
xAC (1 + (S t /S))(1 −  α )
satisfy = VH0 . The exact expression for  is difficult to derive. The reader is
c 1 + (St /S)(1 − α )
referred to a textbook for standard approximations, but is cautioned against using it for analysis without a
careful review of the assumptions.

5
3.4 Statically Unstable Aircraft and Rearmost CG Location
It is interesting to note that several aircraft are designed to be statically unstable, and have to be stabilized
using automatic control systems. This is especially true of fighter aircraft. In general, statically unstable
aircraft are more manoeuvrable than their stable counterparts. There is another benefit to having a statically
unstable airframe, namely that the net lift is higher. We perform a simple analysis to ascertain this. At a
given flight speed, the trim condition is

Lw xac + MAC = Lt lt
 
w t w xAC MAC
=⇒ Lnet = L + L = L 1+ +
lt lt

Clearly, the higher the value of xAC , the greater the net lift on the aircraft. This begs the question: how far
back can the CG be allowed to shift? The answer to this question lies at the heart of control: the CG can
be permitted to move to the rear as long as the pitching moment from the horizontal tail is adequate.
It is evident that as the CG moves backwards, the pitch-up moment from the wing increases rapidly.
In order to trim the aircraft, the horizontal tail must be able to provide a sufficient pitch-down moment.
In particular, if the maximum desired angle of attack is αmax , then the horizontal tail must be able to
balance the pitching moment generated by the wing at αmax while ensuring that its own deflection is within
permissible limits.
Let δ denote the control surface deflection (which could be either the elevator or the complete tail, or
both) and suppose that we need δ ∈ [δmin , δmax ]. The CL of the tail can be written as

CLt = CLα α + CLδ δ

The corresponding control surface deflection is found using Eq. (1):


   
xAC St (xAC /c)CL0 + Cmac
it = 1+ − VH α0 + (2)
c S CLα (VH − (St /S)xAC /c)

If we wish to fly between angles of attack of αl and αu , then we need to ensure that δ(αl ) and δ(αu ) are
within the deflection bounds. Since xAC > cVH0 , we note that δ(α) increases with α. Thus, the rearmost CG
location is then found by solving for xAC when
• δ(αl ) = δmin and δ(αu ) < δmax , or
• δ(αl ) < δmin and δ(αu ) = δmax
The first constraint is usually milder, and the second expression yields the rearmost CG location.

3.5 Positioning the Horizontal Stabilizer: T-Tails and Canards

Figure 5: Examples of T-tails and canards in aircraft. Source: Wikipedia.

Figure 5 shows two alternate longitudinal stabilizer configurations:

6
• T-tail: the horizontal stabilizer mounted on top of the tail. The arrangement keeps the stabilizer out of
the downwash and makes room for rear engines and cargo doors. On the down-side, the static stability
is reduced due to drag and the tail becomes highly susceptible to a dangerous phenomenon called a
deep-stall by being in the wing wake at high α.
• Canard: the horizontal stabilizer is mounted ahead of the wing. Canards improve the maneuverability
of the aircraft significantly. Since they are never in the wing wake, they are quite effective even when
the aircraft pitches to high α. On the flip-side, the use of canards inherently reduces the static stability
of the aircraft. They may also introduce a performance-degrading downwash on the wings.
It is interesting to note that a canard-wing combination is exactly like a wing-HT combination. The static
stability condition can be found using that for the wing - horizontal tail equation by replacing the wing with
the canard, and the horizontal tail by the wing. The analysis is left to the reader who should verify that
• The static stability condition translates into a lower bound on how far ahead of the wing the CG should
be placed.
• Imposing static stability degrades the control authority of the canard.

3.6 Effect of Nonlinearities


In the analysis presented so far, we ignored contributions from the fuselage and other external payload.
Their contributions are usually nonlinear in α. Yet, the control input δ enters as a linear term, so that the
total pitching moment on the aircraft at equilibrium can be written as

M = M (α) + Mδ δ = 0 at trim

The incremental moment after a perturbation ∆α is given by


 
∂M
∆M = ∆α = Mα (α0 )∆α
∂α α0

Therefore, the condition for static stability can be written as Mα (α0 ) < 0, which is the same as the one we
had before with the difference being that the derivative needs to be evaluated at each trim condition.

3.7 Neutrally Stable Configurations and Trim


Let us return to the case where M is linear in α. The equilibrium pitching moment is given by

M = M0 + Mα α + Mδ δ = 0 =⇒ α = − δ

If |Mα | is very small (close to zero), the angle of attack changes by large amounts even for small elevator
deflection. This is highly undesirable. Furthermore, if Mα = 0, then at equilibrium, M = M0 + Mδ δ = 0.
The angle of attack is nowhere in the picture! Therefore, without active control, the aircraft can trim at any
angle of attack, i.e., there is no control whatsoever on the trim value of α and CL .

4 Stick-Free Versus Stick-Fixed Stability


Longitudinal control surfaces are typically actuated by a combination of actuators:

• Electro-mechanical actuators connected to the flight computer


• Hydraulic actuators connected to the flight computer as well as the control column in the cockpit

7
• In small aircraft, mechanical wires and pulleys connected directly to the control column
In all cases except the electro-mechanical actuators, the force exerted by the pilot on the control column is
transmitted to the actuator to move the control surface. The control surface exerts an opposing force on
the actuator which is duly transmitted back to the control column. Under equilibrium conditions, the two
forces are equal and cancel each other.
So far, we assumed that the elevator deflection is constant, for which the pilot would have to hold the
control column in one place manually. The stability that we have looked at so far is therefore called “stick-
fixed” stability. In nominal trim flight, the pilot takes his hands off the control column after “arranging” for
a certain amount of force to be applied to the elevator at all times. This is achieved by using a small flap
on the elevator, called the trim tab. Thus, the elevator is no longer statically deflected and the dynamics of
the elevator deflection further affect the stability of the aircraft. The stability (or lack thereof) is referred to
“stick-free” stability (or instability). Although stick-free flight is rare in the present times, it is still widely
prevalent in general aviation and sports aircraft.
We start our analysis of stick-free stability with the pitching moment equilibrium equation:
M = M0 + Mα α + M δ δ
As we will show presently, the elevator/horizontal tail deflection can be written δ = H0 + Hα α, for some
constants H0 and Hα . Substituting into the pitching moment expression yields
M = (M0 + Mδ H0 ) + (Mα + Mδ Hα )α
The static stability condition changes to ensuring that Mα + Mδ Hα < 0. If Mδ Hα > 0, then we need a much
more negative Mα to ensure stick-free stability. Thus, stick-free stability is a much more stringent attribute
to achieve than stick-fixed stability. Usually, Mδ < 0 for a horizontal stabilizer (can it ever be otherwise?).
Thus, Hα < 0 is destabilizing and vice-versa. We need to derive an approximate expression for H.

(a) An actual trim tab (b) Schematic

Figure 6: Elevator trim tabs and a schematic representation. Source for the image on the left: Wikipedia.

A trim-tab is shown in Fig. 6 together with a schematic representation. In principle, the trim tab behaves
like an independent symmetric airfoil. The force on the tab produces a moment which adds on to the moment
due to the stick force. Our objective is to find the tab deflection angle δtab to achieve zero hinge moment.
The hinge moment consists of two contributions:
• Moment from the elevator: Me = 21 ρV 2 Se CLα
e
(α + δe )xe
• Moment from the trim tab: Mtab = 12 ρV 2 Stab CLα
tab
(α + δtab )xtab
Since equilibrium is achieved when Me + Mtab = 0, the trim elevator angle is given by
tab tab
   
Stab xtab CLα Stab xtab CLα
δe = − e −1 α− e δtab
Se xe CLα Se xe CLα
A comparison with δ = H0 + Hα α clearly shows that Hα < 0. This proves that stick-free configurations are
less stable than stick-fixed configurations.

8
Chapter 5
Lateral-Directional Static Stability
Aditya A. Paranjape

1 Elements of Lateral-Directional (L-D) Static Stability

Figure 1: The three primary angles of interest, all defined with respect to the velocity vector.

Figure 1 shows the three primary rotational degrees of freedom of an aircraft. Notice that these rotations
are defined with respect to the velocity vector. The resulting angles are also thus referred to as “wind axis
angles.” The perturbation ∆α is the “defining perturbation” for longitudinal static stability. On similar lines,
since there are two rotational lateral-directional degrees of freedom, we have two candidate lateral-directional
angles:
• Yaw: perturbations in the the sideslip angle ∆β are the direct analogue of ∆α for lateral-directional
stability.

• Roll: a static perturbation ∆µ (roll) does not alter the orientation of the velocity vector with respect
to the aircraft, and hence does not produce incremental aerodynamic forces and moments by itself.
Therefore, the perturbation ∆µ is not relevant to our discussion on static stability.
Since sideslip arises primarily as a result of the yawing motion, we expect that the yawing moment
derivative Nβ would be of interest. It turns out that the rolling moment Lβ is of interest as well. This
is because a perturbation in the body-axis roll angle (rotation about the body x axis) at non-zero angles
of attack produces a sideslip along with change in µ. This has been shown in Fig. 2, where the sideslip
perturbation is given by ∆β = sin α0 ∆φ. Notice that the perturbation depends on the trim angle of attack
and is zero when α0 = 0.

Important: we will use the symbol L for lift as well as rolling moment, following the flight mechanic
convention. The reader should pay attention to the context to avoid any confusion. Interestingly, the
notation differs for the non-dimensional coefficients of lift and rolling moment, which are denoted using as
CL and Cl , respectively.

1
Figure 2: The change in sideslip due to a perturbation in the body-axis roll angle: ∆β = sin α ∆φ. Note
that φ and µ are very different, if related, quantities.

1.1 General Theory


Figure 1 shows the positive sense of the angles. The sideslip is positive when the velocity vector is to the
right of the aircraft. From Fig. 2, it is also clear that a roll to the right creates positive sideslip when α > 0.
A positive sideslip therefore needs to be countered by (a) a positive yawing moment, and (b) a negative
rolling moment.

Conditions for stability: Following the notation established for longitudinal stability, we infer that for
static stability, we need Nβ > 0 and Lβ < 0.
The above condition is also usually written in terms of the corresponding non-dimensional quantities
1
Cnβ > 0 and Clβ < 0, where Nβ = ρV 2 SbCnβ and likewise for Lβ .
2
In what follows, we will derive the contributions from the wing and the tail to the two derivatives Nβ
and Lβ . The derivations require application of the analytical version of strip theory - the method can be
easily converted into a numerical form as well.

2 Contribution from the Vertical Tail

Figure 3: Vertical tail contribution to the rolling and yawing moments.

2
2.1 Yawing moment
Let bv , cv and Sv denote the span, chord and area of the vertical tail. From Fig. 3, it is clear that β plays
the role of angle of attack as far as the vertical tail is concerned. The yawing moment generated by the lift
Lv produced by the vertical tail tail is
N = Lv (lt − xAC )
where lt is the distance between the tail and the wing AC. Since the vertical tail is symmetric, the yawing
moment would be zero β = 0. The incremental yawing moment due to a small perturbation ∆β is thus given
by
1 2
∆N = ρV Sv (lt − xAC )CLα ∆β
2
1 2
=⇒ Nβ = ρV Sv (lt − xAC )CLα > 0 (1)
2
Therefore, the vertical tail stabilizes the derivative Nβ .

2.2 Rolling Moment


The rolling moment needs to be calculated using strip theory because the spanwise aerodynamic centres are
at varying distances from the x axis. Consider a strip of width dz whose aerodynamic centre is at a distance
z from the rolling axis. Then, the incremental rolling moment from this strip is given by
1
d(∆L) = − ρV 2 cv CLα z dz ∆β
2
Notice the negative sign: this comes from the fact that the lift on the vertical tail, which points to the left
when ∆β > 0, causes the aircraft to roll to the left.
By integrating from z = 0 to z = bv , we get
1
Lβ = − ρV 2 Sv bv CLα < 0 (2)
4
Therefore, a vertical tail makes stabilizing contributions to the Lβ derivative as well provided it is located
above the fuselage.

3 Contributions from the Wing: Important Parameters


The effect of the wing on L-D stability arises from three geometric characteristics of the wing:
• Wing dihedral angle Γ (positive upwards for both wings)
• Wing sweep angle Λ (positive backwards for both wings)

• High wing versus low wing


These have been depicted in Fig. 4. The net influence of a wing is the sum total of the individual contributions
of these three effects, and we will analyse each in isolation.

3.1 Dihedral Effect


The basis of the dihedral effect is the following chain of events which has been visualised in Fig. 5.
1. A perturbation ∆β causes a perturbation in the local angle of attack ∆αl at each spanwise section of
the wing.

3
Figure 4: Wing configuration angles. The dihedral and sweep angles shown here are in the positive sense.

Figure 5: The physical basis for the dihedral effect.

2. The resultant ∆αl is anti-symmetric; i.e., it has opposite signs on the two wings, and it is constant on
each wing.
3. The resultant asymmetry in the lift produces a rolling moment, and a yawing moment is produced due
to the AC of the wing being longitudinally separated from the CG (i.e., xAC 6= 0) (see Fig. 6).
As a matter of convention, the body axis components of V are denoted by u, v, w along the x, y and z
axes, respectively. The angle of attack is defined as tan α = w/u. The dihedral effect essentially allows v to
contribute to the local angle of attack. On the right wing, the local angle of attack is thus given by
w + v sin Γ
αl,R ≈
u
The change due to the sideslip perturbation is, therefore, ∆αl,R = ∆β sin Γ, where we have additionally
assumed that α and β are sufficiently small and u ≈ V . By the same token, the change of the local angle of
attack of the left wing is ∆αl,L = −∆β sin Γ.
The difference in the angles of attack between the left and right wings gives rise to a rolling moment
which can be found using strip theory to be
 
1
∆L = − ρV 2 SbCLα sin Γ ∆β
4
1
Thus, Lβ = − ρV 2 SbCLα sin γ < 0 when Γ > 0. In other words, a positive dihedral angle improves static
4
stability in roll (Lβ ), while a negative dihedral (called anhedral; Γ < 0) reduces static stability in roll.

3.2 Dihedral and Yawing Moment


The analysis in the previous section leads us to believe that the dihedral angle allows the wing to behave
somewhat like the vertical tail. Thus, we expect that the wing would contribute to improving Nβ as well.

4
Figure 6: The moment of the asymmetric lift about the CG produces a rolling moment and a yawing moment.

As we show, this does happen, but the CG’s proximity to the wing makes the wing a secondary contributor
to Nβ .
Suppose the wing AC is ahead of the CG, and let xAC denote the distance between the AC and CG.
Then, the yawing moment on the aircraft due to the dihedral is given by
N = (Liftleft − Liftright ) sin ΓxAC
Since the difference in angles of attack is given by ∆β sin Γ, we get
1
N = − ρV 2 SxAC CLα sin2 Γ∆β
4
1
=⇒ Nβ = − ρV 2 SxAC CLα sin2 Γ
4
Thus, sign(Nβ ) depends only on sign(xAC ): if xAC > 0 then increasing the absolute value of the dihedral
angle leads to a reduction in the static stability, and vice-versa. From the point of view of stability, yet
again, it helps to place the CG ahead of the wing AC.

4 Additional Notes on Lateral-Directional Stability


4.1 Wing-VT Combination
Just as in the case of longitudinal static stability, the wing dihedral setting destabilizes yaw (Nβ ) if the CG
is behind the wing, while the vertical tail has a stabilizing effect. The net Nβ is given by

Nβ = Nβwing + Nβtail (3)


 
1 2 Sv (lt − xAC ) xAC 2
= ρV SbCLα − sin Γ (4)
2 Sb 2b
Sv lt
Define the vertical tail volume ratio Vv = . This gives the following condition for static stability:
Sb
xAC sin2 Γ Sv
 
Static stability : Vc > +
b 2 S
Interestingly, as the dihedral angle increases, the rearmost CG location for static stability (which we could
call the directional neutral point) moves closer to the wing.

5
4.2 L-D Departure Criterion
The criteria Nβ and Lβ are static stability criteria, but they are useful pointers for dynamic stability as well
as other complex phenomena. The term criterion Cnβ ,dyn was defined by Moul and Paulson (1958) as
 
Iz
Cnβ dyn = Cnβ cos α − Clβ sin α
Ix
It can been argued that Cnβ ,dyn < 0 can indicate the likelihood of two dangerous phenomena:
• Wing-rock: rapid, large-amplitude rolling oscillations at high angles of attack.
• Spin: combination of high α, almost vertical flight path and large angular rates.
The condition for the onset of these phenomena is treated very often as Cnβ ,dyn = 0. Notice that Cnβ and
Clβ both depend on α. If the wings have an anhedral (i.e., if Clβ > 0), a directionally-stable aircraft (i.e.,
with Cnβ > 0 can experience departures at relatively low α.

5 Wing Sweep

Figure 7: A swept wing and the coordinate system required for analysis.

Consider a wing with sweep angle Λ, as shown in Fig. 7. Then, the lift on a small strip of the wing
perpendicular to the leading edge
1 1
dLift = ρ(V cos Λ)2 c cos Λ CLn ds = ρ(V cos Λ)2 c CLn dy
2 2
where
• CLn is the coefficient of lift of the normal section
• dy = cos Λ ds, and y ∈ [0, b/2]
There are two contributors to the Lβ derivative: V cos(Λ) and CLn . We will deal with each of these in turn.

5.1 Velocity
Suppose that the aircraft is perturbed by a small sideslip angle ∆β. Then, the effective sweep angle of the
right wing becomes (Λ − ∆β), while that of the left wing becomes (Λ + ∆β). Thus, the rolling moment on
the aircraft is given by
!
1 2 n b/2
Z
ydy cos2 (Λ + ∆β) − cos2 (Λ − ∆β)

L1 = ρV c CL
2 0
1 2
= ρV Sb CLn (− cos Λ sin Λ) ∆β
4
Therefore, we get (
1 2 n < 0, Λ > 0 (sweep-back)
L1β = − ρV Sb CL cos Λ sin Λ
4 > 0, Λ > 0 (sweep-forward)

6
5.2 Coefficient of Lift
We define the normal angle of attack
w tan α α
tan αn = = =⇒ αn ≈
u cos Λ cos Λ cos Λ
Since we perturb only the sideslip and not α (the angle of attack of the FRL is held constant while taking
the partial derivative), we get the following expressions for the angle of attack on the two wings:

n α
Right wing : αR = ,
cos(Λ − ∆β)
n α
Left wing : αL =
cos(Λ + ∆β)

The resulting rolling moment is given by


 
1 1 1
L2 = ρV 2 cos2 (Λ)SbCLnα α −
16 cos(Λ + ∆β) cos(Λ − ∆β)
1 2
=⇒ L2β = ρV SbCLnα α sin Λ
8
We now add the two contributions to obtain the derivative Lβ :
 n 
1 CL sin Λ cos Λ sin Λ
Lβ = L1β + L2β = − ρV 2 Sb − CLnα α
2 2 4

Note that CL = CLn cos2 Λ and CLα = cos ΛCLnα . This gives
 
1 2CL − CLα α
Lβ = − ρV 2 Sb tan Λ
2 4
 
2CL − CLα α CL + CL0
=⇒ Clβ = − tan Λ = − tan Λ
4 4

For stability, we need Λ > 0 (sweep-back); a forward-swept wing is statically unstable.


Note that, in most textbooks, the second term in Clβ is ignored. While the primary contribution does
come from the first term, the second term is not negligible and illustrates the importance of mathematical
rigour in lateral-directional dynamics in particular where it is difficult to isolate perturbations in any given
degree of freedom.

6 High Wing Versus Low Wing

Figure 8: Flow around the fuselage in the event of a lateral-directional perturbation affects the angle of
attack near the root of the wing.

7
The position of the wing relative to the fuselage reference line determines the direction of the flow
experienced by the wing in the event of a lateral-directional perturbation. Consider a high wing, such as
that in Fig. 8, and suppose that the aircraft gets a positive ∆β (without loss of generality). Then, the right
wing experiences an increase in angle of attack close to the root, while the left wing experiences a reduction
in the angle of attack close to the root. This effect is almost negligible near wing tips.
The change in the angle of attack induces a negative rolling moment, so that Lβ < 0. The exact opposite
happens for low wings, and thus low wings are destabilizing.
The change in Clβ due to the location of the wings is difficult to derive analytically. Rather, we use the
following estimates: (
−1 σ = −1 high wing
∆Clβ = σ 0.00917 rad
σ=1 low wing
It should be noted that several books and research papers use the phrase “dihedral effect” to refer to all
three terms looked at above. They are quite similar in nature and from the perspective of design, it helps
to take a combined view of the three terms. While static stability is desirable, too much static stability can
impede maneuverability. Thus, aircraft with high wings use wings in an anhedral configuration, while low
wings are usually accompanied by a positive dihedral setting. The choice of sweep angle is dictated usually
by Mach number considerations, and the other two parameters (dihedral and wing position) are chosen to
provide the remainder of the desired stability.

7 Lateral-Directional Control
7.1 Crosswind Trims
Lateral-direction trim is achieved using the rudder and ailerons together. The rudder and aileron deflections
are conventionally denoted by δr and δa , respectively.
The rudder develops a yawing moment as well as a rolling moment. These exact contributions can be
derived using strip theory and thin airfoil theory. We merely note that the corresponding derivatives Lδr < 0
and Nδr > 0.
The ailerons, for their part, primarily generated rolling moment. As a matter of convention, positive
aileron deflection corresponds to a downward deflection on the left wing and an equal and opposite deflection
on the right wing. Hence, Lδa > 0. The yawing moment from the ailerons can be neglected at low angles of
attack, and we choose to do as much.
In order to trim the aircraft at a sideslip angle β, we need to ensure that
L = Lβ β + Lδa δa + Lδr δr = 0
N = Nβ β + Nδr δr = 0 (5)
which gives the following expressions for δa and δr :

δr = − β
Nδr
 
Lδr Nβ Lβ
δa = − β (6)
Lδa Nδr Lδa
Although aircraft usually trim at β = 0, a non-zero sideslip is used in two distinct scenarios: while trying
to reduce speed rapidly using the increased drag due to sideslip, and while trying landing with crosswinds
while keeping the nose pointed along the runway.

7.2 Control Problems Involving the Ailerons


The control authority of the ailerons, measured by Clδa , can be compromised when the portion of the wings
containing the ailerons is stalled, or due to finite wing effects leading to reduced lift in the aileron portion
of the wings.

8
In the case of stall, aileron deflection leads to Clδa < 0; i.e., a positive aileron deflection may produce a
negative rolling moment. This situation, called aileron reversal, can arise either due to the aircraft pitching
to overly large angles of attack, or due to excessive wing loading at high speeds leading to sectional stall. In
the former case, pilots are trained to use the rudder to recover the aircraft to a controlled condition, while
in the latter case, it usually suffices to reduce the speed by reducing thrust.
The control authority of ailerons can also be reduced also due to finite wing effects, i.e, when the lift
distribution is such that the lift is smaller near the wing tips. This may pose a problem at low flight speeds.
One way to mitigate the problem is to use vortex generators to energize the flow in outboard areas of the
wing and make the lift distribution more uniform. The same effect can be achieved by adding winglets or
wing-tip fences. Alternately, spoilers may have to be employed to make up for the deficiency in the roll
control authority.

7.3 Lateral Control Departure Parameter (LCDP)


Earlier in this chapter, we noted that the criterion Cnβ ,dyn is a bellwether for lateral-directional instabili-
ties. Spin prediction, in particular, is done using Cnβ ,dyn and another parameter called the lateral control
departure parameter (LCDP) defined in its simplest form as

Cnδa
LCDP = Cnβ − Cl > 0 for stability
Clδa β

Notice that Cnδa cannot be ignored at high angles of attack and plays an important role in determining the
value of LCDP.
The safe flight envelope is thus usually approximated by Cnβ ,dyn > 0 and LCDP > 0. Different types
of departure phenomena arise when either one of the conditions fails to apply, spin being the result of
simultaneous failure. Interestingly enough, despite the complex nature of the departure phenomena, the
susceptibility of the aircraft can be predicted reasonably well using purely static criteria.

7.4 Adverse Yaw


For the purpose of ensuring smooth turns, it is essential that the rolling motion induced by the ailerons be
accompanied by a yaw in the same direction; i.e., a positive roll should be accompanied by a positive yaw
and vice-versa. When the yaw is in the opposite sense of roll, it is referred to as adverse yaw.
Adverse yaw is a serious problem in tailless aircraft (e.g., flying wings such as the B-2 bomber). A direct
C
way to prevent adverse yaw is to design for Cnl δa > 0. However, since Cnδa is very sensitive to the wing
δa
geometry, CG location and α, its sign cannot be taken for granted. In fact, a vertical tail goes a long way in
preventing adverse yaw (for reasons which will become apparent later in the book). Since tailless aircraft lack
a vertical tail, it is helpful to place the CG behind the wing AC - the dihedral effect enusres that a positive
rolling moment is accompanied by a positive yawing moment and vice-versa. However, an aft-located CG
compromises directional stability by making Cnβ < 0.
While this poses a stability-versus-control trade-off scenario, a bit of adverse yaw is actually helpful for
piloting. When the airframe has a tendency for adverse yaw, a pilot is required to use the rudder in the
same direction as the ailerons, which is a more natural pair of actions for the human body (a right on the
yoke and right leg down for right rudder, and vice-versa). This illustrates how complex aircraft design can
sometimes get, and the extent to which seemingly minor bells and whistles can have an incommensurate
influence on the design decisions.

9
Chapter 6
Longitudinal Flight Dynamics and Control
Aditya Paranjape

The objective of this chapter is to study the stability and control of an aircraft constrained to move in
its plane of symmetry. Although this is a restricted case of the complete aircraft motion, the results are in
fact valid for the complete aircraft due to the dynamic decoupling that we will see in the next chapter.

1 Review of Systems Theory


We start by reviewing the concepts of equilibrium solutions and stability of linear systems.

1.1 Equilibrium Solutions


Definition: Given a (linear or nonlinear) system ẋ = f (x, u), we define an equilibrium solution as a pair
(x0 , u0 ) of constants that satisfy f (x0 , u0 ) = 0.
For a linear system ẋ = Ax + Bu, we have the following cases for an equilibrium solution:
1. If A is invertible, then for every u0 , there is a unique equilibrium solution x0 = −A−1 Bu0 .
2. If A is non-invertible, then we have exactly one of the following two cases:
• There exists no equilibrium solution if B ∈
/ Range(A)
• There exist an infinite number of equilibria given by x0 = v + n, where v is any vector satisfying
Av = −Bu0 and n satifies An = 0.

1.2 Equilibrium Solutions and Stability


Definition: An equilibrium solution x0 (u0 ) is said to be stable if a perturbation ∆x (e.g., caused by an
external disturbance) is rejected by the system and x(t) → x0 as t → ∞.
Let x(t) = x0 + δx(t), where δx is assumed to be “small.” The small-perturbation dynamics about the
equilibrium x0 (u0 ) are given by  
∂f
ẋ(t) = δ ẋ = δx
∂x x0
 
The equilibrium x0 is stable if the Jacobian ∂f ∂x to be Hurwitz
x0
The “simplest” type of differential equations are linear differential equations

ẋ = Ax + Bu, A ∈ Rn×n

For linear systems, the matrix A is itself the Jacobian. An equilibrium solution is stable if and only if A is
Hurwitz; i.e., it is invertible with all of its eigenvalues in the open left half of the complex plane. Furthermore,
if A is Hurwitz, there is a unique equilibrium point for a given u0 and it is given by x0 = −A−1 Bu0 .

Definition: An equilibrium solution is said to be unstable if it is not stable. In other words, stability of a
system is a dichotomous property.

1
1.3 Modal Decomposition of Linear Systems
Consider the uncontrolled linear system ẋ = Ax. If the geometric multiplicity of A is equal to its algebraic
multiplicity, then the solution x(t) of the linear differential equation can be written in the form
N
X
x(t) = ai eλi t vi (1)
i=1

where the coefficients ai depend on the initial conditions (x(t = 0)). Notice that if λi is complex, then so is
the corresponding vi . The solution x(t) itself is real. If the algebraic multiplicity is more than the geometric
multiplicity, there is a similar expansion for x(t), except that it also involves terms of the form tk eλi t whose
corresponding vector coefficients are not eigenvectors.
In Eq. (1), if the initial condition x0 is such that ai = 0 for all i 6= k for some k and ak 6= 0, then
the solution x(t) = ak eλk t vk ; i.e., the solution x(t) continues to lie along vk for all t if it starts on vk . A
similar situation occurs if the solution starts in a plane spanned by the eigenvectors of complex conjugate
eigenvalues: it continues to remain in the eigenplane forever.
It is thus possible to view the solution of a linear system as a linear combination of its modal solutions.
In certain circumstances, it is possible to obtain the modal solutions in a rather straight-forward way. What
is important here is the stability of the individual modes: if we design the aircraft and its controllers to
stabilise the individual modes, then we would have stabilised the complete dynamics. This simplifies the
problems of stability analysis of complex systems as well as control design.

1.4 Complex Conjugate Modes


A typical second-order system is of the form ẍ + cẋ + kx = 0. Its eigenvalues are given by

−c ± c2 − 4k
λ=
2
Thus, the necessary and sufficient conditions for stability are c > 0 and k > 0. If c2 > 4k, both eigenvalues
are real; else, the eigenvalues are complex conjugate and given by

−c ± i 4k − c2
λ=
2

We define the natural frequency of a second order system as ωn = k, and its coefficient of damping as
c
ζ= . We get the following cases for different values of ζ:
2ωn
• ζ < 0: the system is unstable (c > 0)

• ζ = 1: critical damping; the two roots are equal and given by λ = −ζωn
• 0 < ζ < 1: the system is stable and p exhibits decaying oscillations. The damped natural frequency of
the oscillations is defined as ωd = ωn 1 − ζ 2 .
• ζ > 1: the system is over-damped with real, stable roots. The system does not exhibit any oscillatory
behaviour.
On the modal plane corresponding to complex conjugate eigenvalues, the dynamics are identical to spring-
mass-damper systems and we therefore extend the terminology and the associated concepts to the flight
dynamic modes as well.

2
Figure 1: Free body diagram of an aircraft.

2 Equations of Motion in the Longitudinal Plane


The aircraft has three degrees of freedom in the longitudinal plane:
• Two translational degrees of freedom described by V∞ and γ
• One rotational degree of freedom with α and q as the state variables.
From the free body diagram in Fig. 1, it follows quite readily that
1
V̇ = (T cos α − D) − g sin γ
m
1
γ̇ = (L + T sin α − mg cos γ)
mV
Recall the equations for level flight and climbing performance: these were found by setting V̇ = γ̇ = 0 (i.e.,
constant V and γ).
The rotational dynamics are given by
M
q̇ = (2)
Iyy
where M is the pitching moment which we encountered earlier in Chapter 4.
Finally, note that q = α̇ + γ̇. Thus, we get
1
α̇ = q − (L + T sin α − mg cos γ)
mV
In order to simplify the analysis, we assume that T sin α is negligible and T cos α ≈ T . This gives us the
following EoM:
1
V̇ = (T − D) − g sin γ
m
1
α̇ = q− (L − mg cos γ)
mV
M
q̇ = (3)
Iyy
1
γ̇ = (L − mg cos γ)
mV

3 Linearization
Equilibrium solutions of Eq. (4) are found by setting V̇ = γ̇ = α̇ = q̇ = 0. Thus, flight speed, flight path
angle, and the angle of attack have to be constant. These are the familiar flight conditions encountered
earlier in Chapters 2 and 3. If γ = 0, we get level flight, while a non-zero γ corresponds to trimming climb
or descent.

3
3.1 Preview of the Modal Decomposition
Before formally examining the modal structure of the longitudinal dynamics, it is instructive to look at
the time history of the longitudinal variables for a generic aircraft, found by simulating Eq. (4). The time
histories of V , α and γ are shown in Fig. 2.

(a) α (b) γ

(c) V

Figure 2: Time history of the longitudinal variables for a generic aircraft. Two complex conjugate modes
can be readily discerned from the time history.

Two modes are clearly discernible in Fig. 2:

1. A fast mode in α with a time period of approximately 5 s. This mode is well-damped and is almost
not visible in V and γ.
2. A slow mode in V and γ which is under-damped and has a time period of approximately 55 s. It is
almost missing in the α plot.
It turns out that these modes are quite typical of longitudinal dynamics. The fast mode is also called the
short period (SP) mode, while the slow mode is called the phugoid mode. We will now formally derive
analytical approximations to these modes.

3.2 Small Perturbation Model


The stability of an aircraft is as much a property of the dynamics as it is of specific trim solutions. The
simplest longitudinal trims satisfy q = 0 and are parametrized by constant values of V 0 , α0 and γ 0 . We
know from our analysis in Chapters 2 and 3 that equilbrium solutions satisfy
1
L= ρ(V 0 )2 SCL (α0 ) = W cos γ 0
2
T = D(V 0 , α0 ) + W sin γ 0

Let x0 = [V 0 , α0 , γ 0 , 0] denote the equilibrium condition. In order to linearize about a trim, we perturb the
state variables by sufficiently small values ∆V , ∆α, ∆γ and ∆q.

4
The dynamics of the perturbed variables are linear, since we assume these perturbations to be small
enough for (∆V )2 and similar higher order terms to be zero. The dynamics of the perturbed variables, also
called the linearized dynamics, are given by

CD (α0 )
   
∆V̇ g 0 ∆V 0
= − ∆γ + 2 cos γ + 2k cos γ C Lα ∆α
V0 V0 CL (α0 ) V0
 
g ∆V CLα
∆γ̇ = 0
sin γ 0 ∆γ + 2 cos γ 0 0 + ∆α
V V CL
∆α̇ = ∆q − ∆γ̇
Mα Mα̇ Mq
∆q̇ = ∆α + ∆α̇ + ∆q
Iy Iy Iy

where Mα = ∂M/∂α and so on. These are the same derivatives that we encountered in Chapter 4.
Based on numerical results, and arguments based on physics, it is appropriate to hypothesise that there
exist two modes:

• A fast mode which can be approximated from the dynamics of ∆α and ∆q alone, and
• A slow mode which can be approximated from ∆V and ∆γ alone.

3.3 Short-Period (SP) Dynamics


Short period dynamics represent fast, but well-damped oscillations in pitch. The time scale of the oscillations
is fast enough so that the translational variables are more or less unchanged while the SP oscillations settle
down. Thus, we can study the SP dynamics by simply ignoring the dynamics of V and γ. We can thus write
g CLα
∆α̇ = ∆q − ∆α
V0 CL
Mα Mα̇ Mq
∆q̇ = ∆α + ∆α̇ + ∆q
Iy Iy Iy

These equations can be consolidated into a single second order ODE of the form
2
∆α̈ + 2ζSP ωSP ∆α̇ + ωSP ∆α

where
 
2 Mα Mq g CLα
ωSP = − +
Iy Iy V0 CL
 
Mα̇ + Mq g CLα
2ζSP ωSP = − + (4)
Iy V0 CL

It turns out that the stability of the short period depends largely on ensuring that Mα < 0 and Mq < 0. The
other terms in Eq. (4) make secondary contributions to the stability, although their effect can be significant
at high speeds (since V0 CL ∝ 1/V0 ).
It is worth noting that the modal frequency increases quite rapidly with V0 , as does the damping ratio.
At low speeds, and high angles of attack, the SP mode is affected significantly by aerodynamic nonlinearities,
especially stall.

3.4 Phugoid (P) Dynamics


The phugoid mode consists of poorly damped, long-period oscillations of the point-mass aircraft. We can
derive an approximation for the phugoid mode along the lines of the derivation for the SP mode. We

5
start by assuming that ∆α = ∆q = 0. Furthermore, we normalize the perturbation ∆V by V 0 to keep it
non-dimensional. The phugoid dynamics are then given by
" #
CD (α0 ) ∆V
  0

∆V̇ g −2 cos γ −1
0) V 0
V0 = 0 CL (α (5)
∆γ̇ V 2 cos γ 0 sin γ 0 ∆γ

It follows that the natural frequency of the phugoid mode is given by


 2
CD (α0 )
 g 2  
g
ωP2 = 2 cos γ 0
1 − sin γ 0
≈ 2
V0 CL (α0 ) V0

Unlike ωSP , the natural frequency of phugoid ωP reduces with increasing flight speed and with increasing
γ0 . The time constant of the phugoid mode is typically several tens of seconds.
The damping ratio (with γ 0 ≈ 0) is given by
 
1 CD
ζP = √
2 CL α0
The damping ratio decreases with increasing aerodynamic efficiency, and it typically around 0.04 − 0.07.
Interestingly enough, the damping ratio is minimum at the speed for minimum drag, and increases on either
side of it.
The formulae given here suggest that the phugoid mode should be generically stable, but it is not. The
formulae that we have derived are, thus, strictly approximate in nature and require considerable refinement.
Accurate analytical prediction of phugoid is beyond the scope of this course. However, since the phugoid
has a large time constant and fairly small amplitudes, it is not uncommon for the phugoid mode to be left
uncompensated by active control, especially in combat aircraft.

6
Chapter 7
Equations of Motion
Aditya Paranjape

1 Translational Equations of Motion

Figure 1: The total momentum is found by writing the momentum of the element dm and then summing
over the complete aircraft.

Consider the small element with mass dm with vector r̄ from the origin of the body-fixed B-frame, as
shown in Fig. 1. Then, the total velocity of the element is given by

V̄T = V̄ + ω̄ × r̄

The linear momentum of the element is given by

dp̄ = dm V̄T = dmV̄ + ω̄ × (r̄ dm)

Hence, the total momentum of the body is given by


I I
p̄ = dp̄ = mV̄ + ω × r̄ dm
H
We assume that the origin of B-frame is located at the centre of mass; i.e., r̄ dm = 0. Then, using Newton’s
second law, we get    
dp̄ d(mV̄ )
F̄ = =⇒ F̄ =
dt I dt I
Assuming that the mass m is constant, i.e., the rate of fuel burn is negligibly slow, we get
   
dV̄ dV̄ 
m =m + m ω̄ × V̄ = F̄ (1)
dt I dt B

We express the vectors are in the body (B) frame:


• V̄ = (u, v, w)
• ω̄ = (p, q, r)
• Forces: (X, Y, Z)

1
Substituting into Eq. (1) This gives us the following equations:
X
u̇ + (qw − rv) =
m
Y
v̇ + (ru − pw) = (2)
m
Z
ẇ + (pv − qu) =
m
The forces X, Y Z derive from three sources: propulsive, aerodynamic and gravity.

1.1 Propulsive
Throughout this course, we have assumed that the line of thrust coincides with the xB axis, so that X prop =
T , and Y prop = Z prop = 0.

1.2 Aerodynamic
Lift and drag are the two primary aerodynamic forces. However, their components L and D are written with
respect to the velocity vector. In particlar, they are written in what are called as the wind axes whose axes
are defined as follows:
• xw axis: along V̄∞
• zw axis: along the cross product V̄∞ × ȳB
• yw axis: along the cross product z̄w × x̄w
The rotation from the wind frame to the body frame is achieved by two rotations:
• Rotation about zw by −β to bring the xw axis in the plane of symmetry of the aircraft. The wind axis
frame is then tranformed into a new frame which we denote by W 0 . Note that zw = zw0 .
• Rotation about the y axis of W’ frame by an angle α.
In the wind frame frame, the aerodynamic forces are given by
 
−D
F̄aero,W =  0 
−L
Let R3 (−β) denote the rotation from W to W 0 (the subscript denotes the axis of rotation). In particular
given a vector r̄, its components in W 0 and W are related by r̄w0 = R3 (−β)r̄w . With this notation, we get:
  
cos α 0 − sin α cos β − sin β 0
F̄aero,B =  0 1 0   sin β cos β 0  F̄aero,W
sin α 0 cos α 0 0 1
 aero    
X cos α cos β − cos α sin β − sin α −D
=⇒  Y aero  =  sin β cos β 0  0 
aero
Z sin α cos β − sin α sin β cos α −L
 
L sin α − D cos α cos β
=  −D sin β 
−L cos α − D sin α cos β
Notice that α and β are related to u, v w by the following relations:
w v
tan α = , sin β = (3)
u V∞

2
This can be verified by using the above rotation matrix from W to B: notice that the velocity vector is
[V∞ , 0, 0] in the wind frame. Thus,
u = V∞ cos α cos β, v = V∞ sin β, w = V∞ cos β sin α
All of the forces that we have seen so far depend only on u, v, w. The next force, gravity, will require that
we introduce additional states beyond the velocity and angular velocity components.

1.3 Gravity
The gravitational acceleration is defined by its components [0, 0 g] in the inertial frame (I). The transfor-
mation from I to the body frame B involves a sequence of three rotations:
1. Rotation from I to I1 : about zI by an angle ψ
2. Rotation from I1 to I2 : about yI1 by an angle θ
3. Rotation from I2 to B: about xI2 by an angle φ
It follows that
ḡB = R1 (φ) R2 (θ) R3 (ψ) ḡI
    
1 0 0 cos θ 0 − sin θ cos ψ sin ψ 0 0
=  0 cos φ sin φ   0 1 0   − sin ψ cos ψ 0  0 
0 − sin φ cos φ sin θ 0 cos θ 0 0 1 g
 
−g sin θ
=  g sin φ cos θ 
g cos φ cos θ
Thus, the components of the weight in the body frame are given by
 grav   
X −mg sin θ
 Y grav  =  mg sin φ cos θ  (4)
Z grav mg cos φ cos θ

2 Euler Angle Rates


We have introduced two new states: φ and θ. Notice that ψ does not affect the translational equations of
motion in any way, and therefore ψ̇ is decoupled from the rest of the dynamical equations.
First, we note that ψ̇, θ̇ and φ̇ are defined in three different frames: I, I1 and I2 , respectively. The
angular velocities defined in the three frames are thus [0, 0, ψ̇]I , [0, θ̇, 0]I1 and [φ̇, 0, 0]I2 . Thus, the body
axis components of the sum of these vectors are given by
       
p φ̇ 0 0
 q  = R1 (φ)  0  + R1 (φ)R2 (θ)  θ̇  + R1 (φ)R2 (θ)R3 (ψ)  0 
r 0 0 ψ̇
  
1 0 − sin θ φ̇
=  0 cos φ cos θ sin φ   θ̇ 
0 − sin φ cos θ cos φ ψ̇
Before inverting the above matrix, note that its determinant is cos θ. Thus, the determinant is invertible
only when |θ| < π/2. In that case, inversion yields
φ̇ = p + q tan θ sin φ + r tan θ cos φ
θ̇ = q cos φ − r sin φ
ψ̇ = sec θ(q sin φ + r cos φ)

3
3 Equations of Motion: Rotational
Consider Fig. 1 again. The angular momentum of the element dm is given by

dh̄ = r̄ × dp̄ = dm r̄ × V̄ + ω̄ × r̄

The total angular momentum is then given by


I I  I 
h̄ = dh̄ = dm r̄ × V̄ − dm (r̄×)2 ω̄

The cross product r̄ × ω̄ can be written as


 
0 −r3 r2
r̄ × ω̄ =  r3 0 −r1  ω̄ = S(r̄)ω̄
−r2 r1 0

Similarly, we can write:

−(r22 + r32 )
 
r2 r1 r3 r1
(r̄×)2 ω̄ = S(r̄)S(r̄)ω̄ =  r1 r2 −(r32 + r12 ) r2 r3  ω̄
r1 r3 r2 r3 −(r12 + r22 )

The moments of inertia are defined by


I
r22 + r32 dm

Principal moments : I11 =
I
Cross moments : I12 = − (r1 r2 ) dm

Define the matrix J = {Iij } is given by I


J =− (S(r̄))2 dm

The matrix J is also called the moment of inertia matrix (or moment of inertia tensor). With this notation,
the total angular momentum can be written as

h̄ = J ω̄ (5)

Using Newton’s laws of motion, and with M̄ denoting the vector angular moments, we get the general
equations of motion:
ω̄˙ = J −1 −S(ω̄)J ω̄ + M̄


where ω̄ = (p, q, r) and M̄ = (L, M, N ). Assume that the aircraft is symmetric about the xB − zB plane.
This implies that I12 = I23 = 0. Furthermore, suppose we choose the principal axes so that I13 = 0 as well.
With these simplifications, we get
 
I2 − I3 L
ṗ = qr + (6)
I1 I1
 
I3 − I1 M
q̇ = rp + (7)
I2 I2
 
I1 − I2 N
ṙ = pq + (8)
I3 I3

where I1 = I11 and so on.

4
4 Consolidated Equations of Motion
The equations of motion of a rigid aircraft can be put together from the derivations in this chapter:
X
u̇ = rv − qw +
m
Y
v̇ = pw − ru +
m
Z
ẇ = qu − pv +
 m
I2 − I3 L
ṗ = qr +
I1 I1
 
I3 − I1 M
q̇ = rp +
I2 I2
 
I1 − I2 N
ṙ = pq +
I3 I3
φ̇ = p + q tan θ sin φ + r tan θ cos φ
θ̇ = q cos φ − r sin φ
ψ̇ = sec θ(q sin φ + r cos φ)

Notice that the none of the equations depend on ψ; thus, the equation for ψ̇ is dropped from this list while
studying the stability of an aircraft. The equation should be retained, however, when performance is to be
studied.

4.1 Translational Equations Rewritten in the Wind Frame


It is generally useful to replace the equations for u, v and w with the corresponding equations in V∞ , α and
β since it is the latter set that is of relevance from the point of view of physics.
We start by noting that

u = V∞ cos α cos β
=⇒ u̇ = V̇∞ cos α cos β − V∞ sin α cos β α̇ − V∞ cos α sin β β̇

Likewise,

v̇ = V̇∞ sin β − V∞ cos β β̇


ẇ = V̇∞ sin α cos β + V∞ cos α cos β α̇ − V∞ sin α sin β β̇

Substituting into the above equations of motion and replacing u, v and w in terms of V∞ , α and β yields a
familiar-looking set of equations:
1
V̇∞ = (T cos α cos β − D − W sin γ)
m
1
α̇ = q − tan β(p cos α + r sin α) − (T sin α + L − W cos γ cos µ)
mV∞ cos β
1
β̇ = p sin α − r cos α + (−T cos α sin β + Y + W cos γ sin µ)
mV∞
where
• Y is the aerodynamic side-force acting on the aircraft.
• µ is the wind axis roll angle which we saw earlier in Chapter 5.

5
It remains to describe γ and µ in terms of α, β, θ and φ. Notice that the γ and µ terms arise out of an
attempt to resolve the gravity vector in the wind frame. The Euler angles for the wind frame are χ (velocity
heading), γ and µ. Therefore, we simply transform the gravity vector from the body frame, which we have
already done earlier, to the wind frame:
     
− sin γ cos β sin β 0 cos α 0 sin α − sin θ
 cos γ sin µ  =  − sin β cos β 0   0 1 0   cos θ sin φ 
cos γ cos µ 0 0 1 − sin α 0 cos α cos θ cos φ
  
cos β cos α sin β cos β sin α − sin θ
=  − sin β cos α cos β − sin β sin α   cos θ sin φ 
− sin α 0 cos α cos θ cos φ

This gives

sin γ = sin θ cos β cos α − sin β cos θ sin φ − sin α cos β cos θ cos φ
cos γ sin µ = cos α sin β sin θ + cos β cos θ sin φ − sin α sin β cos θ cos φ
cos γ cos µ = cos α cos θ cos φ + sin α sin θ

5 Special Solutions
It is a fact that most standard aircraft missions can be described as a concatenation of equilibrium solutions
of the 8 dynamical equations described above. It is for this reason that equilibrium solutions are important.
In the preceding chapters, we saw several flight conditions such as level flight, level turn and the steady
sideslip maneuver. Each of these can be found as an equilibrium solution of the above equations of motion.
The sideslip maneuver may also have a non-equlibrium element.
There are three types of variables that make up a maneuver:
1. The prescribed variables: the values of these variables are given. For equilibrium fligth, these are
constant.

2. To-be-determined (TBD) variables: TBD variables are state variables whose values are not prescribed
a-priori. Rather, their values are found by solving the equations of motion.
3. Control variables: these are the control inputs, i.e., thrust, elevator, aileron and the rudder deflections
in a conventional aircraft.

In general, note that at most 4 variables can be prescribed, corresponding to the four control inputs, except
in the wings-level condition where lateral-directional variables vanish automatically and may be treated as
being prescribed zero values.

5.1 Wings-Level Flight


In wings level flight, the distribution of the variables among the three sets is as follows:
• Prescribed zero variables: β = p = r = φ = 0
• Prescribed non-zero quantities: V∞ , γ (i.e., θ − α)

• TBD variables: θ, α
• Control variables: δa = δr = 0; thrust and elevator need to be determined.

6
5.2 Level Turn
The equation for ψ̇ enters the picture for turns: the equilibrium turn rate may be directly equated to ψ̇.
• Prescribed zero variables: β = 0, γ = 0
• Prescribed non-zero quantities: V∞ , turn rate ψ̇
• TBD variables: θ, α, p, q, r, φ
Under equilibrium conditions, we note the following connection between the various quantities which helps
with the calculation of the TBD variables:

p = −ψ̇ sin θ, q = ψ̇ cos θ sin φ, r = ψ̇ cos θ cos φ

This equation ensures that the Euler angle rates φ̇ = θ̇ = 0.


An important point to note here comes from the α̇ equation. When we studied level turns in Chapter 3,
we set L cos µ = W . Can you reconcile this with α̇ = 0, i.e., L − W cos µ = mV∞ q?

5.3 Steady Sideslip Maneuver


A cross-slipping maneuver is performed typically when an aircraft lands. It may be used to slow the aircraft
during descent as well; in fact, it is used routinely by gliders. The distribution of the variables is as follows:
• Prescribed (non-zero): β, γ
• Prescribed (possibly non-equilibrium): V∞ (t); either a constant speed would be prescribed or a rate
V̇∞ . The thrust T can be adjusted to achieve this profile.
• TBD: φ, θ, α
• TBD (but zero): p = q = r = 0
The control variables and the TBD variables are determined as follows:
1. Aileron and rudder inputs by setting Cl = Cn = 0.
2. The aileron and rudder inputs, and the sideslip, generate a non-zero sideforce Y . Thus, the wind
axis roll angle µ must take on a value which ensures equilibrium in β, i.e., β̇ = 0. In fact, ignoring
T cos α sin β allows us to determine µ quite readily.
3. The angle of attack takes on a value found from L = W cos γ cos µ.
4. The thrust can be found from V (t): T = D + W sin γ + mV̇ (t).
5. The pitch angle achieved in the process can be found from the γ and µ equations above.

5.4 Level Acceleration


A wings-level acceleration at constant altitude is an example of a non-equilibrium maneuver. Consider an
aircraft accelerating at a fixed rate V̇d at a fixed altitude (γ = 0). The maneuver can be profiled as follows:
• Prescribed variables: V̇d , γ = 0
• Prescribed/TBD variables: β = p = r = φ = 0
• TBD: α, q, θ
• Control variables: δa = δr = 0; T, δe to be determined.

7
The thrust depends only on the instantaneous drag and acceleration:
1
T = mV̇d + ρV 2 SCD (α)
2
Notice that α changes at each instant to ensure that γ = 0:
 
2W 2W
CL (α) = =⇒ α(t) = CL−1 − CL0
ρV 2 S ρV 2 (t)S

Notice that α is time-varying since V is time-varying. Once α(t) is known, we can find α̇(t). The pitch rate
q(t) is then given by
−1
q(t) = α̇(t) =⇒ δe (t) = (M0 + Mα α(t) + Mq q(t) + Mα̇ α̇(t))
Mδe

Note that the variables calculated here represent the ideal trajectory; the actual trajectory would feature
perturbations about the ideal values which subsequently need to be compensated for by automatic control
systems.

6 Linearization and Decoupling


The flight dynamic equations derived above form a set of nonlinear differential equations. As we saw in
Chapter 6, the stability of the dynamics is not really a global property - rather it is a local property of
equilibrium solutions. Level flight equilibria are of primary interest to us. The results established for level
flight are generally valid for turning flight as well, although nonlinearities in the parent dynamics do bring
about some qualitative and quantitative changes in stability.
We will show here that the dynamics obtained by linearizing the equations of motion decouple of the
longitudinal states [u, w, q, θ] and the lateral-directional states [v, p, r, φ] from each other. Once we have
linearized the dynamics, we will replace [u, v, w] with [V∞ , α, β].
A level flight equilibrium is specified by prescribing the values of u0 , w0 and θ0 . All other variables
are zero at equilibrium. Moreover, if the aircraft is symmetric, the only force and moment derivatives that
matter are listed in Table 6. Note that CXθ , CYφ and CZθ arise exclusively due to the weight of the aircraft.
The derivatives Xq and Zq are almost always non-zero, but can be ignored safely in low angle of attack flight.

2
Table 1: Aerodynamic derivatives; note that q∞ = 0.5 ρV∞ .
Force/Moment Important Stability Derivatives
X = q∞ SCX CXu , CXw , CXq , CXθ
Y = q∞ SCY CYv , CYp , CYr , CYφ
Z = q∞ SCZ CZu , CZw , CZq , CZθ
L = q∞ SbCl Clv , Clp , Clr
M = q∞ ScCm Cmu , Cmw , Cmq , Cmẇ
N = q∞ SbCn Cnv , Cnp , Cnr

Let ∆u denote the perturbation in u, and likewise for other quantities. Substituting into the equations
of motion, we get two decoupled sets of linearized dynamics:

8
1. Longitudinal dynamics
Xu Xw Xθ
∆u̇ = −w0 ∆q + ∆u + ∆w + ∆θ
m m m
Zu Zw Zθ
∆ẇ = u0 ∆q + ∆u + ∆w + ∆θ
m m m
1
∆q̇ = (Mu ∆u + Mw ∆w + Mẇ ∆ẇ + Mq ∆q) (9)
I2
∆θ̇ = ∆q

The reader should verify, as an exercise, that these equations can be transformed into the wind-axis
equations which were derived in Chapter 6. Note that Xθ = −W cos θ and Zθ = −W sin θ. Finally,
the term Mẇ is the equivalent of Mα̇ .
2. Lateral-directional dynamics
Yv Yp Yr Yφ
∆v̇ = w0 ∆p − u0 ∆r + ∆v + ∆p + ∆r + ∆φ
m p m m
1
∆ṗ = (Lv ∆v + Lp ∆p + Lr ∆r)
I1
1
∆ṙ = (Nv ∆v + Np ∆p + Nr ∆r) (10)
I3
∆φ̇ = ∆p + tan θ0 ∆r

These equations can be written in the wind-axis frame quite readily. In fact, the only equation that
changes is for ∆v, with ∆v = V∞ ∆β. This gives us
1 g
∆β̇ = α0 ∆p − ∆r + (Yβ ∆β + Yp ∆p + Yr ∆r) + cos α0 ∆φ
mV∞ V∞
1
∆ṗ = (Lβ ∆β + Lp ∆p + Lr ∆r)
I1
1
∆ṙ = (Nβ ∆β + Np ∆p + Nr ∆r) (11)
I3
∆φ̇ = ∆p + tan α0 ∆r

Since the linearization is about level flight, we replaced Yφ /m = g cos θ0 and we additionally assumed
that u0 ≈ V∞ .
The stability of the complete dynamics about level flight can thus be determined by analysing the
longitudinal and the lateral-directional dynamics independently. In the next chapter, we will analyse the
lateral-directional dynamics along the lines of our analysis for the longitudinal dynamics in Chapter 5.

9
Appendix: Tranformations between Coordinate Frames

Figure 2: Illustration: Rotation about x axis

Transformations between coordinate frames are achieved through a sequential set of rotations. Consider
a single rotation through an angle δ from frame ’A’ to frame ‘B’ about axis 1 as shown in Fig. 2. Let x̄a and
x̄b denote the coordinates of a vector x̄ in frames A and B, respectively.
 
1 0 0
x̄b =  0 cos δ sin δ  x̄a
0 − sin δ cos δ

Denote this transformation by


1
x̄b = RBA (δ)x̄a
For the ith axis:
i i
RBA (δ) = RAB (−δ)
i.e., the transpose of a rotation matrix is also its inverse.
We can now give formulae for coordinate transformation under single rotations:
• Rotation about axis 1:  
1 0 0
1
RBA (δ) = 0 cos δ sin δ 
0 − sin δ cos δ

• Rotation about axis 2:  


cos δ 0 − sin δ
2
RBA (δ) =  0 1 0 
sin δ 0 cos δ

• Rotation about axis 3:  


cos δ sin δ 0
3
RBA (δ) =  − sin δ cos δ 0 
0 0 1

10
Chapter 8
Lateral-Directional Flight Dynamics
Aditya Paranjape

The linearized equations of motion of the lateral-directional dynamics were derived in the previous chap-
ter:
1 g
∆β̇ = α0 ∆p − ∆r + (Yβ ∆β + Yp ∆p + Yr ∆r) + cos α0 ∆φ
mV∞ V∞
1
∆ṗ = (Lβ ∆β + Lp ∆p + Lr ∆r)
Ix
1
∆ṙ = (Nβ ∆β + Np ∆p + Nr ∆r) (1)
Iz
∆φ̇ = ∆p + tan α0 ∆r

We would like to understand the modal decomposition of the lateral-directional (L-D) dynamics along the
lines of the longitudinal dynamics. To simplify our analysis, we will make the following assumptions:
1. α0 ≈ 0. While this assumption limits the validity of our results to low angles of attack, the analysis
nevertheless captures the essential properties of the L-D modes.
2. Yp = Yr = Np = 0. These assumptions are based on physics and will not be discussed any further.
We will thus work with the following linearized equations:
Yβ g
∆β̇ = ∆β − ∆r + ∆φ
mV∞ V∞
1
∆ṗ = (Lβ ∆β + Lp ∆p + Lr ∆r)
Ix
1
∆ṙ = (Nβ ∆β + Nr ∆r) (2)
Iz
∆φ̇ = ∆p

Our first observation is that, unlike the longitudinal modes, there is no obvious way to identify the fast and
the slow modes by isolating the translational and the rotational degrees of freedom. Instead, the way to
arrive at a modal decomposition is to analyse numerical results, postulate upon the nature of the modes and
verify that the postulates are indeed correct.
It turns out that the L-D dynamics decompose into three modes which are visible in the root locus plot
for the F/A-18 HARV, shown in Fig. 1:
• Roll: fast and non-oscillatory; almost entirely p; period T ∼ O(0.1 s)
• Dutch roll: moderately fast; β + r; period T ∼ O(1 s)
• Spiral: slow and non-oscillatory; mostly r; period T ∼ O(10 s)

1
Figure 1: Low α root locus of the L-D dynamics of the F/A-18 HARV

1 Roll Mode
The roll dynamics are given by
Lp 1
∆ṗ = ∆p + (Lr ∆r + Lβ ∆β)
Ix Ix
The roll eigenvalue is given by
Lp
λR =
Ix
with the accompanying stability condition Lp < 0. Clearly, the roll mode is stable as long as the aircraft is
in the pre-stall regime.
Due to the fast roll dynamics, the roll rate converges rapidly to a value found by setting ṗ ≈ 0. This
value is called the static residual of the roll rate and is given by
1
∆ps = − (Lr ∆r + Lβ ∆β) (3)
Lp
The rest of the states “see” this value of roll rate in the course of their evolution.

2 Dutch Roll
The Dutch roll mode predominantly involves β and r. Its dynamics are given by
Yβ g
∆β̇ = −∆r + ∆β + ∆φ
mV V
Nr Nβ
∆ṙ = ∆r + ∆β
Iz Iz
The characteristic equation of the Dutch roll dynamics works out to
   
Yβ Nr 1 Yβ
λ2 − + λ+ Nr + Nβ = 0
mV Iz Iz mV
| {z }
2
ωDR

2
Comparing with the standard spring-mass-damper system of the form ẍ + cẋ + kx = 0, we note the following:
Nr Yβ
• The damping constant c is given by + . Generally, Nr as well as Yβ are negative as long as the
Iz mV
vertical tail is appropriately sized, and the Dutch roll mode damping is ideally not expected to present
any problems at low angles of attack.
Yβ Nr
• The stiffness depends on Nβ + . Clearly, Nβ > 0 is a sufficient condition for ensuring that the
mV
Dutch roll mode is stable, provided both Yβ < 0 and Nr < 0.
The Dutch roll mode manifests itself in the form of oscillations in yaw, as shown in Fig. 2. The yawing
motion is accompanied by a roll rate whose value is equal to the static residual in Eq. (3).

Figure 2: The Dutch roll mode manifests itself in the form of oscillations in yaw. Source:padpilot.eu

Just as in the case of the roll mode, the Dutch roll dynamics are not entirely self-contained. Therefore,
the yaw rate and the sideslip converge to quasi-statically varying values ∆βs and ∆rs which depend on ∆φ,
and are found by solving

 
g
 mV −1  ∆βs
   
− ∆φ
= V
(4)
  
   
 Nβ Nr  ∆r
s 0
Iz Iz
You should check that the static residuals equal
 
−1 g Nr
∆βs = 2 ∆φ
ωDR V Iz
 
1 g Nβ
∆rs = 2 ∆φ
ωDR V Iz
2
where ωDR is the natural frequency of the Dutch roll mode, given by the determinant of the matrix on the
LHS of Eq. (4):  
2 1 Yβ
ωDR = Nr + Nβ
Iz mV

3
3 Spiral Mode
This is the slowest of modes, and its dynamics are based entirely off the static residuals found above. The
spiral dynamics are given by
∆φ̇ = ∆p ≡ ∆ps
where ∆ps is also found from the static residuals ∆rs and ∆βs :
Lr Lβ
∆ps = − ∆rs − ∆βs
Lp Lp
1 1 g
=⇒ ∆φ̇ = 2 (Lβ Nr − Lr Nβ ) ∆φ
ωDR Izz Lp V
The stability condition for the spiral mode is:
Lβ Nr − Lr Nβ
<0
Lp
For an aircraft with a stable roll mode, Lp < 0. Therefore, for spiral stability,

(Lβ Nr − Lr Nβ ) > 0

The sign of Lr is difficult to determine precisely. It gets a positive contribution from the tail and a negative
contribution from the wing. On the other hand, Nr < 0 for the most part; thus, it is important to have
Lβ < 0 for stabilizing the spiral mode.
Figure 3 shows the response of an aircraft if the spiral mode is unstable. An instability in the spiral mode
leads to an uncommanded turn which becomes progressively steeper with time. Depending on Lβ and Nβ ,
the aircraft may either turn steeply with little sideslip, or get into a relatively gentler turn but with a large
sidelsip.

Figure 3: An unstable spiral mode puts the aircraft into an uncommanded turn.
Source: people.rit.edu/pnveme/EMEM682n/DynamicStab/roots lateral.html

4
4 Lateral-Directional Modes at Higher Angles of Attack

(a) Low α

(b) High α

Figure 4: The L-D modes of the F/A-18 HARV for the entire range of pre-stall α.

Figure 4 shows the lateral-directional modes of the F/A-18 HARV with increasing angle of attack. These
modes were plotted without assuming α0 = 0, and α0 was allowed to increase to near-stall values.
The roll mode in this aircraft is slower than in larger cargo and passenger aircraft. In addition, the roll
mode tends to slow down with the angle of attack. This causes it to merge with the spiral mode and produce
an oscillatory mode called the lateral phugoid. Despite its name, the lateral phugoid is very much unlike its
longitudinal counterpart: it is well damped and is relatively faster, with a time constant of about 10 s.
The Dutch roll mode is quite poorly damped, with a coefficient of damping less than 0.2. The Dutch roll
damping can be improved by increasing the size of the vertical tail, which increases Nr as well as Yβ .

5
Chapter 9
Introduction to Manoeuvres and Aircraft Control
Aditya Paranjape

The course so far has covered aircraft performance and stability, and we have seen a hierarchical connec-
tion between them.
Perforamance problems assume that the aircraft can attain and fly at the angle of attack and the bank
angle that arise as solutions for various flight conditions such as maximum speed, fastest sustained turn, etc.
The problem of attaining and maintaining the desired values of angle of attack and other flight parameters
was addressed partially as a “trim and stability” problem. Among other things, we also saw that the
performance of a stable airframe may not be as good as that of an unstable airframe (e.g., recall the
discussion on the lift produced by stable and unstable aircraft). This leads designers to opt for airframes
that are inherently unstable, and these airframes need to be stabilized artificially using control systems to
make the aircraft safe and flight-worthy.

1 Manoeuvres and Pseudo Steady State Dynamics


We looked at level acceleration in Chapter as an example of non-equilibrium manoeuvres. The example
had one special feature: every state that was not specified could be calculated quite easily because of the
longitudinal nature of the manoeuvre. In general, it is not as easy to do so that for general manoeuvres.
As an example, consider a wind axis roll manoeuvre, also called an a velocity vector roll (VVR). In
its simplest form, only the desired roll rate is specified with β = 0, while all other variables need to be
determined from the equations of motion. Determining the exact time history of the state variables “by
hand” is almost impossible, since V , θ and φ are time-varying.
One way to analyse such manoeuvres is to work with a restricted set of equations of motion. For instance,
if the duration of the manoeuvre is very short (less than 5 seconds), then one may assume that the flight
speed is almost constant. Furthermore, if the manoeuvre takes place at reasonable high flight speeds, then
the ratio (g/V ) can be ignored. A cursory look at the equations of motion shows that ignoring g/V comes
with a beneficial collateral consequence: the dynamics of the Euler angles θ and φ get decoupled from the
other equations of motion. As such, we don’t expect these variables to be in equilibrium and their decoupling
opens up the possibility of simplifying the analysis considerably. Let α0 denote the initial angle of attack,
before the manoeuvre is commenced, and let δe0 the initial elevator deflection. This gives:

1
α̇ = q − tan β(p cos α + r sin α) − ρV 2 S(CL − CL0 )
2
Yβ β + Yδa δa + Yδr δr
β̇ = p sin α − r cos α +
mV
Iy − Iz Lp p + Lr r + Lβ β + Lδa δa + Lδr δr
ṗ = qr + (1)
Ix Ix
Iz − Ix Mα (α − α0 ) + (Mq + Mα̇ )q + Mδe (δe − δe0 )
q̇ = rp +
Iy Iy
Ix − Iy Nr r + Nβ β + Nδr δr
ṙ = pq +
Iz Iz

1
The equations stated in in Eq. (1) are called pseudo steady state (PSS) equations and are useful in
studying manoeuvres such as VVR. In fact, the VVR is an equilibrium solution of the above equation.
Prescribe α = α0 and q = β = 0. Then, for equilibrium, we need:
   −1    
δa Lδa Lδr Lp p + Lr r −1 Nδr (Lp + Lr r) − Lδr Nr r
ṗ = ṙ = 0 : =− =
δr 0 Nδr Nr r Lδa Nδr Lδa Nr r
Yδa δa + Yδr δr
β̇ = 0 : p sin α − r cos α + =0 (2)
mV
Iz − Ix
q̇ = 0 : δe = δe0 − rp
Iy
It is amply clear that Eq. (2) can be solved quite easily by hand. Choose p as the free variable. Next, find
δa and δr in terms of p and r. Then, substitute into β̇ = 0 to get a single equation connecting p and r. In
one go, this yields r, δa and δr . The elevator deflection is then found from q̇ = 0.
The stability analysis that was performed for the complete equations of motion can be performed for the
PSS equations as well, with the equilibrium defined appropriately. If level flight were to be analysed using
PSS, it is clear that we would be able to successfully isolate the short period, roll and Dutch roll modes quite
accurately.
It is perhaps not surprising that the equilibrium solutions of the complete set of equations from Chapter 7
are also equilibrium solutions of the PSS equations. However, the PSS equations have additional equilibrium
solutions which are not equilibrium solutions of the complete set of equations. The VVR is an example of
such solutions. Taking this idea a step further, one can conceive other subsets of the original dynamics and
the equilibrium solutions of these subsets can be viewed as manoeuvres in their own right. We will return
to this idea again later in this chapter.

2 Basic Elements of a Control System


An automatic control system is used on aircraft for several reasons:
• Reduce pilot workload by automating routine manoeuvres such as wings-level flight and turns.
• Improve the handling qualities by augmenting the eigenvalues of the flight dynamics.
• Eliminate pilot-induced oscillations by filtering the pilot inputs appropriately.
• Help the aircraft achieve the desired dynamical response, including for the purpose of creating control
system testbeds.

Figure 1: A-380 cockpit and a schematic of the typical flight control system. Source: Wikipedia

Figure 1 shows the schematic representation of an advanced flight control system that is typical of modern
commercial and combat aircraft. It consists of four primary components:

2
• Avionics: control column, autopilot console and the diplays.
• Flight control computer.
• Sensors: sense flight parameters and health monitoring
• Actuators: deflect control surfaces and actuate the engines
The flight control computer is the central node in this system - it provides position and rate commands
to the actuators based on the desired flight profile provided by the pilot (either through manual control
yoke deflection or through the autopilot) and the information it derives from the sensors. The control law is
implemented on the flight control computer as a software and can be of one of two types:
1. Completely closed-loop: there is no direct connection between the pilot’s controls and the actuators.
Rather, the pilot’s stick movements are sent to the flight control computer, and the computer sends
the appropriate commands to the actuators on the control surfaces and in the engines. This control
philosophy is used, for example, on all Airbus aircraft starting with the Airbus A320. It eliminates the
need for mechanical linkages between the control column and the actuators, but renders the aircraft
vulnerable to fatal crashes in the event of a computer malfunction.
2. Combination of closed-loop and open-loop: the control column has a direct link to the actuator position
(typically via a power amplifier). However, stabilizing actuator commands are added on top of the
pilot input by the flight control system. Most Boeing aircraft employ this control design philosophy.
While this design offers the pilot the possibility of taking complete manual control of the aircraft, it
requires additional mechanical components to be built into the aircraft, thereby making the system
more complex.

3 Types of Flight Control Problems


Consider the general nonlinear system
ẋ = f (x, u), y = h(x), z = g(x)
where y denotes the variables which are to be controlled and z denotes the measured output signals. Note
that x ∈ Rn , while z ∈ Rm for some m ≤ n; i.e., not all states are necessarily measured.
Depending on the nature of the problem and the sensed variables, control problems can be classified as
follows.
• State versus output feedback problems: when z = g(x) = x, all of the aircraft states are available
to the flight computer in real time, and it can use them for control. This is called full state feedback. In
contrast, when z ⊂ x, i.e., only some of the state variables are measured, the resulting control problem
is called an output feedback problem. Output feedback problems are generally harder to solve, and the
flight computer will usually make for the limited information by using algorithms which help it predict
the state variables that are not measured.
• Regulation versus tracking problems: a tracking problem is one where the output y(t) is required
to track a reference signal r(t). In such cases, it suffices to ensure that the remaining state variables
do not assume unreasonable values; their exact values are not important subject to the qualification
of boundedness within a reasonable range. On the other hand, if the control objective is to ensure
that y(t) tends to zero asymptotically, then the control problem is referred to as a regulation problem.
Regulation problems are generally easier to solve than tracking problems, although it is usually possible
to convert a tracking problem into a regulation problem by appropriate coordinate transformations.
The design of control systems which fall into either of these categories forms a complete subject in its
own right, and will not be pursued here. Rather, we will look into specific problems that control designers
encounter, and which have their roots in flight dynamics and in the limitations of flight systems. We will
assume that linear control techniques are employed for control design.

3
4 Time Scale Separation
The flight dynamics of most aircraft can be decomposed naturally into simpler blocks, each of which has a dis-
tinct time-scale of its own. This decomposition is akin to the modal decomposition, except that longitudinal-
lateral decoupling is avoided when aircraft are designed for high-rate manoeuvres, including turns and rolls.
The decomposition used for the purpose of control follows a task-based hierarchy:
1. Fast rotational dynamics: p, q, r
2. Acceleration of the point mass: α, β, µ; also engine state η
3. Velocity states: V∞ , γ, χ
4. Position variables: x, y, z
The logic behind this decomposition is that each stage tries to ensure that the slower stage prior to it tracks
the desired reference behaviour, unless the reference signal is meant for that stage itself. For instance, the
velocity commands Vc , γc , χc are chosen so that the aircraft flies along the desired path; it is the job of α, β,
µ and the throttle η to ensure that V tracks Vc and so on. The actual control surface deflections are meant
to ensure that the angular rates track the commanded values.
The benefits of using time scale separation is that it allows for a physically-intuitive control design which,
naturally, leads to relatively simpler formulae for designing each component. The drawback of using time
scale-separation in the above form is that it involves the underlying modal structure and associated coupling
between the dynamics. For example, α and q obey a single time scale (short period dynamics), and ditto for
β and r (Dutch roll). This problem can be dealt with by merging the acceleration states and the rotational
dynamics into a single component - you should recongnize this component as the PSS dynamics seen earlier.

5 Nonlinearities
The flight dynamics are inherently nonlinear, and the nonlinearities are quite significant in the way they
influence stability and control properties of the aircraft. A complete discussion is beyond the scope of this
course, but here is a short list of some major effects:
1. The linearization depends strongly on the flight speed and the angle of attack. Therefore, controllers
have to be designed for several equilibrium conditions and pieced together using an interpolation logic.
2. The “basin of attraction” of a stabilized equilibrium is a function of the equilibrium itself. This
necessitates tight bounds on the tolerable disturbances in certain flight regimes, or a richer grid of
control design points in those regimes.
3. Asymmetric aerodynamics at high angle of attack: the aerodynamics of a perfectly symmetric airframe
flying under symmetric conditions become asymmetric at high angles of attack. This point necessitates
a coupled longitudinal-lateral design at moderate and high angles of attack.
4. Departure phenomena: departure phenomena such as deep stall and spin arise due to the existence
of multiple equilibria for a given value of the control input. The only way to avoid departures is to
prevent the aircraft from flying in the regime in which it is susceptible to departure. While departures
can be prevented using active control, this is not the preferred option because the basins of attraction
of individual controllers tend to be quite small in the departure-prone regions.

6 Sensor and Actuator Limitations


Consider the system ẋ = Ax + Bu. When we design a control law for such systems by writing u = Kx, we
implicitly assume that the control actuator can physically produce any signal given to it. This is usually not
true owing to some physical constraints of the actuator:

4
• Time-delay: in transmitting a signal; reaction time of the actuator
• Internal dynamics of an actuator
• Hysteresis: when an actuator is given a cyclic signal, its output is asymmetric and lags the cycle, as
shown in Fig. 2

The internal dynamics of an actuator are usually neither easy to model perfectly nor easy to measure.
Therefore, feedback of actuator states is ruled out.

Figure 2: Hysteresis. Green: expected cyclic path; Red: actual

As an example, consider the first order system

ẋ = −2x + u

If we desire that the system respond faster, we can add u = −kx. The closed-loop eigenvalue, given by
−2 − k, can be moved as far left as desired. Now, suppose that system has an actuator with dynamics
u̇ = −3u + 3uc , and suppose that we use the same feedback law as above. The closed-loop characteristic
system is given by     
ẋ −2 1 x
=
u̇ −3k −3 u

−5 ± 1 − 12k
The closed-loop eigenvalues are , which suggests that the real part cannot be made any more
2
negative that −2.5. In other words, the convergence rate of the closed- loop is directly bounded by the
actuator dynamics. Does this suggest why one should refrain from using the thrust to control anything but
the phugoid and the spiral modes (and that too if no other option is available)?
A similar problem occurs when the aircraft sensors have a finite bandwidth. Consider again the system

ẋ = −2x + u

This time, suppose that the state x is measured by a sensor with its internal dynamics, so that

u = −kxs , ẋs = −3xs + 3x

This creates the same sitation as that of the previous slide; once again, the closed-loop eigenvalues cannot
have a real part any more negative that −2.5.
The discussion in this section leads us to conclude that the actuators and sensors should typically be
faster than the mode that they are trying to control/sense. In particular, the limitations from actuators and
sensors also limit the extent of instability that an airframe can be designed for.

You might also like