You are on page 1of 62

Control of a Reusable Launch Vehicle

JOHAN KNS

Examensarbete
Stockholm, 2011

Abstract
This report examines different control design methods, linear as well as nonlinear,
for a suborbital reusable launch vehicle. An investigation of the natural vehicle
behavior is made, after which a baseline linear controller is constructed to fulfill
certain handling quality criteria. Thereafter the nonlinear cascade control methods
block backstepping and nonlinear dynamic inversion via time scale separation are
examined and used to construct two nonlinear controllers for the vehicle. Optimal
controllers, in terms of three different criteria, are found using the genetic optimization algorithm differential evolution. The optimal controllers are compared, and it
is found that nonlinear dynamic inversion via time scale separation performs better
than block backstepping with respect to the cases investigated. The results suggest control design by global optimization is a viable and promising complement to
classical methods.
Keywords: Flight control systems, nonlinear control, global optimization

Acknowledgments
I am indebted to my supervisors Fredrik Berefelt and John W.C. Robinson at the
Swedish Defence Research Agency, FOI, for their continual guidance and support.
I would also like to thank Professor Xiaoming Hu at the Division of Optimization
and Systems Theory at the Department of Mathematics at KTH Royal Institute of
Technology for his advice and support.

Contents

Aerodynamic Nomenclature

1. Background
8
1.1. Scope and Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2. Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2. Vehicle Model
11
2.1. Actuator Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3. Vehicle Characteristics
3.1. Control Effectors . . . . . .
3.2. Open Loop Behavior . . . .
3.2.1. Pull Up . . . . . . .
3.2.2. Velocity Vector Roll
3.3. Results . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

14
14
15
16
16
16

4. Baseline Controller
21
4.1. Longitudinal Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.2. Lateral Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.3. Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5. Nonlinear Control
5.1. Motivating Idea . . . . . . . . .
5.2. Block Backstepping . . . . . . .
5.2.1. Extension of the Model
5.3. Nonlinear Dynamic Inversion .
5.4. Application to Aircraft . . . . .
5.4.1. BBS . . . . . . . . . . .
5.4.2. NDI-TSS . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

34
34
35
37
39
40
42
43

6. Optimization
44
6.1. Differential Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.2. Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

Contents

6.3. Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
A. Derivations and Proofs
56
A.1. Dynamics of , , and . . . . . . . . . . . . . . . . . . . . . . . . . . 56
A.2. Invertibility of the Virtual Control Gain Matrix
. . . . . . . . . 58

Aerodynamic Nomenclature

a
e
r
J
n

F aero
Mb
u
v

g
l
m
m
n
p
q
r
u
V

angle of attack
sideslip angle
aileron angle
elevator angle
rudder angle
moment of inertia tensor
undamped actuator frequency
heading
actuator time constant
elevation
bank
control surface angles
rotational velocity
aerodynamic force
moment
control surface inputs
velocity
actuator damping coefficient
acceleration due to gravity
rolling moment
airframe mass
pitching moment
yawing moment
roll rate
pitch rate
yaw rate
velocity along body-fixed x-axis
airspeed

Contents
v
w
xe
ye
ze

velocity along body-fixed y-axis


velocity along body-fixed z-axis
x-position in Earth coordinates
y-position in Earth coordinates
z-position in Earth coordinates

CHAPTER

Background

Since the advent of the Space Age, the need for more inexpensive and reliable access to
space has been ever-increasing. The technological and economical barriers involved in
launching vehicles into space have made space access prohibitive to many prospects,
commercial and otherwise. In recent years, significant efforts have been made in
the research and development of fully reusable launch vehicles (RLVs) in order to
provide lower-cost and more reliable access to space (see [4]), not only in scientific
contexts such as transportation to and from the International Space Station, but
also for novel ideas such as space tourism. These initiatives have been undertaken
by several organizations in various parts of the world, including NASA in the United
States and the European Space Agency (ESA) and Virgin Galactic in Europe.

Figure 1.1.: Geometry of the ALPHA.

One project aimed at facilitating space access is FAST20XX (Future High-Altitude


High-Speed Transport 20XX), an ESA research program intended to evaluate and

1.1. Scope and Objectives

lay the groundwork for future high-altitude high-speed transportation. The initiative consists of two sub-projects, one concerning the implementation of the suborbital
aircraft ALPHA (Airplane Launched PHoenix Aircraft) of Fig. 1.1, whose primary
mission is essentially to provide the means for space tourism, and the other with
a hypersonic passenger aircraft, SpaceLiner. Both projects have many technological
obstacles, including exceedingly high temperatures upon atmospheric reentry and excessive accelerations (possibly unsuitable for humans). To tackle these complications,
the projects have been divided into smaller tasks and distributed to 17 specialist companies and institutions in various countries across Europe. This study was conducted
at one of these institutions: the Swedish Defence Research Agency, FOI.

1.1. Scope and Objectives


FOIs task is to design control and stability augmentation systems for the manual
steering system of the ALPHA vehicle. Since the vehicle expends all available fuel
during its climb to the upper atmosphere, no engine power is available during its
descent. At high altitudes, the low air pressure combined with the vehicles lack of
engine power makes the vehicle difficult to control and best suited for steering by
autopilot. The design of the manual steering system is therefore restricted to low
altitudes. Its role as a backup for a potentially malfunctioning autopilot provides not
only the means to control the vehicle during some unforeseeable disturbance or event
with which the autopilot is unable to cope, but also as a psychological reassurance
for the pilot and the passengers.
In order to spare the ALPHA from unnecessarily high surface temperatures upon
its descent through the upper atmosphere, its trajectory has been defined in such a
way as to, among other things, minimize thermal load. At high altitudes where the
aircraft is flying at speeds of several Mach, deviations from this trajectory cannot be
allowed due to these factors. When the aircraft has reached an altitude of ca. 13 km,
it performs a braking maneuver to dispel kinetic energy. After this point, the vehicle
flies more like a regular aircraft and can be treated as such. This study is therefore
restricted to the state trajectory defined between the altitudes of 12.5 km (the altitude
reached immediately after the braking maneuver) and 0 km (see Fig. 1.2). During this
final descent, the nominal Mach number does not exceed 0.5. In order to allow for
maneuvers that possibly cause a positive acceleration while still staying under the
numerically problematic transonic region, the maximum Mach number of interest
that is studied is set to 0.8.
The study begins with an investigation of the airframes behavior (Chapter 3), followed by the design of a baseline linear controller consisting of two decoupled subcontrollers: one for the aircrafts longitudinal modes and one for its lateral modes
(Chapter 4). Thereafter two nonlinear controllers are constructed using cascade control methods (Chapter 5), specifically backstepping and nonlinear dynamic inversion

10

Chapter 1. Background

using time scale separation. The performance of these controllers is investigated using
global optimization tools (Chapter 6).
120
Ascent and Initial Descent
Final Descent

Altitude [km]

100

80
60
40
20

200

400

600
Time [s]

800

1000

1200

Figure 1.2.: The reference trajectory.

1.2. Notation
Vectors and matrices are written in bold font. Vectors are by default column vectors.
The Euclidean norm of a vector v is written as kvk. A variable preceded by a
signifies a deviation from the nominal value. A primed variable signifies a new
variable related to the unprimed variable of the same name. Rotational matrices are
expressed as T ab , where a and b are the source and destination reference frames,
respectively. The derivative of a function f : Rm Rn with respect to a vector
x Rm is denoted by the Jacobian matrix f /x, which at row i and column j
has the derivative fi /xj . The transpose of a matrix A is written AT . The open
interval between a and b is denoted by (a, b) and the closed, left-open, and right-open
by [a, b], (a, b], and [a, b), respectively. The set R+ denotes the strictly positive real
axis. A diagonal square matrix A Rnn with elements a1 , a2 , . . . , an along the
diagonal is written diag(a1 , a2 , . . . , an ). No notational distinction is made between a
function and its Laplace transform; the distinction is evident by the context in which
the function appears.

CHAPTER

Vehicle Model

The model of the airframe forms the foundation of all flight simulations and control
designs made in this study. To construct the model, two right-handed Cartesian
coordinate systems are utilized: the aircraft-fixed and the Earth-fixed coordinate
systems. The aircraft-fixed coordinate system has its origin at the aircraft center of
mass and its x-, y-, and z-axes pointing through the nose, through the starboard
wing, and down, respectively. The Earth-fixed coordinate system has its x-, y-, and
z-axes pointing north, east, and down, respectively. The origin is usually set at the
starting point of a flight, at ground level.
Using these coordinate systems, the following standard six-degree-of-freedom (henceforth 6-DOF) model describing the translation and rotation of the airframe in its
own coordinates is used as a foundation for the model (the derivations of which can
be found in several standard references, e.g. [10, 15]):

0
1
v = F aero + T eb 0 v
(2.1)
m
g
= J 1 (M b J ),

(2.2)


T
where v = u v w is the vehicles velocity, m its mass, F aero R3 the aero
T
dynamic force, g the magnitude of the acceleration due to gravity, = p q r
the angular velocity, J R33 the moment of inertia tensor (here assumed positive

T
definite), and M b = l m n the external moment1 .
1

The variable m can be assumed to refer to the aircraft mass unless explicitly referred to as pitching
moment.

12

Chapter 2. Vehicle Model

To describe the vehicles orientation and position in the Earth-fixed coordinate system, the following six kinematic equations are needed, which can be found in [10]:

1 sin tan cos tan


= 0
cos
sin

0 sin / cos cos / cos


x e
y e = T be v,
ze

(2.3)

(2.4)

where xe , ye , and ze are the positions in the Earth coordinate system. The angles ,
, and are the Euler angles representing heading, elevation, and bank, respectively,
which describe three successive transformations used to rotate the Earth frame to
the aircraft frame, namely rotations about (1) the Earth z-axis, (2) the y-axis of the
rotated coordinate system, and finally (3) the x-axis of the second rotated coordinate
system. The order in which the rotations are performed is important since one triple
of angles does not give a unique orientation for all rotational orders. The order given
above (heading, elevation, bank) is a convention used in the aircraft community.
The force F aero and the moment M b in Eqs. (2.1) and (2.2) are functions of the
aircraft state as well as its position and orientation. In this study, they are approximated using aerodatadata collected from CFD2 calculations and wind tunnel
measurements. These data are parameterized by, among other variables, Mach number. The lowest Mach number measured is 0.2, which is used as a lower Mach number
limit for designs and simulations in this study as well.
Additional variables that are used throughout this text are the airspeed V , angle of
attack , and sideslip angle , which in terms of the defined states can be expressed
as
V = kvk
w
= arctan
u
v
= arcsin .
V

(2.5)
(2.6)
(2.7)

Conversely, the following also holds:

cos cos
v = V sin .
sin cos

(2.8)

The angles of attack and sideslip are shown visually in Figs. 2.1 and 2.2, respectively.
Their domains are set to (2 , 24 ) and (30 , 30 ), respectively, in accordance with
aerodata limits.
2

Computational fluid dynamics.

13

2.1. Actuator Dynamics

x
v

Figure 2.1.: The angle of attack.

Figure 2.2.: The sideslip angle.

2.1. Actuator Dynamics


In order to create a more realistic model of the airframe, the 6-DOF system presented
above is augmented to take the control surface actuator dynamics into account. These
are modeled with first order dynamics as
= u,

(2.9)

where is the actuators time constant and and u the vectors of actuator positions
and control inputs, respectively. The time constant for the ALPHAs actuators is
unknown and is therefore approximated by taking that of the actuators of the similar
NASA X-38 vehicle, which according to [7] have an undamped natural frequency
n = 26 s1 and damping ratio = 0.707, i.e. = 1n 54 ms.

CHAPTER

Vehicle Characteristics

For an appropriate control design to be made, the basic characteristics and behavior
of the aircraft need to be investigated. For example, an airframe may be naturally
stable and thus not require a stability augmentation system, or it may be naturally
unstable and have insufficient control authority. This analysis is done by observing
how the airframe responds to control surface deflections as well as determining the
conditions necessary for the vehicle to be in an equilibrium state.

3.1. Control Effectors


A control system is based on ability to manipulate the airframe state, which is done
using its control effectors. From the start of its descent, the ALPHA has depleted
its fuel and thus has no thrust available. The remaining means for control of the
aircraft are its aerodynamic surfaces. These are given with their abbreviated names
and functions in Table 3.1 and shown with their positions in Fig. 3.1. Positive surface
deflections are defined to follow the industry convention of using the right-hand rule
(see [10, 15]) and are shown in Table 3.2. The limits of the actuators are given in
Table 3.3 and their maximum allowable rate is conservatively set to rmax = 45 s1 .
The inboard as well as outboard surfaces on the wings can be used as both elevators
and ailerons. To avoid potentially unnecessarily complicated control allocation logic,
the inboard elevators EI are chosen for pitch control and the outboard ailerons AO
for roll control, i.e. EO and AI are not engaged throughout this design. This choice
is based on the fact that the outboard ailerons, despite having roughly the same
surface area as their inboard counterparts, are farther from the vehicles main axis

15

3.2. Open Loop Behavior


Control surface

Name

Primary control

Rudder
Body flap
In- and outboard elevators
In- and outboard ailerons
Speed brakes

RU
BF
EI & EO
AI & AO
SB

Yaw
Pitch (primarily for trimming)
Pitch
Roll
Velocity

Table 3.1.: Control surfaces and their functions.

Control surface

Positive deflection

Rudder
Elevators and body flap
Ailerons

Trailing edge left


Trailing edge down
Starboard trailing edge down

Table 3.2.: Positive directions of control surface deflections.

and therefore can exert more moment using the same deflection. Unless otherwise
specified, elevator henceforth refers to EI and aileron to AO.

Speed brake
Rudder

Body flap
Inboard elevator/aileron
Outboard elevator/aileron
Figure 3.1.: The ALPHA and its control surfaces.

Using this knowledge of the control surfaces, one can begin to investigate the aircraft
open loop behavior.

3.2. Open Loop Behavior


The aircrafts open loop behavior is its natural behavior without a control system
engaged. This behavior, or characteristics and limitations, of the airframe is in-

16

Chapter 3. Vehicle Characteristics


Control surface

Lower limit [deg]

Upper limit [deg]

10
0
20
10
15
10
20

10
40
20
10
20
10
20

RU
SB
AO
AI
BF
EO
EI

Table 3.3.: Control surface limits.

vestigated in the context of equilibrium states, i.e. states in which the aircraft is
trimmed (which are also used as linearization points in the linear control design in
Chapter 4). It is first useful to make the observation that the most general steady
state motion (in terms of the body equations) an aircraft can make is helical. This
motion can be decomposed into rotations about the lateral axis (pull ups) and about
the velocity vector (velocity vector rolls).

3.2.1. Pull Up
In flight with constant angle of attack and constant pitching rate the aircraft is said to
be performing a pull up. Contained in this class of maneuvers are not only upward and
downward rotations but also straight and level flight. The states are found by solving

T

T

q(x)

the equation (x)


= 0, where x = v T T xe ye ze is the
aircraft state.

3.2.2. Velocity Vector Roll


A velocity vector roll is achieved when the angles of attack and sideslip as well
as angular velocity are held constant. This is tantamount to a constant rotation
of the airframe about its velocity vector, a movement integral to maneuvers such as


T T = 0,

coordinated turns. The trim states are found by solving (x)

(x)
(x)
where x is as defined above.

3.3. Results
Figures 3.2 and 3.3 show the longitudinal control surface deflections (body flap and
elevator) required for steady state pull ups of different normal accelerations at different Mach numbers. First, the body flap is used to attempt a steady state. If it
saturates, the elevator is engaged. The essential trend is that with higher dynamic

17

3.3. Results

pressure, less deflection is needed. In particular, the body flap is capable of keeping
the vehicle in steady state at high dynamic pressures. The more the height and/or
speed increases, however, the more the elevator needs to be engaged.
Elevator [deg]
3

2.5

2.5

2
nz

nz

Body Flap [deg]


3

1.5

1.5

0.5

0.5

0
0.2

-15

0
0.2

0.4
0.6
Mach

-10

-5

-15

0.4
0.6
Mach

-10

-5

Figure 3.2.: Body flap and elevator deflections needed for a steady state pull up at different
normal acceleration factors nz = qV /g and Mach numbers. The Machnz region that is solid
white in both graphs represents untrimmable states. Altitude: 0 km.

Figures 3.4 and 3.5 show the control surface deflections and angles of attack and
sideslip for different roll rates at 0 km and 12.5 km, respectively. It is apparent in
Fig. 3.5 that the airframe cannot sustain a roll at low dynamic pressures (see the
region between Mach 0.2 and 0.4, for example). The limiting factor appears to be
predominantly the angle of attack boundaries (2 < < 24 ) at lower roll rates
and the elevator at higher roll rates. In other words, there is not enough lift to
keep the airframe in the roll. At high roll rates, one might expect the ailerons to
saturate. However, it is apparent from both figures that the limiting factor is the
rudder limitations.

18

Chapter 3. Vehicle Characteristics

Elevator [deg]
3

2.5

2.5

2
nz

nz

Body Flap [deg]


3

1.5

1.5

0.5

0.5

0
0.2

-15

0
0.2

0.4
0.6
Mach

-10

-5

-10

0.4
0.6
Mach

-5

Figure 3.3.: Body flap and elevator deflections needed for a steady state pull up at different
normal acceleration factors nz = qV /g and Mach numbers. The Machnz region that is solid
white in both graphs represents untrimmable states. Altitude: 12.5 km.

19

3.3. Results

Roll rate [deg/s]

Roll rate [deg/s]

Roll rate [deg/s]

Roll rate [deg/s]

Roll rate [deg/s]

Elevator [deg]
6
4
2
0
-2

100
0
0.2

0.3

0.4

0.5
0.6
Mach
Rudder [deg]

0.7

0.8

8
6
4
2
0

100
0
0.2

0.3

0.4

0.5
0.6
Mach
Aileron [deg]

0.7

0.8
0
-5
-10
-15

100
0
0.2

0.3

0.5
0.6
Mach
Angle of Attack [deg]
0.4

0.7

0.8

12
10
8
6
4
2

100
0
0.2

0.3

0.4

0.5
0.6
Mach
Sideslip Angle [deg]

0.7

0.8

0.5
100

0
-0.5

0
0.2

0.3

0.4

0.5
Mach

0.6

0.7

0.8

Figure 3.4.: Control surface deflections as well as angles of attack and sideslip needed for
velocity vector rolls of varying roll rates and Mach numbers. Altitude: 0 km.

20

Chapter 3. Vehicle Characteristics

Roll rate [deg/s]

Roll rate [deg/s]

Roll rate [deg/s]

Roll rate [deg/s]

Roll rate [deg/s]

Elevator [deg]
15
10
5
0
-5

100
0
0.2

0.3

0.4

0.5
0.6
Mach
Rudder [deg]

0.7

0.8

8
6
4
2
0

100
0
0.2

0.3

0.4

0.5
0.6
Mach
Aileron [deg]

0.7

0.8
0
-5

100

-10
-15

0
0.2

0.3

0.4

0.5
0.6
Mach
Angle of Attack [deg]

0.7

0.8

20
100
0
0.2

10
0.3

0.4

0.5
0.6
Mach
Sideslip Angle [deg]

0.7

0.8

0.5
0
-0.5
-1

100
0
0.2

0.3

0.4

0.5
Mach

0.6

0.7

0.8

Figure 3.5.: Control surface deflections as well as angles of attack and sideslip needed for
velocity vector rolls of varying roll rates and Mach numbers. Altitude: 12.5 km.

CHAPTER

Baseline Controller

In order to have a means for evaluating the performance and response of the nonlinear
controller, a baseline linear controller is designed for the family of general linearized
systems


f (x, u)
f (x, u)
x =
x +
u,
(4.1)
x x=x0
u x=x0
u=u0

u=u0

where x Rn is the aircraft state, u Rp the control input, and f : Rn Rp Rn


its dynamics, of the original nonlinear dynamics
x = f (x, u)

(4.2)


T
uT
at several equilibrium points xT
in the flight envelope. The variables and
0
0
dynamics are defined in more detail below.
The airframes longitudinal and lateral modes are in the linear approximation assumed to be decoupled. The linear controller design can therefore be divided into
two design tasks: pitch channel control (longitudinal) and roll channel control (lateral). Since the vehicle in question is not intended to make any extreme maneuvers,
this approximation is valid for most of the vehicles flight and the controllers relying
on it can be expected to behave well.
The design procedure for each of the control loops of the two channels is as follows:
1. A loop structure is posited with gains left undefined.
2. Desired handling qualities (HQs) are constructed as closed loop pole boundaries
in the complex plane. Because of its ubiquity, a subset of the HQ requirements
outlined by the US Department of Defense in [2] is used (shown in Table 4.1).

22

Chapter 4. Baseline Controller


In terms of the classifications presented therein, the ALPHA can be categorized
as a class II airplane (medium weight, low-to-medium maneuverability), and
the flight phase of interest is consistent with what is classified as category
B (nonterminal Flight Phases that are normally accomplished using gradual
maneuvers and without precision tracking).
3. For several points in the flight envelope (parameterized by altitude and Mach
number):

T
uT
a) The system is linearized around a trim point (see Section 3.2) xT
.
0
0
b) The HQ boundaries are mapped into gain parameter space by Theorem 1.
Here, the polynomial p(s, q) defined in Eq. (4.3) is the denominator of the
relevant transfer function and the parameter vector q is a vector containing
the loop gains. The variable parameterizes the HQ boundaries in the
complex plane. Equation (4.5) is solved for q to find the parameter space
boundaries.
c) The region that fulfills the HQs is located and an arbitrary point inside it
is chosen for the gains.
4. Gain scheduling: the controllers are glued together into a global controller
in the flight envelope by means of interpolation of loop gains using scheduling
variables, here altitude and Mach number.

Theorem 1. Let a : Rm Rn+1 be a function of a parameter vector q Rm and


p : C Rm R a polynomial family parameterized by q as


p(s, q) = 1 s . . . sn a(q).
(4.3)
Then q belongs to the set of roots
{q : p(() + j(), q) = 0, [ , + ]}

(4.4)

if and only if



 
0
d0 () d1 () . . . dn ()

a(q ) =
0
0
d0 () . . . dn1 ()

(4.5)

for some [ , + ], where


d0 () = 1

(4.6)

d1 () = 2()

(4.7)



di () = 2()di1 2 () + 2 () di2 , i = 2, 3, . . . , n.

(4.8)

Proof. See [1].

23

4.1. Longitudinal Control


Flight mode

HQ

Minimum

Maximum

Short period mode

Damping
CAP
Im s 6= 0
Damping
Natural frequency n
n
Im s 6= 0
Time constant
Im s = 0

0.3
0.085

2
3.6

Dutch roll mode

Roll subsidence mode

0.08
0.4
0.1
0

1.4

Table 4.1.: Desired handling qualities.

4.1. Longitudinal Control


The task of this pitch channel control and stability augmentation system is to implement the HQs for the aircrafts longitudinal modes: the short period mode, characterized by a rapid oscillation of the angle attack but with nearly constant airspeed
and altitude, and phugoid mode, a slower pitch, altitude, and airspeed oscillation
without significant change of angle of attack. Phugoid mode, however, is of such a
long period that it can easily be controlled by the pilot. This discussion is therefore
limited to control of the higher frequency short period mode.
The desired HQs for short period mode are summarized in Table 4.1 and consist of
conditions on the damping (that the closed loop poles lie in a certain cone) and on
the control anticipation parameter (CAP). In addition, the mode is required to be
oscillatory since the framework of mapping boundaries from the complex space is
incompatible with damping ratios > 1. The CAP (see [3, 10]) is the ratio of initial
pitch acceleration to steady state normal acceleration. In other words, it is a measure
of how much acceleration to anticipate from an initial one. Mathematically, it can
be written as
n2sp
CAP =
,
(4.9)
nz /
where nsp is the undamped natural frequency of the short period mode and nz the
normal acceleration (i.e. the acceleration in the negative z-direction) in multiples of
g. This equation is solved for nsp for the minimum and maximum CAP, resulting
in an annulus in the complex plane. Together, the HQs form the shape shown in
Fig. 4.1.
Relevant quantities for the short period mode are angle of attack, pitch rate, normal
acceleration, and position of the elevator actuator. It is well-known that pilots are
more sensitive to pitch rate at low speeds and to acceleration at high speeds (see
[18]). The authors of [18] therefore propose a combination of the two quantities.

24

Chapter 4. Baseline Controller

Im

n,min
n,max
min
max

0
Re

Figure 4.1.: Longitudinal HQ boundaries. Inactive boundaries not shown.

This combination, called C , is given by the linear combination nz + Kq. In steady


state, i.e. in the equilibrium state with constant angle of attack, the relation
q=

nz g
V

(4.10)

holds. To find a suitable K, the two components of C are set equal to each other:
nz = Kq = K

nz g
,
VCO

(4.11)

where VCO is the velocity at which the two quantities should have equal influence.
The authors of [18] choose VCO = 122 m s1 , resulting in K = 12.4.
Similar to the longitudinal control loops in [12] and [15], a control loop for C is
constructed, shown in Fig. 4.2. Using the variables mentioned above, the transfer
function from elevator to nz can be expressed as Gnz (s) = Nnz (s)/D(s) and from
elevator to q as Gq (s) = Nq (s)/D(s) using the reduced linearized system
x = Alng x + B lng ue ,

(4.12)

T

T

where x = q nz e , x = q nz e , and the state and control matrices Alng R44 and B lng R4 , respectively, are Jacobian matrices deter
mined by numerical linearization. The transfer function from Ccmd
to C can thus

25

4.2. Lateral Control


be derived from the control loop as
GC (s) =

Ccmd

[Nnz (s) + 12.4Nq (s)]KI


.
sD(s) + [Nnz (s) + 12.4Nq (s)] (KI + KP s)

KI /s

EI

ALPHA
nz

KP
C

(4.13)

q
12.4

Figure 4.2.: Longitudinal control loop.

Theorem 1 is applied to the denominator of the transfer function in Eq. (4.13) to


map the boundaries to parameter space, after which acceptable gains KI and KP can
be chosen.

4.2. Lateral Control


In addition to the longitudinal controller, a (somewhat more advanced) lateral controller must be implemented. The controller has the following objectives:
1. To let the pilot initiate a coordinated turn1 simply by specifying the turn rate
(the rate of change of the heading ). The control loop calculates the bank
angle and control surface deflections.
2. To fulfill the HQ requirements for the lateral modes: Dutch roll mode, a coupled
roll and yaw movement, roll subsidence mode, a mode of pure rolling, and spiral
mode. The spiral mode is usually so slow that it is suppressed unconsciously
by the pilot. However, it is required that its root stay real. For the Dutch roll,
a damping ratio of at least 0.08, a natural frequency of 0.4, and a product of
the two of at least 0.1 are required. The roll subsidence mode needs to be kept
non-oscillatory (i.e. one real eigenvalue) and have a time constant of at most
1.4 s. The Dutch roll mode is kept oscillatory (as the short period mode is in
the longitudinal controller) by setting the maximum damping ratio max = 1,
i.e. it is at most critically damped.
To fulfill these objectives, the control loop in Fig. 4.3 is constructed with an inner
loop similar to the roll-angle hold autopilot presented in [15]. The input is the desired
1

A coordinated turn is, as defined in [15], a turn where there is no aerodynamic force along the
lateral axis. Furthermore, the airspeed and angular velocity need to be constant for the aircraft
to be in steady state.

26

Chapter 4. Baseline Controller

(n )min
n,min
n,min
Real roll root
Roll time constant

Im

1/1.4

Re

Figure 4.3.: Lateral HQ boundaries. Inactive boundaries not shown.

KI /s
cmd Parameter cmd

Lookup

RUcmd

K1

K2

AO

ALPHA
Figure 4.4.: Lateral control loop.

The bank angle and rudder position RU needed for turn coordination
turn rate .
for the specified are computed offline for a range of different turn rates at different
parts of the flight envelope. This computation also includes the necessary values for
the longitudinal states, which are passed to the integrated C controller (not shown
in Fig. 4.3).
Through this loop, the transfer function from cmd to is
G(s) =

K1 N (s)(s + KI )
,
sD(s) sK2 Np (s) + K1 N (s)(s + KI )

(4.14)

where the transfer function from AO to is N (s)/D(s) and from AO to p is


Np (s)/D(s). These transfer functions are derived from the reduced linearized system
for the lateral channel, which for the Dutch roll, roll subsidence, and spiral modes

27

4.3. Results

depends on the sideslip angle , roll rate p, yaw rate r, roll angle , and aileron
actuator position:
x = Alat x + B lat ua ,

(4.15)

T
T


where x = p r a , x = p r a , and the state
and control matrices Alat R55 and B lat R5 , respectively, are Jacobian matrices
determined by numerical linearization.
As in the longitudinal case, Theorem 1 is used to map the boundaries (the shapes
of which are shown in Fig. 4.3) from the complex plane to parameter space to find
satisfactory K1 and K2 . Before applying Theorem 1, however, KI is set to 1/50 (chosen by trial-and-error) to keep the parameter space two-dimensional for simplicity.
Contrary to the longitudinal controller, the lateral controller needs to fulfill the HQs
for multiple modes. Therefore, the boundaries mapped into parameter space need to
be satisfied simultaneously. In other words, the region of intersection of all allowable
regions needs to be found.

4.3. Results
Figure 4.5 shows the mapping of the denominator of the longitudinal controllers
transfer function in Eq. (4.13) to parameter space for three different Mach numbers at
one altitude. Apparently, the two active conditions are maximum natural frequency
and minimum damping, the region S between which is that of acceptable values. This
region shrinks with increasing Mach number as it becomes more difficult to damp
oscillations and lower the natural frequency. For several altitudes and Mach numbers,
a point inside the region is chosen for usage in interpolation in a gain scheduler, i.e. a
function parameterized by the flight envelope position (altitude and Mach number)

T
which gives a vector KP KI as output.
Figures 4.6 and 4.7 show the responses to a unit step in commanded C for two
altitudes at different Mach numbers using the gain scheduled longitudinal C control
loop. The vehicle exhibits a fast and well-damped response to the input under the
regulation of the longitudinal control loop.
Proceeding in a similar manner as in the the longitudinal control loop implementation, the denominator of the lateral controllers transfer function in Eq. (4.14) is
mapped to parameter space for three different Mach numbers at one altitude, shown
in Figure 4.8. As in the longitudinal case, the region S shrinks as the Mach number
increases. The active boundaries in this case, however, are the ones that ensure real
roll subsidence and imaginary Dutch roll roots as well as a satisfactory time constant

T
of the roll subsidence mode. A gain scheduler with the vector of gains K1 K2 as
output is made by choosing a point inside S for several altitudes and Mach numbers
and then interpolating between them.

28

Chapter 4. Baseline Controller

KI

0
-5

-10
-3

-2

-1
KP

-5

KI

KI

-10
-3

-5
-10

-2

-1
KP

-3

-2

-1
KP

Figure 4.5.: Boundaries for the longitudinal parameters mapped by Theorem 1 at an altitude
of 3 km for, from left to right and top to bottom, Mach numbers 0.4, 0.6, and 0.8. The map
for Mach 0.2 is missing since the aircraft is untrimmable with only the body flap at that point.
Origins of boundaries as defined in Fig. 4.1.

Governed by this gain scheduled controller, two coordinated turns with different turn
rates at different parts of the flight envelope are made, the results of which are
shown in Figures 4.9 and 4.10. In both cases the turn rates reach the commanded
rates, and so do the bank angles (with a small steady state error, however). The
overshoots in are quite large, but the damping ratios are within the acceptable
range. With the exception of some excursions in the transient region when the
vehicle is entering the turn, the aerodynamic side force Faero,y = mgny tends to zero,
which, along with the condition of constant angular velocity and constant airspeed,
makes the turn coordinated.

29

4.3. Results

C*

1
Response
Command

0.5
0

10

15

10

15

10

15

Time [s]

C*

1
0.5
0

5
Time [s]

C*

1
0.5
0

5
Time [s]

Figure 4.6.: C unit step responses at an altitude of 3 km at, in order from top to bottom,
Mach 0.4, 0.6, and 0.8.

30

Chapter 4. Baseline Controller

C*

1
Response
Command

0.5
0

10

15

10

15

10

15

Time [s]

C*

1
0.5
0

5
Time [s]

C*

1
0.5
0

5
Time [s]

Figure 4.7.: C unit step responses at an altitude of 9 km at, in order from top to bottom,
Mach 0.6, 0.7, and 0.8.

31

4.3. Results

15

K2

10
5
0
-60

-20

-40

K1

10

10
S

5
0
-60

K2

15

K2

15

-40

-20
K1

5
0

0
-60

-40

-20

K1

Figure 4.8.: Boundaries for the lateral parameters mapped by Theorem 1 at an altitude of
3 km for, from left to right and top to bottom, Mach numbers 0.4, 0.6, and 0.8. The map for
Mach 0.2 is missing since the aircraft is untrimmable with only the body flap at that point.
Origins of boundaries as defined in Fig. 4.3.

32

Chapter 4. Baseline Controller

[deg/s]

20
Response
Command

10
0
0

10
Time [s]

15

20

10
Time [s]

15

20

10
Time [s]

15

20

[deg]

100
50
0

ny

0.5
0
-0.5

Figure 4.9.: Responses to a step in cmd to 4 s1 at an altitude of 3 km at Mach 0.4. The


1
variable ny is the lateral aerodynamic acceleration in terms of g: ny = mg
Faero,y .

33

4.3. Results

[deg/s]

20
Response
Command

10
0
0

10
Time [s]

15

20

10
Time [s]

15

20

10
Time [s]

15

20

[deg]

100
50
0

ny

0.5
0
-0.5

Figure 4.10.: Responses to a step in cmd to 5 s1 at an altitude of 9 km at Mach 0.6.


1
The variable ny is the lateral aerodynamic acceleration in terms of g: ny = mg
Faero,y .

CHAPTER

Nonlinear Control

The advantages of using nonlinear control methods over linear ones are manifold,
in particular the class of controllers of systems in cascade form on which this study
is focused. This class of controls includes, for example, backstepping and nonlinear
dynamic inversion. In contrast to linear control methods, these cascade methods
are not based on linear approximations of the systems dynamics and are thus built
on a more solid foundation. Since the nonlinear system description is valid globally
(in terms of the flight envelope), gain scheduling is not required. Furthermore, the
implementation of these controllers is straightforward.
This is not to say, however, that these controllers are devoid of difficulties. For
example, the dynamics must be in strict-feedback form. Additionally, there is no
clear-cut manner in which to control output variables as in the linear case.
Despite these difficulties, the benefits of this methodology outweigh its drawbacks in
terms of performance and effort required in implementation. This is true particularly
in the case of aircraft, whose motion is essentially governed by a chain of integrators:
commanded control surface deflection rates are integrated into moments, moments
are integrated into rotational velocities, which finally are integrated into orientational
angles. This inherent property makes aircraft well-suited for cascade designs.

5.1. Motivating Idea


Consider the system
xi = fi (x1 , x2 , . . . , xi ) + gi (x1 , x2 , . . . , xi )xi+1 ,

i = 1, 2, . . . , n

(5.1)

35

5.2. Block Backstepping

of order n, with input xn+1 . The idea behind the recursive backstepping and nonlinear dynamic inversion via time scale separation methods is to control xi by using
xi+1 as a virtual control, which is chosen in such a way as to steer xi to the desired
point. The first virtual control x2 is chosen in such a way as to steer x1 to the origin
while at the same time giving it certain desired dynamics f1des (x1 ). Next, x3 is chosen
to steer x2 to its desired value. This process continues recursively for all remaining
states xi , i = 3, 4, . . . , n, and can be done using varying degrees of accuracy, as is
evident below.

5.2. Block Backstepping


Block backstepping (BBS), which can be seen as the most general method in the
class of cascade-based controls since it relies on the fewest assumptions possible, is
a Lyapunov-based1 design technique for deriving control laws for systems in strictfeedback form. An early presentation of the scalar form of the method can be found
in [9], and of the generalized vector form, or block form, applied in a flight setting in
[14]. As in other cascade methods like nonlinear dynamic inversion, the method of
backstepping essentially allows one to supplant the dynamics of systems with desired
ones.
Following the terminology presented in [11, 17], a BBS controller is constructed and
extended with input and output integrators. For simplicity, consider the system
x = f (x) + g(x)y

(5.2)

y = h(x, y) + k(x, y)u,

(5.3)

where x Rm , y Rn , u Rp , the functions f : Rm Rm and h : Rm Rn


Rn are smooth and vanish at the origin, and the functions g : Rm Rmn and
k : Rm Rn Rnp smooth and nonsingular. The objective is to stabilize the origin
of x and prescribe new dynamics f des (x) (the conditions for which are given below)
for the x-subsystem given by Eq. (5.2). This is done by choosing a virtual control
law for the x-subsystem y des (x) that satisfies2
f (x) + g(x)y des (x) = f des (x).

(5.4)

= y y des (x). This is needed since the state y


and introducing an error variable y
can deviate from the desired state y des (x). The objective of assigning new dynamics
, which in turn results in the state y
for x amounts to stabilizing the origin of y

1
2

Lyapunov theory is covered in several standard references, e.g. [8].


Here one could choose to, as in [6], only cancel the unstable dynamics in f to lessen control effector
movement. The trade-off is less freedom when postulating new dynamics. This study therefore
cancels all dynamics regardless of the potential stability of terms therein.

36

Chapter 5. Nonlinear Control

) = h(x, y
+y des (x)) and k(x,
) =
chasing the desired state y des (x). Define h(x,
y
y
des
3
+ y (x)). Equations (5.2) and (5.3) can then be written as
k(x, y
x = f des (x) + g(x)
y

) + k(x,
)u y des (x, y
).
y
= h(x,
y
y

(5.5)
(5.6)

can, in the sense of Lyapunov, be globally asymptotically


The origins of x and y
stabilized by constructing a quadratic control Lyapunov function (clf)
1
1 T
Qy y

) = x T Qx x + y
V (x, y
2
2

(5.7)

with positive definite weighting matrices Qx Rmm and Qy Rnn , and choosing
a control law u such that the clfs time derivative
) = xT Qx x + y
T Qy y
V (x, y



= xT Qx f des (x) + g(x)
y



T Qy h(x,
) + k(x,
)u y des (x, y
)
+y
y
y

(5.8)

is negative definite along the solution trajectory. The control law needs to cancel
potentially indefinite terms in V and is additionally exploited to synthesize new
) for y. One possible control law is
error dynamics hdes
(x, y
y


1
T
des

)1 hdes
(x,
y
)

h(x,
y
)

Q
g(x)
Q
x
+
y
(x,
y
)
,
u = k(x,
y

x
y

(5.9)

where it is assumed that the desired dynamics f des (x) for the x-subsystem and
) for the y
-subsystem are smooth and chosen such that xT Qx f des (x) and
hdes
(x, y
y
T Qy hdes
), respectively, are negative definite (e.g. f des (x) = Ax, where A
y
(x, y
y
nn
R
is negative definite). Combining Eqs. (5.5), (5.6), and (5.9), one arrives at the
system
x = f des (x) + g(x)
y
1
des
) Q g(x)T Qx x,
y
= hy (x, y

(5.10)
(5.11)

= 0, i.e. when y has caught up with


which gives x the desired dynamics when y
des
y (x).

) is the derivative of y des (x) along the solution trajectory, i.e.


The function y des (x, y


y des (x)
y des (x)
y des (x)  des
d  des
y (x) =
x =
(f (x) + g(x)y) =
f (x) + g(x)
y ,
dt
x
x
x
-dependence.
hence the y

5.2. Block Backstepping

37

5.2.1. Extension of the Model


Extension of this framework to include, for example, higher derivative states or integrator states is straightforward and requires only recursive application of the method
described above.
Adding Derivative States
Derivative states can be added to account for inputs that are integrated before entering the model, hence their name input integrators. To augment the model with
an input integrator, one can proceed as follows. Note that the objective is still to
give the x-subsystem new dynamics. Consider, for example, the dynamics
u = f D (x, y, u) + g D (x, y, u)v,

(5.12)

where v Rq , f D : Rm Rn Rp Rp vanishes at the origin, and g D : Rm


= u udes (x, y
) is formed and
Rn Rp Rpq nonsingular. First, the error u
des
des

, u
) = f D (x, y
+ y (x), u
+ u (x, y
)) and g
D (x, y
, u
) =
the functions f D (x, y
+ y des (x), u
+ udes (x, y
)) defined, where udes (x, y
) is the virtual control law
g D (x, y
for the y-subsystem chosen to satisfy Eq. (5.9) in place of u. This error variable
, where Qu Rpp , which is
T Qu u
is then used in a positive definite function 21 u
appended to the clf given in Eq. (5.7) to include the new dynamics:
1 T
, u
) = V (x, y
) + u
.
Qu u
VD (x, y
2

(5.13)

As a result of including the extended dynamics, u cannot be controlled directly.


Therefore, y
is first rewritten to explicitly contain the deviation:

) + k(x,
)u y des (x, y
)
y
= h(x,
y
y
des

) + k(x,
)(u u (x, y
) + udes (x, y
)) y des (x, y
)
= h(x,
y
y

(5.14)

) + k(x,
)
)udes (x, y
) y des (x, y
),
= h(x,
y
y
u + k(x,
y
))4
leading to the clf time derivative (with terms already canceled by udes (x, y
, u
) = xT Qx x + y
T Qy y
T Qu u

V D (x, y
+ u
= xT Qx f des (x)



T Qy hdes

+y
(x,
y
)
+
k(x,
y
)
u

y


, u
) + g
D (x, y
, u
)v u des (x, y
, u
) .
T Qu f D (x, y
+u

(5.15)

, u
) is the derivative of udes (x, y
) along the solution trajectory. The u
The function u des (x, y
dependence appears because of the dependence thereon by y
.

38

Chapter 5. Nonlinear Control

The f des and hdes


terms are, as before, negative definite and need not be canceled.
y
New terms appear not only in u
, but also in y
. To fulfill the requirement of negative
definiteness, the control law is chosen as

D (x, y
, u
)1 f des
, u
) f D (x, y
, u
)
v=g
(x, y
D,u

(5.16)
T
des

Q1
k(x,
y
)
Q
y
+
u
(x,
y
,
u
)
,
y

u
, u
) is the desired error dynamics for u, chosen such that the function
where f des
(x, y
D,u
des
T

u Qu f D,u (x, y , u) is negative definite. This results in the system


x = f des (x) + g(x)
y
des

) + k(x,
)
y
= hy (x, y
y
u Q1 g(x)T Qx x

(5.17)

, u
) Q1
)T Qy y
.
u
= f des
(x, y
D,u
k(x, y
u

(5.19)

(5.18)

Adding Integrator States


Similarly to adding derivative states, the model can be extended to include integrator states, or output integrators. The motivation for including them is not only
to eliminate steady state errors, but also to synthesize (approximate) second order
dynamics. For simplicity, derivative states are excluded from this presentation. As
above, the objective is still to prescribe new dynamics for the x-subsystem. Consider
the system
= f I () + g I ()x,
(5.20)
where Rr , f I : Rr Rr vanishes at the origin, and g I : Rr Rrm nonsingular.
Prescribing new dynamics for this equation would require completely relinquishing
the synthesized dynamics for the x-subsystem. The objective is therefore to merely
stabilize the equation under the assumption that T Q f I () is negative definite for
a positive definite weighting matrix Q . First, the clf is augmented with a negative
definite function of the new state:
1
).
) = T Q + V (x, y
VI (, x, y
2

(5.21)

Then, the time derivative is calculated5 :


) = T Q + xT Qx x + y
T Qy y
V I (, x, y

= T Q (f I () + g I ()x)
+ xT Qx (f des (x) + g(x)
y)



T Qy h(x,
) + k(x,
)u y des (x, y
) .
y
y
+y
5

Note that it is allowed that all lower states also be functions of , e.g. h = h(, x, y).

(5.22)

5.3. Nonlinear Dynamic Inversion

39

This function is made negative definite along the solution trajectory by using the
augmented function
T
f des0 (, x) = f des (x) Q1
x g I () Q

(5.23)

in place of the previously defined f des (x).

5.3. Nonlinear Dynamic Inversion


Nonlinear dynamic inversion via time-scale separation (NDI-TSS) is another technique in the cascade family of controls, which, like BBS, allows one to modify a
systems dynamics. It has been used in settings similar to this study, e.g. in [7]. It
has been shown in [17] that NDI-TSS is very closely related to BBS and that the
two can be regarded as belonging to the same class of control methods. Unlike BBS,
however, the NDI-TSS control method is not formulated using Lyapunov theory, even
though its stability has been proved therewith in [13]. This proof, and the control
method in general, are predicated on the assumption that the subsystems considered
are evolving on different time scales. In other words, fast variables need to be fast
enough to be able to be considered as control inputs to slow variables.
In accordance with the terminology presented above, a system with input and output
integrators is stabilized under the separation of time scales assumption by choosing
) and the control law v as
the virtual control law udes (x, y



) h(x,
)
) = k(x,
)1 hdes
y
(5.24)
udes (x, y
y
(x, y
y


, u
)
, u
) f D (x, y
, u
)1 f des
D (x, y
(5.25)
v=g
(x, y
D,u
in place of the BBS control laws in Eqs. (5.9) and (5.16), respectively. The virtual
control law y des (x) is as defined in Eq. (5.4) and the desired dynamics for the xsubsystem as in Eq. (5.23). With the exception of certain terms being absent, the
control laws look very similar to those of the BBS controller. They are in fact a
simplified case of the BBS control laws. For each subsystem, the time derivative
of the desired signal as well as the inter-subsystem term that guarantees a negative
definite clf are omitted. Comparing the closed loop system
x = f des (x) + g(x)
y
des

) + k(x,
)
)
y
= hy (x, y
y
u y des (x, y
, u
) u des (x, y
, u
)
u
= f des
(x, y
D,u

(5.26)
(5.27)
(5.28)

with that resulting from the BBS method given in Eqs. (5.17) to (5.19), one sees
that the omission of the terms in the control laws leads to the time derivatives of the
virtual controls remaining. These terms are small in comparison to the other terms

40

Chapter 5. Nonlinear Control

when there is time scale separation. As noted in [17], in the limit of complete time
scale separation as well as when Qy = k1 I, Qu = k2 I, k1 , k1 /k2 0, the BBS
and NDI-TSS methods formally converge.
In terms of stability, BBS is clearly superior since it is based on fewer assumptions.
The cost for this guaranteed stability is two extra terms, one of which involves a
potentially complicated differentiation.

5.4. Application to Aircraft


As previously stated, the aircraft model given in Chapter 2 lends itself to cascade
control due to its structure. One way of formulating an aircrafts dynamics so that
it fits the structure required to utilize backstepping and nonlinear dynamic inversion
is to choose the errors between the angles of attack and sideslip, bank angle, rotational velocity vector, the moment, and their respective commanded values as state
variables. The following fully nonlinear controllers6 are specializations of the laws
presented above.
Let

T
=
= cmd

des

(, )

b = M b M des (, ,
),

M
b

(5.29)
(5.30)
(5.31)
(5.32)

the signal needed


where cmd is the vector of commanded angles, M des
b (, , )
des

and (, ) the signal needed to stabilize the origin


to stabilize the origin of ,
The controller functions in such a way as to drive the angles of attack and
of .
sideslip as well as bank angle to their commanded values using the vehicles rotational
velocity, which in turn is driven to the appropriate value by exerting moments using
the control surfaces. This is possible insofar as the actuators and their rates stay
within their bounds.
The orientational dynamics can be expressed as
= f (, ) + g(),

They are also completely decoupled from the linear controller presented in Chapter 4.

(5.33)

41

5.4. Application to Aircraft


where


sec 0 
1
sin
0
cos

0
1
f (, ) =
F tot
cos sin cos sin sin
mV
0
0

cos tan
1
sin tan
sin
0
cos
g() =
1
sin tan cos tan
F tot = F tot (, ).

(5.34)

(5.35)
(5.36)

For derivations of these equations, see Section A.1. The dependence of the force
F tot and the function f on the vector of actuator positions is due to deflections of
the control surfaces not purely exerting moments on the vehicle, but also (relatively
small) forces. This dependence makes a standard BBS design impossible and is in
this study neglected so that F tot = F tot () and f = f ().
= f (
+ cmd )
cmd 7 and g
= g(
+ cmd ). Then, differentiation
()
Let f ()
of Eq. (5.30) yields
+g

= f ()
()

+g
des (, )
+g
.
()
()

= f ()

(5.37)

Using Eq. (2.2), the dynamics for the rotational velocity can be written as
= h() + kM b ,

(5.38)

where h() = J 1 ( J ) and k = J 1 (J is assumed positive definite). Letting


)
and differentiating Eq. (5.31), the error dynamics of the
= h(
+ des (, ))
h(
rotational velocity is
)
)
des (, ,
+ kM b

= h(
des
)
)
)
b.
(, ,
+ kM des (, ,
+ kM
= h(

(5.39)

In order to account for model errors and synthesize approximate second order dynamics, a (virtual) integrator of the orientational error states is included in the controller
as a state R3 :

= K + ,
(5.40)
where K = diag(K, , K, , K, ) has strictly positive diagonal elements. Driving

this state to zero forces to cmd as t . The term K makes the integrator leak and is necessary for the positive definiteness of the clf, or, indirectly,
to ensure stability of the origin of .
7

Here, the formulation is that of a tracking problem. As such, the system is nonautonomous (due
to the time dependence of cmd ). However, the Lyapunov stability arguments presented above,
which are formulated for a setpoint problem, can be shown to be valid in a tracking setting as
well by a generalization of the autonomous case as shown in Theorems 4.8 and 4.9 in [8].

42

Chapter 5. Nonlinear Control

Since the actuators are modeled as first order systems, the moment exerted on the
vehicle cannot be controlled directly. However, by controlling the actuator derivatives, one is essentially controlling the moments derivative. Therefore, the model is
extended with the dynamics8
b = .
M
(5.41)
Combining Eqs. (5.37) and (5.39) to (5.41) and differentiating Eq. (5.32), the complete system can be written as

= K +
+g
des (, )
+g

= f ()
()
()

(5.42)

)
)
+ kM
b
des (, ,
+ kM des

= h(
b (, , )

(5.44)

b =
M

des (, ,
,
b ).
M
M
b

(5.43)

(5.45)

This system is in strict-feedback form and can therefore be controlled using BBS or
,
b
and M
NDI-TSS. To choose a control law, desired dynamics for the states ,
are chosen as the linear functions
des
= K
Q
f
(, )

des ()

= K
h

des
b ) = K M
b,
jM
b (M
Mb

(5.46)
(5.47)
(5.48)

respectively, where K
= diag(Kp , Kq , Kr ), and K M
b =
= diag(K , K , K ), K
diag(Kl , Km , Kn ) have strictly positive diagonal elements and Q , Q
R+ . Then,
is chosen in accordance with Eq. (5.4) as
des (, )



= g()
1 f des

des (, )
(,
)

f
(
)
,
(5.49)

is invertible, or, as can be seen in Section A.2, for all


which is possible when g()

satisfying || 45 , a condition fulfilled for all maneuvers performed in this study.


At this point in the construction of the controller, the BBS and NDI-TSS methods
diverge.

5.4.1. BBS
can be chosen:
From Eq. (5.49), the virtual control M des
b (, , )


Q

des
1 des
des
T

=k
h()
+ (, , )

() ,
M b (, , )
h ()
g
Q
8

(5.50)

Determining the signal to send to the actuators can be decoupled from the design and solved as a
b = M b ( u) for , where M b = M b ()
separate problem, namely solving the equation M

and Eq. (2.9) has been used.

43

5.4. Application to Aircraft


b is stabilized by choosing the control law
where Q R+ . Finally, M
des
b) + M
des (, ,
,
b ) Q kT ,

M
= jM
b (M
b
QM
b

(5.51)

where QM
b R+ .
Combining the system given by Eqs. (5.42) to (5.45) with the desired dynamics given
by Eqs. (5.46) to (5.48) and the control laws given by Eqs. (5.49) to (5.51), the closed
loop system becomes

= K +
Q + g

= K
()

Q T
b
g
() + kM

= K
Q
b Q kT .
b = K M

M
Mb
QM
b

(5.52)
(5.53)
(5.54)
(5.55)

5.4.2. NDI-TSS
The NDI-TSS controller construction is completed in a similar manner, the differ )
and
ence being that the time derivatives of the desired dynamics, des (, ,
des

M b ), as well as extra terms to ensure a negative definite clf time


M b (, , ,
and control input
derivative are omitted so that the virtual control M des
b (, , )
become
 des

()
)
= k1 h
h(

(5.56)
M des

b (, , )
des

b ),
= jM
b (M

(5.57)

respectively, in place of Eqs. (5.50) and (5.51). This results in the closed loop system

= K +
+g

= K
()

(5.58)

)
b
des (, ,
+ kM

= K

(5.60)

b
b = K M
M
Mb

des (, ,
,
b ).
M
M
b

(5.59)

(5.61)

CHAPTER

Optimization

In nonlinear control, analytical tools such as Lyapunov theory are used extensively.
The usage of these tools leads to control laws which guarantee stability, but which
often are conservative. This, in turn, implies potentially excessive control loop gains
and unnecessary control action.
The field of mathematical optimization provides a means of analyzing and removing
the shortcomings of relying solely on analytical methods as well as making their
outcomes more applicable to the real world. Many control problems can be posed as
or augmented with the constrained nonlinear optimization problem
min

J()
(6.1)

subject to h() = 0,
where J : X R is the objective function (for engineering applications generally not
convex), X = { Rn : h() = 0} the feasible set, and h : Rn Rm the m equality
constraints (see [5] for an in-depth discussion of the subject).
In general, one is interested in the global minimum. Since the function J is potentially
non-convex, global optimization methods need to be employed. In practice, these are
much more complex than local optimizers. This, coupled with the fact that designs
in aircraft control often depend on many parameters, implies programs which in
the past would have been essentially unsolvable due to their numerical complexity.
However, with the exponential increase in computing power over the last few years,
solving them is no longer unfeasible, even when they depend on many parameters.
This leads to a new dimension in control design which can be used to solve myriad
problems.

45

6.1. Differential Evolution

6.1. Differential Evolution


In this study, the global optimization method used to solve some of these problems
is differential evolution (DE), a global unconstrained optimization method invented
by Storn and Price (see [16]) in the class of genetic optimization algorithms. Like all
real-world global optimization methods, DE is a heuristic one. As such, it provides no
guarantee of finding the global extremum. However, it has been shown in [16] to be
an effective and competitive algorithm for many classical optimization problems.
The objective is to locate the point of the global extremum of a function J : Rn R
without constraints. Drawing on the principles of biological evolution, the algorithm attempts this by letting a fixed-size population of agents (candidate solutions)
successively evolve into new populations by means of stochastic transformation. Incidentally, having a population with constituents that can evaluate the function independently of each other makes the algorithm scalable and well-suited for parallel
computing. The objective function values given by the agents in a population are
always less than or equal to the objective function values given by their respective
parents. There is currently no standard way of determining whether the objective
function value is good enough. Here, the algorithm is terminated when the number of generations reaches a certain point. Analogous to the biological function of
genetic recombination, DE also incorporates a process known as crossover to increase
population diversity. Algorithm 1 shows the steps of the method.
Algorithm 1. Unconstrained Differential Evolution
0. Create the initial generation with a population consisting of N 4 agents
x1 , x2 , . . . , xN chosen randomly1 from Rn . Let I := {1, 2, . . . , N } be the set of
indices. Define the termination criterion by defining the maximum number of
generations kmax N.
For each generation k = 1, 2, 3, . . .:
1. If k = kmax , terminate.

T
2. For each xi = xi,1 xi,2 . . . xi,n , i I:
a) Choose three mutually distinct indices u, v, w randomly from I \ {i}.
b) Mutation: generate a vector
x0 = xu + F (xv xw ),

(6.2)

where F (0, 2] is an amplification factor.

The agents in the population are not required to be unique. Therefore the population is not a set.

46

Chapter 6. Optimization
T

with
c) Crossover: generate a recombinant vector y = y1 y2 . . . yn
elements
(
x0j
for j {l {1, 2, . . . , n} : rl kCO } {r}
yj =
(6.3)
xi,j otherwise,
j {1, 2, . . . , n}, where r1 , r2 , . . . , rn are chosen randomly from [0, 1], kCO
is the crossover constant chosen from [0, 1] to facilitate diversity, and
r is an integer chosen randomly from {1, 2, . . . , n} to ensure that not all
elements are taken from xi .
d) If J(y) < J(x), then set z i := y. Otherwise set z i := xi .
3. Set xi := z i

i I.

6.2. Application
Other studies, e.g. [14], have shown the efficacy of global optimization methods in
shaping time domain responses. Here, the feasibility of the methodology presented
above is illustrated with three design objectives concerned with inherent system properties, formulated as separate minimizations of (a) the systems L2 gain, (b) actuator
flutter, and (c) the maximum actuator rate. Each minimization is done for both
BBS and NDI-TSS at several points i in the flight envelope using the control laws
given in Section 5.4. The properties are optimized using the DE algorithm with
crossover constant kCO = 0.8, amplification factor F = 0.5, and population size
N = 10. For the remainder of this section, a dot in a subscript implies the expression
is valid for both the BBS and NDI-TSS cases, e.g. f = g implies fBBS = gBBS and
fNDITSS = gNDITSS (but not fBBS = gNDITSS ).
In order for the simulations to be realistic, constraints need to be added. The normalized time domain response of a cmd = 2 pull up (i.e. a pitching down motion
of 2 ) at point i in the flight envelope
yi, (t) =

i, (t) 0
,
cmd

(6.4)

where 0 is the initial (trimmed) value of the angle of attack, is required to lie within
the bounds presented in Fig. 6.1. Furthermore, constraints are added to keep the
elevator from saturating (see Table 3.3 for its limits) and its rate below rmax (see
Section 3.1). In order for the time derivative of the commanded value (which is used
by the controllers) to be finite, the step is replaced with a ramp of duration 0.5 s, so
that
(
t
cmd 0.5
for 0 t < 0.5
cmd (t) = 0 +
(6.5)
cmd
for 0.5 t,

47

6.2. Application
which, normalized, is
u(t) =

cmd (t) 0
.
cmd

(6.6)

The remaining commanded values cmd and cmd in the vector cmd are set to their
trimmed values. The constraints are formulated as equality constraints and collected
in the non-negative function

e,i, (t) e,max

(t)

e,min
e,i,
XZ T

max 0, e,i, (t) rmax dt,


h ( ) =
(6.7)

yi, (t) f+ (t)


i

f (t) yi, (t)


which is zero on the feasible set and non-zero on the complement of the feasible set.
Here is the vector of parameters (defined below for BBS and NDI-TSS), T the
simulation time, e,i, (t) the elevator deflection starting at point i in the flight envelope, e,max and e,min the maximum and minimum elevator deflections, respectively,
f+ (t) and f (t) the upper and lower boundaries shown in Fig. 6.1, respectively, and
the maximum is performed componentwise. The functions e,i, (t) and yi, (t) have
an implicit dependence. Since the DE algorithm does not provide a built-in way
of minimizing a function subject to constraints, one can encourage it to keep the
search within the feasible set X by penalizing the objective function when it leaves
X by adding the penalty function
1
h ( )T h ( )
2

(6.8)

to the objective function, where R55 is a positive definite diagonal matrix of


penalty parameters chosen in such a way that the penalty function dominates over
the objective function on Rn \ X. Note that the domain of the objective function
now needs to encompass all of Rn .
The vector of parameters (all defined in Section 5.4) that define the search space is
given by

T
BBS = Q
(6.9)
Q
QM
b K Kq Km
for BBS, i.e. the weights of the blocks in the clf and the magnitudes of the poles of
the desired dynamics. NDI-TSS does not require the weights for the nor the M b
blocks, so the parameter vector becomes

T
NDITSS = Q
.
K Kq Km

(6.10)

Both the BBS and NDI-TSS optimizations have Q = 1.


Analogous to the H norm for linear systems (see [19]), the L2 gain is a property
worthy of consideration when choosing the parameters of a system since it gives

48

Chapter 6. Optimization
1.4

Normalized step response

1.2
1
0.8
0.6
0.4
0.2
0

3
Time [s]

Figure 6.1.: Time domain response envelope as used in [7].

a measure of the sensitivity of an output to an input. A system with input u


2
n
Lm
2 ({0} R+ ) resulting in an output y L2 ({0} R+ ) is said to have an L2 gain
of at most R+ when, for all u and all finite T > 0,
Z T
Z T
2
2
ky(t)k dt
ku(t)k2 dt.
(6.11)
0

Here, the output y(t) is replaced with the pitching moment m(t) caused by the input
u(t), which in turn is replaced with the normalized commanded angle of attack u(t),
so that the gain gives a measure of how violently the controller performs maneuvers.
Two simplifications are made for a rough numerical approximation of the gain:
1. Only a single input is considered.
2. T is set to 6 s, the maximum time for which the time domain response envelope
is defined (see Fig. 6.1). This limit is sufficient to capture the essential behavior
of the response.
The objective is to find the maximum of the gains with respect to the examined
points i of the flight envelope, i.e.
!1/2
RT
2
0 mi, (t) dt
f, ( ) = max
(6.12)
RT
2
i
0 u(t) dt
2

n
For X R, Ln
2 (X) = {f : X R :

R
X

kf (t)k2 dt < }.

49

6.2. Application

is used as an unconstrained objective function, where mi, (t) is the pitching moment
(implicitly dependent on ) starting at point i. Essentially, one is minimizing the
pitching moment the controller requires for the vehicle to reach the commanded angle
of attack.
Another useful criterion to consider when choosing control loop gains is the movement
of the actuators. For energy saving purposes as well as to minimize wear and tear
on the actuators, one would like to minimize their movement. Here, the sum of the
elevator actuator time derivatives L2 norms
fL2 , ( ) =

X Z
i

1/2
2

|e,i, (t)| dt

(6.13)

is used as an unconstrained objective function, where e,i, (t), implicitly dependent


on , is the elevator deflection starting at point i.
Finally, the maximum of the actuator ratesakin to their maximum poweris minimized. This is done using the maximum of the uniform norm of the actuator rates
with respect to the examined points i of the flight envelope
f, ( ) = max |e,i, (t)|
i,t

(6.14)

as the unconstrained objective function.


In conclusion, optimizations are performed for BBS and NDI-TSS for the following
three parameters:
1. L2 gain from commanded angle of attack to pitching moment for comparing
moment exertion on the airfame using the objective function
1
J, ( ) = f, ( ) + h ( )T h ( ).
2

(6.15)

2. Actuator flutter for comparing energy usage of the actuators using the objective
function
1
JL2 , ( ) = fL2 , ( ) + h ( )T h ( ).
2

(6.16)

3. Actuator rates for comparing maximum power of the actuators using the objective function
1
J, ( ) = f, ( ) + h ( )T h ( ).
2

(6.17)

50

Chapter 6. Optimization

6.3. Results
The results of the three optimizations described above are shown in Tables 6.1 to 6.3.
For each controller, several generations are listed with the function values given by the
optimal agents in each respective population, as well as the dispersion of the agents
normalized by the bounds of the parameters. The dispersion is a useful measure to
consider when determining when to terminate the optimization. The less dispersed
the agents become, the more certain one can be about having found a local minimum
(that may also is a global one).
All of the BBS optimizations are done with completely random initial populations (as
is evident by thein comparison to the final valuesvery large function values in the
early generations, caused by penalty terms), and the optimizer manages to find the
feasible set in each minimization. In the case of the NDI-TSS optimizations, however,
the optimizer has difficulty locating the feasible set. Therefore, the optimizations are
restarted with one agent in the initial population in the feasible set. The remaining
agents are, as in the case of the BBS optimizations, chosen randomly.
Generation

Best function value

Dispersion

BBS

1
50
100
500
1000
5000

5.422 1010
3.386 102
3.358 102
3.349 102
3.349 102
3.349 102

6 101
1 101
3 102
2 103
8 104
9 1017

NDI-TSS

1
50
100
500
1000
5000

4.799 102
7.537 101
7.462 101
7.459 101
7.459 101
7.459 101

6 101
2 101
3 102
2 106
5 107
6 1017

Table 6.1.: L2 gain minimization.

The results from the L2 gain optimizations are summarized in Table 6.1. The NDITSS controller reaches a smaller value than the BBS controller, suggesting it in
general uses less pitching moment to reach the values it is given as commands. The
differences in behavior of the two optimized controllers can be seen visually in Fig. 6.2.
The optimal BBS controller is much faster than the optimal sluggish but economical
NDI-TSS controller. The trade-off for this speed is more elevator deflection, hence
more pitching moment exertion.
As in the L2 gain optimizations, the NDI-TSS controller is capable of a smaller L2
norm, in this case measuring the actuator energy usage, than the BBS controller. The

51

6.4. Conclusion
Generation

Best function value

Dispersion

BBS

1
50
100
500
1000
5000

1.238 107
2.804 101
2.674 101
2.533 101
2.513 101
2.512 101

7 101
3 101
3 101
3 104
1 106
6 109

NDI-TSS

1
50
100
500
1000
5000

3.586 101
5.013
4.848
4.820
4.820
4.820

6 101
2 101
7 102
1 102
1 103
2 1017

Table 6.2.: L2 norm minimization.

results of the optimizations are shown in Table 6.2. The superiority of the optimal
NDI-TSS controller is visually apparent in Fig. 6.3 by the movement of the elevator.
It keeps the elevator fairly constant throughout the maneuver, whereas the optimal
BBS controller outputs some relatively large deflections in the transient region. Just
like the controllers optimal in L2 gain, the BBS controller is fast and the NDI-TSS
controller sluggish.
Finally, the uniform norm optimizations, i.e. power minimizations, are summarized
in Table 6.3. As in the previous two cases, NDI-TSS is more conservative. Unlike
the other two optimizations, one can see from the response in Fig. 6.4 that the
BBS controller is slower and has a larger overshoot. Also, the elevator movement is
smoother, which is what one would expect when minimizing elevator power.
In all three cases, the respective optimal NDI-TSS controllers are better in the
sense that the their optimizations reach lower objective function values than their
optimal BBS counterparts.

6.4. Conclusion
As can be seen in the responses of the optimal controllers, the NDI-TSS controllers
are very sluggish and seem to not even reach steady state. A fairer comparison
between BBS and NDI-TSS could be made if the time domain boundaries were made
stricter and if the objective function also were a function of the deviation of the
response from a desired one, such as a typical second order response. Furthermore,
more extreme maneuvers could be investigated. These investigations may reveal
BBS to be the superior control method of the two. In addition to these changes, the

52

Chapter 6. Optimization
Generation

Best function value

Dispersion

BBS

1
50
100
500
1000
5000

5.423 107
3.509 101
3.355 101
1.160 101
1.157 101
1.156 101

6 101
1 101
2 101
2 105
1 106
1 1016

NDI-TSS

1
50
100
500
1000
5000

4.217 101
4.330
4.305
4.290
4.290
4.290

6 101
1 101
2 101
7 104
8 1010
8 1010

Table 6.3.: Uniform norm minimization.

objective functions could be made more accurate. For example, the L2 gain can be
better approximated by testing more inputs, and the norms examined by testing more
initial conditions, i.e. performing maneuvers in more points of the flight envelope.
Global optimization using genetic algorithms proved a useful tool not only in choosing
parameters for BBS and NDI-TSS controllers, but also in comparing them. This
method of design allows for the efficient choice of system parameters for a host of
objectives, e.g. minimization of energy loss and actuator movement. These results
suggest that global optimization in nonlinear control is a viable and promising design
methodology.

replacemen
6.4. Conclusion

53

Normalized step response

1.2
1
0.8
0.6
BBS
NDI
Command
Boundaries

0.4
0.2
0

3
Time [s]

3
Time [s]

EI

0
-5
-10

Figure 6.2.: Time domain responses and elevator deflections using BBS and NDI-TSS controllers with parameters optimized for minimum moment exertion.

replacemen
54

Chapter 6. Optimization

Normalized step response

1.2
1
0.8
0.6
BBS
NDI
Command
Boundaries

0.4
0.2
0

3
Time [s]

3
Time [s]

EI

0
-5
-10

Figure 6.3.: Time domain responses and elevator deflections using BBS and NDI-TSS controllers with parameters optimized for actuator energy conservation.

replacemen
6.4. Conclusion

55

Normalized step response

1.2
1
0.8
0.6
BBS
NDI
Command
Boundaries

0.4
0.2
0

3
Time [s]

3
Time [s]

EI

0
-5
-10

Figure 6.4.: Time domain responses and elevator deflections using BBS and NDI-TSS controllers with parameters optimized for actuator power conservation.

APPENDIX

Derivations and Proofs

A.1. Dynamics of , , and


Using Eqs. (2.1), (2.6), and (2.8) and

F tot

0
= F aero + T eb 0 ,
mg

(A.1)

the dynamics for the angle of attack is derived as follows:


w
d
arctan
dt
u
1
uw wu
=
1 + ( wu )2
u2
1
(wV
cos cos uV
sin cos )
= 2
V v2
1
= 2
(wV
cos cos uV
sin cos )
V (1 sin2 )
sec
=
(w cos u sin )
V


 F tot
sec 
sin 0 cos
=
v
V
m

q sin cos r sin



 F tot
= sec sin 0 cos
r cos cos p sin cos
mV
p sin q cos cos




sec
sin 0 cos F tot + cos tan 1 sin tan .
=
mV

(A.2)

A.1. Dynamics of , , and

57

Similarly, using Eqs. (2.1), (2.7), and (2.8) and the above definition of F tot , the
sideslip dynamics is:
v
d
arcsin
=
dt
V
1
V v v V
=q
2
V2
1 v
V2


1
v 
v V
V
V 2 v2

1
v T 
=p
v

v v
V2
V 2 (1 sin2 )


sec 
v
0 1 0 2 v T v
=
V
V


T 
sec 
0 1 0 sin cos cos sin sin cos
=
v
V

1 
cos sin sec tan sin sin sin v
=
V


 F tot
1 
cos sin cos sin sin
=
v
V
m

1 
cos sin cos sin sin F tot
=
mV

q sin cos r sin




+ cos sin cos sin sin r cos cos p sin cos
p sin q cos cos



1 
cos sin cos sin sin F tot + sin 0 cos .
=
mV
=

(A.3)

The combination of Eqs. (A.2) and (A.3) along with Eq. (2.3) leads to the relation

= f (, , , ) + g(, , ),
(A.4)

where


sec 0 
1
sin
0
cos
0
1
f (, , , ) =
F tot
cos sin cos sin sin
mV
0
0

cos tan
1
sin tan
sin
0
cos .
g(, , ) =
1
sin tan cos tan

(A.5)

(A.6)

58

Appendix A. Derivations and Proofs

A.2. Invertibility of the Virtual Control Gain Matrix

Proposition 1. The virtual control gain matrix g() given in Eq. (5.35) is invertible
for all satisfying || 45 and all , [30 , 30 ].
Proof. Assume g() is singular for some satisfying || 45 . Then its determinant
det g() = cos tan sin tan cos
sin (cos tan + sin tan sin tan ) cos
= cos2 tan sin tan sin2 sin tan tan

(A.7)

sin cos tan cos


= ( tan sin sin cos ) tan cos
is zero, so that
cot =

tan sin + sin cos


.
cos

However, the domain of and is [30 , 30 ], so that






 sin


| cot | = tan sec tan
cos


tan sec




tan
q
= (tan sec )2 + tan2
q
(tan 30 sec 30 )2 + tan2 30
v
!2
u
!2
u
3 2
3
t

=
+
3 3
3

= 7/3,
i.e.

(A.8)

(A.9)

3 7
|| arctan
> 45 .
(A.10)
7
This is an obvious contradiction of the initial assumption and thus completes the
proof.

Bibliography

[1] J
urgen Ackermann. Robust Control. Springer Verlag, 1997.
[2] Anonymous. Military specification flying qualities of piloted airplanes. Specification, Department of Defense, United State of America, November 1980.
MIL-F-8785C.
[3] Anonymous. Department of defense handbook flying qualities of piloted aircraft. Handbook, Department of Defense, United States of America, December
1997. MIL-HDBK-1797.
[4] Anonymous. The space launch initiative: Technology to pioneer the space frontier. Technical Report FS-2002-04-87-MSFC, NASA Marshall Space Flight Center, April 2002.
[5] Dimitri P. Bertsekas. Nonlinear Programming. Athena Scientific, New Hampshire, second edition, 1999.
[6] Ola H
arkeg
ard. Flight Control Design Using Backstepping. Licentiate thesis
no. 875, Department of Electrical Engineering, Linkoping University, SE-581 83
Linkoping, Sweden, March 2001.
[7] Daigoro Ito, Jennifer Georgie, John Valasek, and Donald T. Ward. Reentry
vehicle flight controls design guidelines: Dynamic inversion. Technical report,
NASA, March 2002.
[8] Hassan K. Khalil. Nonlinear Systems. Prentice Hall, third edition, 2002.
[9] Petar V. Kokotovic. The joy of feedback: Nonlinear and adaptive. Control
Systems, IEEE, 12(3):717, June 1992.
[10] Warren F. Phillips. Mechanics of Flight. John Wiley & Sons, Inc., second edition,
2010.
[11] John W.C. Robinson. Block backstepping for nonlinear flight control law design.
In D.G. Bates and M. Hagstr
om, editors, Nonlinear Analysis and Synthesis
Techniques for Aircraft Control, pages 231257. Springer, Berlin, 2007. Lecture
Notes in Control and Information Sciences Series, Vol. 365.

60

Bibliography

[12] D. Saussie, L. Saydy, and O. Akhrif. Longitudinal flight control design with handling quality requirements. Aeronautical Journal, 110(1111):627637, September
2006. Paper No. 2989.
[13] Corey J. Schumacher, Pramod P. Khargonekar, and N. Harris McClamroch.
Stability analysis of dynamic inversion controllers using time-scale separation.
In Proceedings of the AIAA Guidance, Navigation, and Control Conference and
Exhibit, Boston, Massachusetts, August 1998. AIAA paper 1998-4322.
[14] Marc L. Steinberg and Anthony B. Page. Nonlinear adaptive flight control
with genetic algorithm design optimization. International Journal of Robust
and Nonlinear Control, 9:10971115, December 1999.
[15] Brian L. Stevens and Frank L. Lewis. Aircraft Control and Simulation. John
Wiley & Sons, Inc., second edition, 2003.
[16] Rainer Storn and Kenneth Price. Differential evolution a simple and efficient
heuristic for global optimization over continuous spaces. Journal of Global Optimization, 11:341359, December 1997.
[17] Johan Thunberg and John W.C. Robinson. Block backstepping, NDI and related
cascade designs for efficient development of nonlinear flight control laws. In
Proceedings of the AIAA Guidance, Navigation and Control Conference and
Exhibit, Honolulu, Hawaii, August 2008. AIAA paper AIAA 2008-6960.
[18] Harold N. Tobie, Elden M. Elliott, and Lawrence G. Malcom. A new longitudinal
handling qualities criterion. In Proceedings of the National Aerospace Electronics
Conference, pages 9399, Dayton, Ohio, May 1966.
[19] A.J. van der Schaft. L2 -gain analysis of nonlinear systems and nonlinear state
feedback H control. IEEE Transactions on Automatic Control, 37(6):770784,
1992.

Examensarbete E357 i
Optimeringslra och systemteori
September 2011

www.math.kth.se/optsyst/

You might also like