Vous êtes sur la page 1sur 133

Integrated Modeling and Robust

Control for Full-Envelope Flight of


Robotic Helicopters

Marco La Civita

Department of Mechanical Engineering


Carnegie Mellon University
Pittsburgh, Pennsylvania

Dissertation submitted in partial fulfillment of the requirements


for the degree of Doctor of Philosophy
December, 2002
i

c
2002 by Marco La Civita. All rights reserved.
ii

Abstract

Robotic helicopters have attracted a great deal of interest from the university, the in-
dustry, and the military world. They are versatile machines and there is a large number
of important missions that they could accomplish. Nonetheless, there are only a hand-
ful of documented examples of robotic-helicopter applications in real-world scenarios.
This situation is mainly due to the poor flight performance that can be achieved and
— more important — guaranteed under automatic control. Given the maturity of con-
trol theory, and given the large body of knowledge in helicopter dynamics, it seems
that the lack of success in flying high-performance controllers for robotic helicopters,
especially by academic groups and by small industries, has nothing to do with heli-
copters or control theory as such. The problem lies instead in the large amount of time
and resources needed to synthesize, test, and implement new control systems with the
approach normally followed in the aeronautical industry.
This thesis attempts to provide a solution by presenting a modeling and control
framework that minimizes the time, cost, and both human and physical resources nec-
essary to design high-performance flight controllers. The work is divided in two main
parts. The first consists of the development of a modeling technique that allows the
designer to obtain a high-fidelity model adequate for both real-time simulation and
controller design, with few flight, ground, and wind-tunnel tests and a modest level
of complexity in the dynamic equations. The second consists of the exploitation of
the predictive capabilities of the model and of the robust stability and performance
guarantees of the H∞ loop-shaping control theory to reduce the number of iterations of
the design/simulated-evaluation/flight-test–evaluation procedure. The effectiveness
of this strategy is demonstrated by designing and flight testing a wide-envelope high-
performance controller for the Carnegie Mellon University robotic helicopter.
iii

Acknowledgments

I would like to thank my advisors Bill Messner and Takeo Kanade for their constant
support and advice during my research years at Carnegie Mellon University.
I am equally in debt with several other people besides my advisors.
Mark Tishler at NASA Ames NASA/Army Rotorcraft Division for sharing with me
some of his legendary knowledge of helicopter dynamics and control, and for funding
part of my research (NASA Grant NAG2-1441).
George Papageorgiou for his invaluable technical advice and guidance on robust
control, and for being a member of my Ph.D. committee.
Adnan Akay for making sure that my stay at the Department of Mechanical En-
gineering was the most comfortable possible, and for being a member of my Ph.D.
committee.
Omead Amidi for having started the Autonomous Helicopter Project and for lead-
ing it in the best way possible, for his availability, and for his generosity.
Bernard Mettler for bringing me into this fascinating research area not only men-
tally, but also physically, literally driving me with his car across the United States from
Pittsburgh to NASA Ames.
Ryan Miller and Todd Dudek for their assistance and for working extra hours to
allow me to do the flight tests presented in this thesis.
Noah Falk for designing the graphical interface to my flight simulator.
Mark Bedillion for proofreading this thesis.
I would also like to acknowledge some of the many people which have made en-
joyable my life in Pittsburgh: my Italian friend Andrea Gambotto, the Pittsburgh De-
partment of Leisure, Fun, and other Pachangas, and my family — in Italy and Spain —
that I always felt unconditionally present and supportive no matter the distance that
physically separates us.
Last, but not least, Rocı́o: I would not have accomplished any of this without her
presence. Gracias, compañera.
Contents

1 Introduction 1
1.1 Motivations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Modeling background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Control background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Outline of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2 Integrated modeling 11
2.1 First-principles non-linear model . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.1 Fuselage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.1.2 Gravity forces and Euler angles . . . . . . . . . . . . . . . . . . . . 14
2.1.3 Main rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.4 Actuators and stabilizer bar . . . . . . . . . . . . . . . . . . . . . . 21
2.1.5 Tail rotor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.1.6 Aerodynamic forces . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.1.7 System of equations . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2 Linear models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.1 Trimming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.2.2 Linearization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

iv
Contents v

2.3 Model and flight-test frequency responses . . . . . . . . . . . . . . . . . . 28


2.4 Identification of physical parameters . . . . . . . . . . . . . . . . . . . . . 29
2.5 Reduced-order linear models . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.6 MOSCA-model evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.6.1 Identified physical parameters . . . . . . . . . . . . . . . . . . . . 33
2.6.2 Flight-test comparisons . . . . . . . . . . . . . . . . . . . . . . . . 34
2.7 Conclusions on integrated modeling . . . . . . . . . . . . . . . . . . . . . 35

3 Robust hover control 40


3.1 H∞ loop shaping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.1 Classical loop shaping . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.1.2 Singular values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.1.3 Coprime factor uncertainty . . . . . . . . . . . . . . . . . . . . . . 43
3.1.4 Design procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2 Inner-loop design and analysis . . . . . . . . . . . . . . . . . . . . . . . . 47
3.3 Outer-loop design and analysis . . . . . . . . . . . . . . . . . . . . . . . . 58
3.4 Controller implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
3.5 Flight tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.6 Loading analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.7 Conclusions on robust hover control . . . . . . . . . . . . . . . . . . . . . 76

4 Robust full-envelope flight control 77


4.1 Observer form for scheduling . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.2 Controller structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.3 Inner-loop design for fixed operating points . . . . . . . . . . . . . . . . . 83
4.4 Scheduling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.5 Scheduled inner-loop evaluation . . . . . . . . . . . . . . . . . . . . . . . 93
4.6 Outer loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
4.7 Controller implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Contents vi

4.8 Flight tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

5 Conclusions 116
5.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.2 Suggestions for future work . . . . . . . . . . . . . . . . . . . . . . . . . . 117

Bibliography 118
Chapter 1
Introduction

1.1 Motivations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Modeling background . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.3 Control background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.4 Outline of contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

1.1 Motivations

Robotic helicopters have attracted a great deal of interest from the university, the indus-
try, and the military world. The number of situations in which they could be deployed
and used successfully is limited only by men’s imagination. The following list cites a
few of the possible scenarios.

Search and rescue: “More than 3 million people owe their lives to the special charac-
teristics of these aircrafts.”1 Robotic helicopters have the possibility to increase
dramatically the number of situations where people’s lives are currently saved
with the use of manned full-scale helicopters. These platforms will be able to
1 M.E. Rhett Flater, A Letter to the President George W. Bush, Washington Post, 27 March 2001.

1
1.1. Motivations 2

search quickly and systematically a very large area and could be more readily de-
ployed in weather conditions that would normally prevent human piloted search
and rescue and could be sacrificed in very dangerous conditions (fire, radioactive,
and chemical accident areas) to save human lives.

Surveillance: Robotic helicopters could perform a variety of around-the-clock surveil-


lance operations and report interesting or unusual activity.

Law Enforcement: Robotic helicopters could fly overhead to aid police forces in dan-
gerous high-speed chases or criminal search operations.

Inspection: Robotic helicopters could inspect high voltage electrical lines, bridges, and
dams in remote locations and monitor traffic cost-effectively.

Aerial Mapping: Robotic helicopters could build more accurate topological maps than
conventional aircraft with substantial cost savings by flying in smaller and more
constrained areas.

Wildlife Observation: “In just two acres of Amazonian rain-forest there are about four
times more species of trees than can be found on the whole of the North Amer-
ican continent.”2 Robotic helicopters could be a breakthrough in the human-
based identification and mapping of tropical trees species and the monitoring of
wildlife.

Cinematography: Robotic helicopters could automatically track subjects with on-board


vision-based object trackers, and could fly a prescribed path with high accuracy
to help in planning shots and producing special effects.

In spite of the importance of these applications, there are only a handful of doc-
umented examples of robotic-helicopter applications. This situation is mainly due to
2 Miroslav Honzák, http://www.passporttoknowledge.com/rainforest/Interact/Remote/Remote.html
1.1. Motivations 3

the poor flight performance that can be achieved and — more important — guaran-
teed under automatic control. Autonomous flight is typically implemented through
Guidance, Navigation, and Control (GNC) systems. The success of guidance and nav-
igation components relies strongly upon what level of performance is deliverable by
the control component. This is because the open-loop dynamics of helicopters are un-
stable, multivariable, highly coupled, non-minimum phase, and vary widely across the
flight envelope. Thus, the limits in the flight-control component of GNC systems jeop-
ardize the overall performance, and have prevented robotic-helicopter applications in
real-world missions.
This situation is quite surprising given the maturity reached by control theory and
given the large body of knowledge and experience in helicopter dynamics. It seems
that the lack of success in effective flight of robotic helicopters, especially by academic
groups and by small industries, has nothing to do with helicopters or control theory
as such, but it is mainly due to the limited resources in terms of costs and man hours
needed to design, test, and implement new control systems. In the aeronautical in-
dustry (assuming that an instrumented vehicle is available), the flight-controller design
process for an aerial vehicle is composed of the following activities:

• The development of an accurate mathematical model for the vehicle. This activity
may require (for the actual model generation and/or its validation) the combi-
nation of extensive knowledge of flight dynamics, numerous piloted flight tests,
wind tunnel tests, and static and dynamic ground tests.

• The use of the model for controller synthesis with model-based controller design
techniques.

• The use of the model in simulated flight tests for performance evaluation of the
synthesized model- or non-model–based controller.

• The flight testing of the control laws on the real vehicle and the performance eval-
uation.
1.2. Modeling background 4

• The iteration of controller design, simulated and real flight, and evaluation until
the performance is satisfactory. The length of the iteration process is inversely
proportional to the level of stability and performance robustness guaranteed by
the controller and/or to the predictive capabilities of the model.

The length, complication, and cost of these activities is exacerbated by rotorcraft-based


vehicles, which are inherently more challenging to model and difficult to control with
respect to their fixed-wing counterparts.
This thesis attempts to overcome these problems by presenting a modeling and con-
trol framework that greatly reduces the time, cost, and both human and physical re-
sources needed to design high-performance control systems for robotic helicopters. The
approach consists of

• the development of a modeling technique that allows the designer to obtain a


high-fidelity model adequate for both real-time simulation and controller design,
with few flight, ground, and wind-tunnel tests and a fair level of complication in
the dynamic equations;

• the exploitation of the predictive capabilities of the model and the robust stabil-
ity and performance guarantees of H∞ loop-shaping control theory to reduce the
number of iterations of the design/simulated-evaluation/flight-test–evaluation
procedure.

The effectiveness of this strategy will be demonstrated by designing and flight testing a
wide-envelope high-performance controller for the Carnegie Mellon University robotic
helicopter.

1.2 Modeling background

Traditionally, the full-scale helicopter community has addressed controller design and
simulation separately, with two different but complementary modeling techniques: sys-
1.2. Modeling background 5

tem-identification and first-principles modeling.

First-principles techniques typically produce both real-time and non-real-time flight


simulation models, which have non-linear dynamics, extended flight envelope simula-
tion capabilities, and which provide the capability of extracting linear models at various
trim operating points. The main drawbacks of these techniques are (1) the need for ex-
tensive knowledge of helicopter flight mechanics, (2) the difficult and inefficient manual
tuning of many physical parameters to assure a good match with flight-test data, and
(3) the low accuracy of the linear models extracted — for controller design — from the
real-time non-linear models. Highly accurate models are needed for high-bandwidth
controller design [Tischler, 1990].

The literature reports several examples of development and validation of first-prin-


ciples non-linear models for full-scale helicopters [Zhao and Curtiss, 1988, Schrage
et al., 1988, Talbot et al., 1982, Eshow et al., 1988, Ballin and Dalang-Secrétan, 1991,
Kim et al., 1993], and small-scale helicopters [Weilenmann and Geering, 1994, Gavrilets
et al., 2001]. All these models suffer from the low accuracy of the extracted linear mod-
els, especially in the off-axis responses, and thus are inappropriate for high-bandwidth
control design. Failure in capturing the off-axis response characteristics, which results
in 180-degree phase errors in the bandwidth region, is a common denominator also in
the non-linear models.

System-identification techniques use time-response data or frequency-response data


from flight tests to produce linear real-time models, which are very accurate in the
neighborhood of discrete operating points. The main problem of system identification
techniques is the difficulty in simulating the full flight envelope and thus designing a
control system for extended flight-envelope operations.

At Carnegie Mellon University (CMU), Mettler et al. [2002] have already investi-
gated the application of well known system-identification techniques used in the full-
scale helicopter community (Comprehensive Identification from FrEquency Responses,
CIFER [Tischler and Cauffman, 1992]) to the CMU Yamaha R-50 small-scale helicopter
1.2. Modeling background 6

Figure 1.1: CMU Robotic Helicopter.

shown in Figure 1.1. The research produced and validated two linear models for hover
and cruise flight of the Yamaha R-50. Kim and Tilbury [1998], Morris et al. [1994], Bruce
et al. [1998], and Shim et al. [2000] gave other examples of system identification for
small-scale helicopters, and Tischler and Cauffman [1992] and Tischler and Tomashof-
ski [2002] for full-scale helicopters. The accuracy of these linear model ranges from fair
to very good. Linear models enable the design of controllers, but then simulation with
a non-linear model is required before actual flight tests.
It is evident that a high-fidelity real-time full–flight-envelope simulation model that
integrates the characteristics of the linear models obtained from system identification,
and of the non-linear model from first-principles modeling, would be of great value in
helicopter modeling. This model would allow a user

• to simulate the helicopter in real-time and with high fidelity in its full flight enve-
lope;

• to extract, at any operating point, linear models adequate for high-bandwidth


1.3. Control background 7

controller design with linear control theory;

• to use non-linear dynamic inverse controller design techniques;

• to evaluate quickly, safely, and accurately model- or non-model based full-envelope


flight controllers.

Furthermore, if the model were a physical parametric model, it would make it possible

• to understand and evaluate the effect of changes in helicopter configuration on its


dynamics;

• to examine what are the inherent dynamic limitations of the vehicle and if and
how they can be removed by a properly designed controller;

• to analyze the controller stability and performance robustness to physical uncer-


tainties with deterministic techniques such as µ-analysis [Doyle, 1982] or statisti-
cal methods such as stochastic robustness analysis [Stengel and Ray, 1991].

1.3 Control background

The design of controllers for robotic helicopters has been documented by Weilenmann
et al. [1994], Amidi et al. [1998], Mettler et al. [2000], Koo et al. [1998], Hoffmann et al.
[1998], Prasad et al. [1999], Hovakimyan et al. [2001], Shim et al. [2000], and Sprague
et al. [2001]. A large set of design techniques, from classical control to neural-based
adaptive control, has been reported. Many of these controller-design studies have used
over-simplified models of the helicopters for design and simulation and have achieved
quite modest performance: the flight modes are limited to hover and low-speed straight
flight, or tracking accuracy decreases considerably as the speed increases and maneu-
vering flight is attempted. The absence of accurate models has also hindered the un-
derstanding of specific control problems in robotic helicopters (e.g., large system delays
1.3. Control background 8

and rotor/fuselage coupling) leading to the design of poor controllers or the incorrect
assessment of certain approaches to control design.
Weilenmann et al. [1994] compared, on an indoor bench, several multivariable tech-
niques including H∞ . H∞ control can systematically handle uncertain and multi-input-
/multi-output plants, and it promises fast and low-cost design of good-performance
flight control laws. The use of H∞ control for the design of stability augmentation and
guidance systems for fixed- and rotary-wing aircrafts has been the subject of much re-
search.
One of the effective H∞ -based design procedures is called H∞ loop shaping, which
was proposed by McFarlane and Glover [1992]. It combines the traditional notions
of bandwidth and loop gain together with modern ideas of robustness into a single
framework. Particular advantages of the method are (1) the ease with which the con-
troller weighting functions can be tuned to give an appropriate system bandwidth, and
(2) the use of a generalized stability margin that supersedes gain and phase margins.
Furthermore, controller gain scheduling and anti-windup are easily addressed in the
H∞ loop-shaping framework by expressing the H∞ loop shaping controller as an exact
plant observer plus state feedback [Sefton and Glover, 1990].
The H∞ loop-shaping approach has now been used quite widely, and a consider-
able body of experience has been accumulated. A careful examination of the liter-
ature showed that controllers designed using H∞ loop shaping are the only type of
H∞ controllers that have been flown on a piloted aircraft to date. A gain-scheduled
H∞ loop-shaping longitudinal controller flew on DERA’s3 experimental Harrier air-
craft in 1993 [Hyde, 1995]. A fixed-gain two-degree-of-freedom H∞ loop-shaping con-
troller flew on the National Research Council of Canada’s Bell 205 airborne simulator in
1997 [Smerlas et al., 1998]. A self-scheduled linear parameter-varying H∞ loop-shaping
longitudinal controller flew on DERA’s experimental Harrier aircraft in 1999 [Papageor-
giou and Glover, 2000]. In addition to the above flight tests, a fixed-gain two-degree-
3 DERA, now QinetiQ, was the UK’s Defence Evaluation and Research Agency.
1.4. Outline of contents 9

of-freedom H∞ loop-shaping controller designed for the Westland Lynx multirole com-
bat helicopter [Walker and Postlethwaite, 1996], and a fixed-gain integrated flight and
propulsion H∞ loop-shaping controller designed for DERA’s experimental Harrier air-
craft [Bates et al., 2000] have been evaluated in piloted simulation. All five control laws
were designed in a short amount of time according to AGARD, Gibson, MIL, and ADS-
33 specifications, and performed satisfactorily, reinforcing the argument that H∞ loop
shaping could indeed benefit the aeronautical industry. Nonetheless, H∞ loop shaping
has not been applied to robotic helicopters, and there have been no documented exam-
ples of flight-tested gain-scheduled H∞ loop-shaping controllers for either manned or
unmanned helicopters.

1.4 Outline of contents

What remains of this thesis is organized as follows:

Chapter 2: Integrated modeling. This chapter presents a modeling technique, called


MOdeling for flight Simulation and Control Analysis (MOSCA), which integrates
first-principles and system-identification techniques. The approach consists of
developing a medium-complexity, parametric, physics-based, non-linear model,
and automatically tuning its parameters exercising system-identification methods
on extracted linear models in the frequency domain. The accuracy of the model is
determined through comparisons with flight tests both in the frequency and time
domains.

Chapter 3: Robust hover control. Chapter 3 introduces the H∞ loop shaping design
method and shows how H∞ loop shaping can be used to design a hover controller
for a robotic helicopter. The controller stability and performance robustness is
evaluated with some standard aeronautical-industry requirements and by flying
several complex maneuvers off of the hover design point. Finally, robustness un-
der varying loading conditions is evaluated using the parametric model.
1.4. Outline of contents 10

Chapter 4: Robust full-envelope flight control. This chapter presents a scheduling pro-
cedure for a H∞ loop shaping controller. The procedure makes use of current en-
gineering practice in implementing scheduled controllers, but it is formalized to
a level that makes it easy to automate the whole process from establishing the
variables that need to be scheduled to cover the target flight envelope, to eval-
uating the minimum number of fixed operating-point controllers needed to sat-
isfy the stability margin specification throughout the flight envelope. Compliance
with some standard aeronautical requirements is then checked, and it is shown
that satisfying robust stability using the H∞ loop-shaping stability margin metric
also delivers excellent closed-loop performance. An analysis of the observer plus
state-feedback structure of the H∞ loop shaping controller is used to justify and
gain physical insight into the excellent performance. The scheduled controller is
then evaluated in real flight with several steps and maneuvers at different speeds.

Chapter 5: Conclusions. This chapter summarizes the main contributions and provides
some suggestions for future work.
Chapter 2
Integrated modeling

2.1 First-principles non-linear model . . . . . . . . . . . . . . . . . . . . . 12

2.2 Linear models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.3 Model and flight-test frequency responses . . . . . . . . . . . . . . . . 28

2.4 Identification of physical parameters . . . . . . . . . . . . . . . . . . . 29

2.5 Reduced-order linear models . . . . . . . . . . . . . . . . . . . . . . . . 32

2.6 MOSCA-model evaluation . . . . . . . . . . . . . . . . . . . . . . . . . 33

2.7 Conclusions on integrated modeling . . . . . . . . . . . . . . . . . . . 35

This chapter presents a novel modeling technique called MOdeling for Flight Simu-
lation and Control Analysis (MOSCA). MOSCA combines the benefits of first-principles
and system-identification techniques. The technique consists of (1) generating an ade-
quate first-principles non-linear model, and (2) using global optimization methods and
extracted linear models to tune automatically selected uncertain physical parameters
of the non-linear model to match frequency responses from flight tests. MOSCA yields
non-linear models that are adequate for simulation over a large portion of the heli-
copter’s flight envelope, and high-fidelity linear models needed for high-bandwidth
controller design.

11
2.1. First-principles non-linear model 12

2.1 First-principles non-linear model

As shown in Figure 2.1, the first step of the MOSCA approach is to obtain a first-
principles non-linear model of the robotic helicopter of interest. The CMU Yamaha
R-50 (see Figure 1.1) has been used in this thesis as the test vehicle.

Control Inputs
First−Principles Non−Linear Responses
Parametric Model

Linearization

Hover Flight Forward Flight


Linear Model Linear Model

Frequency Responses Frequency Responses

Hover Flight Tests Forward Flight Tests


Frequency Responses Frequency Responses
Optimization

Updated Physical Parameters

Figure 2.1: Flow chart of MOSCA integrated approach to modeling.

The R-50 is a small-scale helicopter designed to be remotely controlled by a human


pilot for crop-dusting operations. It has a two-bladed rotor of 1.535 m radius with a
Bell-Hiller stabilizer bar. The dry weight is 44 kg. CMU’s instrumentation package for
autonomous flight (IMU, differential GPS, magnetic compass, laser range sensor, and
CPU with I/O board) adds 23 kg.
The non-linear model generated for the CMU R-50 represents explicitly the follow-
ing: (1) fuselage, (2) gravity forces and Euler angles, (3) main rotor, (4) dynamic inflow,
(5) actuators, (6) stabilizer bar, (7) tail rotor, and (8) aerodynamic forces. The modeling
2.1. First-principles non-linear model 13

of the subsystems is based on work by others [Johnson, 1994, Zhao and Curtiss, 1988,
Schrage et al., 1988, Talbot et al., 1982, Eshow et al., 1988, Ballin and Dalang-Secrétan,
1991, Kim et al., 1993, Pitt and Peters, 1981, Keller, 1996] with the relevant difference
given by the non-American left-blade–advancing rotation of the main rotor, which ne-
cessitated the derivation from scratch of all the equations of motion. What follows
shows the modeling of the mentioned subsystems.

2.1.1 Fuselage

The dynamic model of the fuselage uses rigid-body equations. The translational and
rotational velocities of the fuselage in the body frame are (see Figure 2.2)
   
u p
   
vb =  v  , ωb =  q  (2.1)
   
   
w r

where u, v, w are the forward, lateral, and down velocities, and p, q, r are the roll, pitch,
and yaw angular rates. The Newton-Euler equations for the fuselage expressed in the
body frame are
1
v̇b = F − ωb × v b (2.2)
m b
ω̇b = J −1 ( M b − (ωb × Jωb )) (2.3)

where m is the mass of the fuselage,


   
F Mb x
 bx   
F b =  Fby  , and M b =  Mb y  (2.4)
   
   
Fbz Mb z

are the applied forces and moments, and


 
J 0 Jxz
 xx 
J=0 Jyy 0 (2.5)
 
 
Jzx 0 Jzz
2.1. First-principles non-linear model 14

y, v, q

x, u, p

x, u, p

z, w, r

Figure 2.2: Body frame with translational (u, v, w) and rotational (p, q, r) velocities.

is the inertia matrix. F b and M b are

Fb = F mr + F tr + F g + F ae
(2.6)
M b = M mr + M tr + M g + M ae

that is, they are the sum of the forces and moments due to the main rotor (F mr , M mr ),
tail rotor (F tr , M tr ), gravity (F g ), and fuselage aerodynamic forces (F ae , M ae ).

2.1.2 Gravity forces and Euler angles

The expression of gravity forces in the body frame is


 
− sin θ
 
F g = mg  sin φ cos θ  (2.7)
 
 
cos φ cos θ
2.1. First-principles non-linear model 15

where ψ, φ, and θ are the yaw, roll, and pitch Euler angles. Euler angles represent the
three consecutive rotations that transform the earth frame into the body frame (see Fig-
ure 2.2). The earth frame is fixed and has the x, y, and z axes parallel to the North, East,
and Down directions respectively. The body frame is parallel to the earth frame for ψ, θ,
and φ = 0. The equation for the Euler angles variations in time is
   
φ̇ 1 tan θ sin φ tan θ cos φ
   
 θ̇  = 0 cos φ − sin φ  ωb (2.8)
   
   
ψ̇ 0 sin φ/ cos θ cos φ/ cos θ

2.1.3 Main rotor

The motion of the rotor blades generates the main-rotor forces and moments transmit-
ted to the fuselage. The main assumptions made to derive the dynamic equations of
the main rotor are the following: (1) the blades are rigid in bending and torsion, (2) the
flapping angle is small, (3) there is no reverse flow region and no blade stall, (4) the tip
loss factor is equal to 1, and (5) compressibility effects are ignored.
These assumptions grant validity of the model up to an advance ratio µ of 0.3 [Tal-
bot et al., 1982]. The advance ratio µ is defined as the ratio of the main-rotor in-plane
velocity to the blade-tip speed:
V cos α
µ= (2.9)
ΩR
where V is the airspeed, α is the angle of attack of the rotor, Ω is the main-rotor ro-
tational speed, and R is the rotor radius. For the R-50, µ = 0.3 corresponds to V ≃
40 m/s (144 Km/h), which exceeds the R-50’s maximum speed. Thus, the model is
adequate to represent the full flight envelope of the vehicle.
The first step to obtain the dynamic equations of the main rotor is to derive the
expression of the sum of all the moments on a blade with respect to the flapping hinge.
The development of such an expression is simplified by the use of the hub-wind frame.
The hub-wind frame has its origin at the center of the main-rotor hub, the z axis aligned
2.1. First-principles non-linear model 16

with the main-rotor shaft with positive direction down, the x axis aligned with the
relative wind component perpendicular to the shaft, and the y axis oriented as required
to form a right-handed orthogonal frame.
The moment components on a single blade are due to the aerodynamic forces, cen-
trifugal forces, blade inertia, Coriolis acceleration, restraint moment at the hinge, weight,
and fuselage vertical and angular accelerations.
The dynamic equation for a blade has the following form:
Z R−e
1
β̈ = Mh dr (2.10)
Jβ 0

where β is the out-of-plane flapping angle, Jβ the blade moment of inertia, and e is
the blade hinge offset from the center of the rotor. Mh dr is the sum of the moment
components with respect to the flapping hinge calculated at each “strip” dr along the
blade (strip theory). Mh is a complicated function of the rotor physical parameters
(e.g., blade twist, hinge spring, blade chord, etc.), flapping angle β and flapping angle
velocity β̇, blade velocity of rotation Ω, blade azimuth angle Ψ, fuselage velocities and
accelerations (translational and rotational), induced velocity, and blade pitch angle.
Once the expression for Mh is obtained, β (and its derivatives) is substituted with
the following expression (and its derivatives):

β(t) = a0 (t) − a1 (t) cos Ψ + b1 (t) sin Ψ; (2.11)

Retaining only the first harmonic terms in Equation 2.10 after the former substitution,
allows the description of the flapping motion of the blades with three elementary mo-
tions of the tip of the blades (tip-path plane). In fact, a0 (t), a1 (t), and b1 (t) represent
respectively the coning, the pitching, and the rolling motion of the tip-path plane.
A non-linear first-order system of differential equations for the flapping coefficients
can be obtained by separating the constant terms and the terms with cos Ψ and sin Ψ in
Equation 2.10,  
0 I
ȧmr =   amr + f
mr (2.12)
−K mr − Dmr
2.1. First-principles non-linear model 17

where amr = [a0 a1 b1 ȧ0 ȧ1 ḃ1 ]T and K mr , Dmr , and f mr depend on the rotor physical
parameters, the blades’ velocity of rotation, the induced velocity, the blade pitch angle,
and the fuselage velocities and accelerations (translational and rotational). In detail, the
elements of K mr are
e mβ 1

1 1 1
K mr (1, 1) = + γ ε K l + 1 − γ ε2 µ2 K l + γ ε µ2 K l − γ K l −
Jβ 6 8 4 8


1
γ µ2 K l Ω 2 + (2.13)
8 Jβ
 
1 1 2
K mr (1, 2) = − γ ε µ + γ ε µ Ω2 (2.14)
8 4
 
1 1
K mr (1, 3) = − γ µ K l + γ ε µ K l Ω2 (2.15)
6 4
 
1 1
K mr (2, 1) = − γ µ + γ ε µ Ω2 (2.16)
6 4
e mβ

1 1 1 1
K mr (2, 2) = γ ε K l + γ ε µ2 K l − γ K l + − γ ε2 µ2 K l −
6 8 8 Jβ 16


1
γ µ2 K l Ω 2 + (2.17)
16 Jβ
 
1 2 1 2 1 2 1 1 1
K mr (2, 3) = − γε + γεµ − γµ − γ ε µ − γ + γ ε Ω2 (2.18)
2 2
4 8 16 16 8 3
 
1 1
K mr (3, 1) = − γ µ K l + γ ε µ K l Ω2 (2.19)
3 2
 
1 1 1 1 1 1
K mr (3, 2) = − γ ε2 µ2 − γ µ2 + γ ε µ2 + γ ε2 − γ ε + γ Ω2 (2.20)
16 16 8 4 3 8
e mβ

1 1 3 3
K mr (3, 3) = γ ε Kl − γ Kl + − γ ε2 µ2 K l + γ ε µ2 K l − (2.21)
6 8 Jβ 16 8


3
γ µ K l Ω2 +
2
(2.22)
16 Jβ
the elements of Dmr are
 
1 1 2 1
Dmr (1, 1) = − γε+ γε + γ Ω (2.23)
3 4 8
Dmr (1, 2) = 0 (2.24)
 
1 2 1 1
Dmr (1, 3) = γε µ+ γµ− γεµ Ω (2.25)
4 12 4
Dmr (2, 1) = 0 (2.26)
2.1. First-principles non-linear model 18

 
1 1 2 1
Dmr (2, 2) = − γε+ γε + γ Ω (2.27)
3 4 8
Dmr (2, 3) = −2 Ω (2.28)
 
1 2 1 1
Dmr (3, 1) = γε µ+ γµ− γεµ Ω (2.29)
2 6 2
Dmr (3, 2) = 2Ω (2.30)
 
1 1 2 1
Dmr (3, 3) = − γε+ γε + γ Ω (2.31)
3 4 8
and the elements of f mr are

1 1 1 1
f mr (1) = − ā0 + γ µ θlonw − γ ε θcol + γ ε λ0 + γ µ θa1s +
6 6 4 6
1 1 1 1 e m β ā0
γ θcol + γ θtw − γ ε µ θlonw + γ ε µ λs − +
8 10 4 8 Jβ
1 1 1 1 1
γ µ2 θcol − γ ε θtw + γ µ2 θtw − γ ε µ θa1s + γ ε2 µ2 θcol −
8 8 12 4  8
1 1 1 1
γ ε µ2 θcol − γ ε µ2 θtw − γ µ λs − γ λ0 Ω 2 +
4 8 12 6

1 1 1 γ ε wh
γ ε µ p cos β w − γ µ q sin βw − +
8 12 4 R

1 1 γ wh 1
γ ε µ q sin βw + − γ µ p cos β w Ω +
8 6 R 12
m β ẇ m β g cos φ cos θ m β uq m β pv
− − + (2.32)
Jβ Jβ Jβ Jβ

1 1 1 1
f mr (2) = − γ ε 2 µ 2 θ b1 s − γ µ2 θlatw − γ µ 2 θ b1 s − γ ε2 µ2 θlatw −
16 16 16 16
1 1 1 1 1
γ ε λc + γ λc − γ θlatw − γ θb1s − γ ε µ ā0 +
6 8 8 8 4 
1 1 1 1 1
γ µ ā0 + γ ε µ θlatw + γ ε θlatw + γ ε θb1s + γ ε µ θb1s Ω2 +
2 2
6 8 6 6 8

1 1
2 q sin β w − γ q cos βw + γ p sin β w + 2 p cos β w +
8 8
e m β p cos β w e m β q sin βw

1 1
2 +2 + γ ε q cos βw − γ ε p sin βw Ω −
Jβ Jβ 6 6
q̇ cos βw + ṗ sin β w (2.33)

1 1 1 3 1
f mr (3) = γ θa1s + γ µ θcol − γ ε µ θtw − γ ε µ2 θlonw − γ ε µ θcol −
8 3 3 8 2
1 2 1 1 3 1
γ ε µ λ0 − γ ε θa1s − γ λs − γ ε µ2 θa1s + γ θlonw +
4 6 8 8 8
2.1. First-principles non-linear model 19

3 1 3 3
γ µ 2 θ a1 s + γ ε µ λ 0 + γ ε 2 µ 2 θ a1 s + γ ε2 µ2 θlonw +
16 2 16 16 
1 3 1 1 1
γ ε λs + γ µ θlonw + γ µ θtw − γ µ λ0 − γ ε θlonw Ω2 +
2
6 16 4 4 6
e m β q cos βw

1 γ µ wh
−2 − 2 q cos βw + + 2 p sin βw +
Jβ 4 R
1 1 γ ε µ wh 1 e m β p sin β w
γ ε p cos β w − − γ q sin βw + 2 +
6 2 R 8 Jβ
γ ε2 µ w h 1

1
γ ε q sin βw + 1/4 − γ p cos βw Ω − ṗ cos βw −
6 R 8
q̇ sin βw (2.34)

In the preceding expressions m β is the blade mass moment about the flapping hinge,
Kl the blade pitch/flap coupling ratio, ε = e/R, K β the hinge-spring stiffness, γ the
blade Lock number, ā0 the blade pre-coning angle, λ0,c,s the inflow components, βw
the sideslip angle (the angle between the inertial velocity vector and the relative wind
vector in the x-y plane of the hub-wind frame), and wh the velocity of the rotor hub
in the body frame. The subscripted θs represent the various contributions to the main-
rotor blade pitch angle θmr ; so, for example, θtw is the pitch due to the blade twist, and
θlonw is the pitch due to the longitudinal cyclic input in the hub-wind frame. The blade
Lock number γ is defined as
ρ c α c R4
γ= (2.35)

where ρ is the air density, c is the blade cord, and cα is the blade lift-curve slope.
The dynamic inflow λ is defined as the ratio between the induced velocity and
blade-tip velocity. The representation of the rotor inflow is the one suggested by Pitt
and Peters [1981] with the addition of the wake-distortion term proposed by Keller
[1996]. Pitt and Peters [1981] assumed that the distribution of the inflow λ is
r r
λ(t, r, Ψ) = λ0 (t) + λs (t) sin Ψ + λc (t) cos Ψ (2.36)
R R
The first-order differential equation that describes the inflow dynamics is then

L M λ̇ + λ = L c + k (2.37)
2.1. First-principles non-linear model 20

where
     
λ
 0
c
 T  0 
 pw 
λ =  λs  , c =  cL  , and  − Kr Ω 
k= (2.38)
    
qw 
    
λc cM Kr

represent the inflow components, the thrust and moment coefficients due to aerody-
namic forces on the rotor, and the wake distortion terms respectively. The expressions
for L and M are
8
 
 3π Ω 0 0 
 
 16 
M= 0 − 0  (2.39)

 45 π Ω 

 16 
0 0 −
45 π Ω

1 15 π χ
 
0 tan

 2 vt 64 vm 2 

 4 
L=
 0 − 0 
(2.40)
(1 + cos χ) vm

 
 
 15 π 4 cos χ 
0 −
64 vm (1 + cos χ) vm

where χ is the rotor-wake skew angle,

wh
q
vt = µ2 + ( λ0 − µ3 )2 , with µ3 = , (2.41)
ΩR

and
µ2 + (2 λ0 − µ3 )(λ0 − µ3 )
vm = . (2.42)
vt

For a twisted rotor blade, M (1, 1) = 128/(75πΩ). Both the dynamic inflow and
the Keller’s wake-distortion terms allow a better representation of the coupling effects
between the rotor dynamics and the fuselage.

The forces and moments transmitted by the rotor to the fuselage are calculated using
2.1. First-principles non-linear model 21

the elementary strip forces and moments. For example, Fmrz (main-rotor thrust) in
 
Fmrx
 
F mr =  Fmry  (2.43)
 
 
Fmrz

is
Z 2πZ R−e
N
Fmrz = Fz dr dΨ (2.44)
2π 0 0

where N is the number of blades, and Fz is the sum of all the forces acting on a rotor
blade in the z direction and calculated at each strip dr along the blade.
F mr and M mr , and Equations 2.12 and 2.37, all depend on blade pitch angle θmr ,
which is effected by the stabilizer-bar flapping motion and the actuator positions. The
following section describes these two subsystems.

2.1.4 Actuators and stabilizer bar

In a helicopter without a stabilizer bar, four actuators are used to change the pitch angle
θmr of the main rotor blades and the pitch angle θtr of the tail rotor blades to generate
forces and moments on the fuselage. The expression for θmr is

θmr = θcol + θlon sin Ψ + θlat cos Ψ (2.45)

where the subscripts col, lon, and lat indicate the blade-pitch components due to the
collective, longitudinal, and lateral inputs (δcol , δlon , and δlat ) respectively. Second-order
systems approximate satisfactorily the dynamics of the actuators. For the collective
control, for example, the equation is

θ̇col = Acol θcol + Bcol δcol (2.46)

where  
θcol
θcol =   (2.47)
θ̇col
2.1. First-principles non-linear model 22

and δcol is the input to the actuator. The matrices in Equation 2.46 are
   
0 1 0
Acol =  , Bcol =  . (2.48)
−ωn2 col −2 ζ col ωncol ωn2 col

Frequency-domain identification is used to determine the natural frequency ωn and


damping ζ of each actuator from sweep experiments.
When the stabilizer bar is present (as in the R-50), the expression for θmr is

θmr = θcol + θlon sin Ψ + θlat cos Ψ + ks ( a1s sin Ψ + b1s cos Ψ) (2.49)

where a1s and b1s represent the flapping motion of the stabilizer bar, and the constant
ks models the mechanical gearing between this flapping motion and the main-rotor
blades’ pitch angle.
The purpose of the stabilizer bar is to help the human pilot in controlling the he-
licopter. Usually, it is present on small-scale helicopters, which have faster dynamics
compared to full-scale ones, and acts as a lagged-rate feedback system [Miller, 1950].
The stabilizer bar on the R-50 has two peripheral paddles that produce negligible
forces on the fuselage due to the their small surface. Its dynamics are modeled as an
additional rotor with a set of equations similar to Equation 2.12:
 
0 I
ȧs =   as + f
s (2.50)
−K s − Ds

Since the stabilizer bar is teetered there is no coning term in Equation 2.50 and as =
[a1s b1s ȧ1s ḃ1s ]T . The expressions for K s , Ds , and, f s are much simpler than the corre-
sponding ones for the main rotor:

γs µ 2 R2i γs R4i
     
0 −1 + −1 
16 R2o 8 R4o
K s = Ω2 


 γs µ 2 R 2
 
γs R4i
  
i
−1 − −1 0
16 R2o 8 R4o
(2.51)
2.1. First-principles non-linear model 23

R4
   
γs
1 − 4i −2
 8 Ro 
Ds = Ω   (2.52)
R4i
 
 γs 
2 1−
8 R4o

4 2
1 1 γs θslonw Ri 1 γs µ2 θslonw Ri
f s ( 1) = − γs θslonw + + −
8 8 Ro 4 16 Ro 2
1 γs q cos βw Ri 4

1 1
γs µ θslonw Ω2 + − γs q cos βw +
2
+
16 8 8 Ro 4
!
1 1 γs p sin β w Ri 4
γs p sin β w + 2 p cos β w + 2 q sin β w − Ω−
8 8 Ro 4
q̇ cos βw + ṗ sin β w (2.53)
4
1 γs µ λ̄s Ri 2 1 1 γs θlonsw Ri 1
f s ( 2) = − 2
+ γ s µ λ̄ s − 4
+ γs θlonsw −
4 Ro 4 8 Ro 8
!
2
3 γs µ2 θlonsw Ri 3 2 2 1 γs q sin βw Ri 4
+ γ s µ θ lonsw Ω + +
16 Ro 2 16 8 Ro 4
1 γs p cos β w Ri 4 1 1
4
+ 2 p sin βw − γs q sin βw − γs p cos βw −
8 Ro 8 8
2 q cos β w ) Ω − ṗ cos β w − q̇ sin βw (2.54)

Many of the variables in the former equations have the same meaning as for the main
rotor except for Ri and Ro that are the radial position of the inward and outward bound-
aries of the paddles, and λ̄s that is the static inflow ratio. The stabilizer-bar blade-pitch
angle is
θs = θslon sin Ψ + θslat cos Ψ. (2.55)

where θslon and θslat represent the blade pitch due to the longitudinal and lateral inputs
respectively.

2.1.5 Tail rotor

The tail rotor has smaller dimensions compared to the main rotor. Its dynamics are
much faster so they can be fairly represented by a steady-state solution of the flapping
2.1. First-principles non-linear model 24

equation. This is achieved by zeroing the derivatives of the blade flapping in a flapping
equation similar to the one for the main rotor (Equation 2.12):
 
0 I
ȧtr =   atr + f
tr (2.56)
−K tr − Dtr

where atr = [ ȧ0tr ȧ1tr ḃ1tr a0tr a1tr b1tr ]T . Eliminating ȧtr and ȧ0tr , ȧ1tr , and ḃ1tr from atr ,
Equation 2.56 becomes:
−1
ātr = K tr f tr (2.57)

where ātr = [a0tr a1tr b1tr ]T . The expression for K tr and f tr are very similar to the ones
given for the main rotor, the only difference being a steady-state representation for the
induced velocity as for the stabilizer bar. Forces and moments (F tr , M tr ) are calculated
as for the main rotor (Equation 2.44). The pedal actuator is used to change the blade
pitch angle θtr :
θ̇tr = Atr θtr + Btr δped (2.58)

where δped is the pedal input and the entries in Atr and Btr (see Equation 2.48) are
obtained with system identification.

2.1.6 Aerodynamic forces

The aerodynamic forces and moments (F ae , M ae ) acting on the fuselage are modeled as
proposed in Talbot et al. [1982]. It can be assumed that F ae and M ae depend on angle of
attack α and/or sideslip β w . For example, the expression for the drag force Faex is

Faex = qf (d0α + d1α α + d2α α2 + d2βw β2w ) (2.59)

where qf is the dynamic pressure and d0α , d1α , d2α , and d2βw are experimental aerody-
namic coefficients. Talbot et al. [1982] give aerodynamic coefficients for a Cobra AH-1G.
Given the absence of wind tunnel tests for the Yamaha R-50, the use of scaling rules pro-
vides an estimate of the coefficients in Equation 2.59 from the ones of the AH-1G.
2.1. First-principles non-linear model 25

2.1.7 System of equations

Summing the forces and moments from all the subsystems gives F b and M b (see Equa-
tion 2.6). Combining Equations 2.2, 2.3, 2.8, 2.12, 2.37, 2.50, Equation 2.46 (three of them
for collective, longitudinal, and lateral inputs), and Equation 2.58 results in the system
of non-linear equations that describes the coupled dynamics of the helicopter:

ẋ = g ( ẋ, x, u, t) (2.60)

The 30-dimensional state vector x is

x = [ u v w p q r φ θ ψ a0 a1 b1 ȧ0 ȧ1 ḃ1 λ0 λs λc


(2.61)
T
a1s b1s ȧ1s ḃ1s θlon θ̇lon θlat θ̇lat θcol θ̇col θtr θ̇tr ]

and the input vector is


 
δlon
 
 δlat 
 
u=
 
 (2.62)
 δcol 
 
δped
Identifying all the terms containing ẋ in g ( ẋ, x, u, t), the Equation 2.60 is expressed
as [Kim et al., 1993]
ẋ = Γ(t) ẋ + g ( x, u, t) (2.63)

which is equivalent to

ẋ = ( I − Γ(t))−1 g ( x, u, t) ≡ f ( x, u, t) (2.64)

For simulation, an additional non-linear system of equations is needed to express


the dependence of the desired output variables y (such as forward ground velocity V
or flight-path angle γ) from the state vector:

y = h( x, u, t) (2.65)

In case the desired outputs are all the helicopter states, h( x, u, t) is simply x.
2.2. Linear models 26

2.2 Linear models

After obtaining the non-linear model, the second step in the MOSCA procedure (see
Figure 2.1) is to extract linear models at the operating points for which flight-test fre-
quency responses are available. Notice that, for simplicity, Figure 2.1 shows only two
operating points, but the approach can be extended to any number of them as long as
the corresponding flight-test data are available.

2.2.1 Trimming

To extract the linear models, the non-linear model must be trimmed first. The trim state
is obtained by solving (see Equation 2.64)

f ( x, u) = 0 (2.66)

that is a non-linear system of 30 equations with 34 unknowns (30 for x plus 4 for u).
To have a square system, which can be solved efficiently utilizing the Powell Hybrid
Method, four equations must be added. These four equality constraints specify the
operating point at which the trim is desired. In fact, a particular flight condition is
specified with V, γ, and βw , and it is independent of ψ. V is the ground speed of the
helicopter, γ the flight-path angle, and βw the sideslip angle. For example, forward
straight level flight at 10 m/s is specified with

V = 10

γ = 0

βw = 0

ψ = 0
2.2. Linear models 27

and sideways straight level flight at the same velocity with

V = 10

γ = 0
π
βw =
2
ψ = 0

V, γ, and β w are related to the state vector x; constraining these variables translates into
constraining the state vector. Hence, it is possible to derive the missing equations to
solve Equation 2.66. The four equations are:

ub = Vx cos θ cos ψ + Vy cos θ sin ψ − Vz sin θ (2.67)

vb = Vx (sin φ sin θ cos ψ − cos φ sin ψ) + (2.68)

Vy (sin φ sin θ sin ψ + cos φ cos ψ) + Vz sin φ cos θ

wb = Vx (cos φ sin θ cos ψ + sin φ sin ψ) + (2.69)

Vy (cos φ sin θ sin ψ − sin φ cos ψ) + Vz cos φ cos θ

ψ = 0 (2.70)

where

Vx = V cos γ cos βw (2.71)

Vy = V cos γ sin βw (2.72)

Vz = −V sin γ (2.73)

It is also possible to obtain solutions for other flight configurations such as coordi-
nated turn, instantaneous constant pitch rate (θ̇), and instantaneous constant roll rate
(φ̇). For example, the trim for a forward coordinated level turn at V = 20 m/s and
turning velocity (ψ̇) of 30 deg/s is specified by imposing the same conditions as for
forward straight level flight for V, γ, β w , and ψ, and by requesting that the value of
the state-derivative vector component ψ̇ in the system of Equations 2.66 is equal to 30
deg/s instead of 0 deg/s.
2.3. Model and flight-test frequency responses 28

2.2.2 Linearization

Linear models at the calculated trim are derived with a first order Taylor series expan-
sion of Equations 2.64 and 2.65.
   
∂g ∂g
∆ ẋ ≃ ∆x + ∆u (2.74)
xtrim ∂x ∂u utrim
   
∂h ∂h
∆y ≃ ∆x + ∆u (2.75)
∂x xtrim ∂u utrim

where

∆ ẋ = ẋ − ẋtrim , ∆x = x − xtrim (2.76)

∆y = y − ytrim , ∆u = u − utrim (2.77)

The partial derivatives in Equations 2.74, 2.75 are computed using a two-point finite-
difference scheme with an adaptive step size.
With x̃˙ = ∆ ẋ, x̃ = ∆x, ỹ = ∆y, and ũ = ∆u, Equations 2.74, 2.75 can be rewritten in
the standard form for linear systems:
    
x̃˙ A B x̃
 =   (2.78)
ỹ C D ũ

2.3 Model and flight-test frequency responses

The linear-model frequency responses can be calculated with

G (s) = C (sI − A)−1 B + D. (2.79)

MOSCA requires also frequency responses from flight tests (see Figure 2.1), which were
calculated using CIFER with the same data set used by Mettler et al. [2002]. This data
set consists of several sweeps (used for frequency responses calculations) and doublets
(used for time domain verification) on all the inputs for two different flight conditions:
hover and cruise flight at 15 m/s.
2.4. Identification of physical parameters 29

2.4 Identification of physical parameters

It is a well known problem that low- and medium-complexity first-principles models


(the model assembled in the former sections belongs to these categories) are not accu-
rate enough to be used for controller design. This is particularly true for rotorcrafts that
exhibit far higher complexity, coupling, and non-linearity than fixed-wing aircraft.
The flight-control community has been using the following approaches to tackle this
problem:

1. Extensive and cumbersome manual tuning of the large number of physical pa-
rameters in the non-linear model to improve their overall matching capabilities.

2. Extraction of linear models at the target design points, and direct automatic tun-
ing of the A and B matrices in Equation 2.78 with the use of flight-test data and
system identification.

3. Development of physics- and non-physics-based parametric linear models suc-


cessively tuned with flight-test data and system identification.

4. Development of high-complexity non-linear models (often non-real time) to in-


crease the matching capabilities without extensive tuning.

Each one of these approaches has its own drawbacks. The intent of the MOSCA
approach is to automate the practically infeasible manual tuning process in approach 1,
to deliver a high-fidelity real-time non-linear model adequate for linear and non-linear
model-based controller design.
Automatic tuning can be achieved by identifying the uncertain parameters in the
time or in the frequency domain. The use of time domain identification for non-linear
models, presents an important shortcoming: the identification of the uncertain parame-
ters is severely biased by the necessity of identifying also both the trimming conditions
and the initial dynamic states of the model. In fact, the parameters of the non-linear
2.4. Identification of physical parameters 30

model, the trim, and the initial states all contribute to the dynamic response of the sys-
tem, which makes it difficult, if not impossible, to identify correctly the system parame-
ters. Another problem with time domain identification arises if the linear or non-linear
system is unstable as in the case of a helicopter. Instability creates problems both in
executing long flight-test runs to include the low-frequency dynamics, and during the
simulation phase for the identification.
For linear systems, frequency domain techniques eliminate the need to identify the
trim and the initial states, and the collection of flight data is easily accomplished with
sweeps over the frequency range of interest. These are the main reasons why identi-
fication from frequency responses is so popular in the aeronautical industry with ro-
bust software packages such as CIFER [Tischler and Cauffman, 1992]. Frequency do-
main identification has not been used directly to identify the parameters of a non-linear
model because for non-linear cases a frequency response represents only the best lin-
ear system approximating the dynamic behavior in the neighborhood of a particular
operating point.
MOSCA uses the frequency domain to identify the parameters of a non-linear model
in an intuitive and physically meaningful way. Suppose that the NFR flight-test fre-
quency responses G for each input/output pair of interest are available at each of the
NOP operating points. MOSCA identifies the parameters of the non-linear model by
(1) extracting linear models (Equation 2.78) at the NOP operating points, (2) calculating
their frequency responses (Equation 2.79) and selecting the ones of interest, and (3) min-
imizing a single cost function that includes the mismatches between the NFR calculated
and flight-test frequency responses at each one of the NOP operating points. Figure 2.1
illustrates this procedure assuming that only flight tests for hover and cruise flight are
available.
The expression for the cost function to minimize is

NOP NFR
J ( p) = ∑ ∑ wκ κTn,i ( p)W n,iκn,i ( p) + wχ χTn,i ( p)W n,i χn,i ( p) (2.80)
n =1 i =1
2.4. Identification of physical parameters 31

where p is the vector of selected parameters of the non-linear model that we want
to identify, NFR is the number of frequency responses, NOP is the number of operat-
ing points, κn,i is the magnitude error between flight-test and calculated frequency re-
sponses
κn,i = |Gn,i (s)data | − |Gn,i (s)model | (2.81)

and χn,i is the phase error between flight-test and calculated frequency responses

χn,i = 6 Gn,i (s)data − 6 Gn,i (s)model (2.82)

W n,i is a weighting matrix to emphasize most accurate data (i.e., data with high coher-
ence), and wκ and wχ are weights that determine the trade-off between magnitude and
phase errors [Tischler and Cauffman, 1992].
A single cost function results in a unique set of optimized model parameters, which
makes the model accurate in the flight-envelope region covered by the flight exper-
iments. MOSCA allows the inclusion of any number of operating points as long as
flight-test frequency responses are available. MOSCA uses a simplex-method algorithm
[Rowan, 1990] to minimize J ( p ).
An analogy to a geometric problem in two dimensions helps to understand better
how MOSCA works. The process (tuning a non-linear parametric model to flight tests
in the frequency domain) is analogous to finding the values of a parametric curve (the
non-linear model) such that the curve is tangent to N known straight lines (frequency
responses from flight tests that describe a linear approximation of the non-linear sys-
tem) at a number N of known points (the operating points). To form the cost function
one would compute the derivatives (extract the linear models) of the parametric curve
(the non-linear model) at each of the N points, and calculate the errors with the known
tangents (the frequency responses from flight tests).
The key to the success of this approach is that the non-linear parametric model must
capture the salient characteristics of the given dynamic system. This allows the solution
to be a physically meaningful set of parameters .
2.5. Reduced-order linear models 32

2.5 Reduced-order linear models

The full-state linear models contain high frequency dynamics that may be less impor-
tant in capturing the dynamic behavior of the helicopter in the crossover region of in-
terest for control. Reduced-order linear models can be derived from Equation 2.78 in
two steps. The state vector x̃ is partitioned in x̃1 (states to be kept) and x̃2 (states to be
eliminated):     
x̃˙1 A A12 B1 x̃
   11  1 
 ˙  
 x̃2  =  A21 A22 B2   x̃2  (2.83)
 
    
ỹ C1 C2 D ũ
Then x̃2 , which contains the faster dynamics, can be eliminated substituting its dynamic
equation with the respective steady-state solution; that is, x̃˙2 is set to zero,

0 = A21 x̃1 + A22 x̃2 + B2 ũ

and x̃2 is solved for x̃1 :


−1 −1
x̃2 = − A22 A21 x̃1 − A22 B2 ũ (2.84)

Hence, the reduced-order model is


    
−1 −1
x̃˙1 A11 − A12 A22 A21 B1 − A12 A22 B2 x̃1
 =   (2.85)
−1 −1
ỹ C1 − C2 A22 A21 D − C2 A22 B2 ũ
The reduced-order model used for comparisons in the next section was obtained by
eliminating the following states:

x2 = [ a0 ȧ0 ȧ1 ḃ1 λ0 λs λc ȧ1s ḃ1s ]T (2.86)

This results in a model with states

x1 = [ u v w p q r φ θ ψ a1 b1 a1s b1s θlon θ̇lon θlat θ̇lat θcol θ̇col θtr θ̇tr ]T (2.87)

Except for the actuators, this reduced model has basically the same dimensions of the
CMU R-50 linear model identified with CIFER by Mettler et al. [2002]. The reduced
linear model describes the helicopter with rigid body dynamics for the fuselage, and
first-order tip-path–plane dynamics for the main rotor and the stabilizer bar.
2.6. MOSCA-model evaluation 33

2.6 MOSCA-model evaluation

This section reports the set of identified model parameters and comparisons among full-
state linear models, reduced-order linear models, and flight-test data in the frequency
and time domains. Results from the non-linear model are not presented since, as ex-
pected, they were always marginally better than the results from the linear models at
every operating point.

2.6.1 Identified physical parameters

Table 2.1 reports some of the measured and optimized values for a typical MOSCA
run for the CMU R-50. The optimized parameters are close to the measured ones. The
pitch moment of inertia exhibits the largest variation (30%); this is probably due to
an incorrectly measured value1 , or to the transfer of the uncertainty of other physical
parameters to the pitch moment of inertia.

The fact that the identified physical parameters do not lose their physical meaning
is a key result, which shows the power of the integration of first-principles modeling
and system identification with global optimization in the frequency domain. Previous
attempts [Hong and Curtiss, 1993] in tuning a non-linear helicopter model with the use
of time responses have always led to unrealistic values of the model’s parameters. Fur-
thermore, Hong and Curtiss [1993] performed the optimization only for hover, which
does not guarantee that the optimized set of values is able to increase the fidelity of the
original non-linear model over a wider portion of the flight envelope. MOSCA, which
uses flight-test frequency responses at multiple operating points, results in a non-linear
model that is guaranteed to be highly accurate within the boundary of the flight enve-
lope covered with all the experiments.

1 When the the moments of inertia were approximately measured, the R-50 had a different configuration

from the configuration during flight tests.


2.6. MOSCA-model evaluation 34

Table 2.1: Optimization results from MOSCA (British System of weight and measure)

Namea Measured Assumed Identified Description


se s lon 0.40 0.329 stab-bar cyclic stick to blade pitch
se s lat 0.45 0.414 stab-bar cyclic stick to blade pitch
Ibet 0.71 0.707 blade moment of inertia
Kbet 578.4 510.9 hinge spring
Kl 0.15 0.169 pitch-flap coupling
epsln 0.034 0.042 normalized hinge offset
xh 0.0 0.003 x hub position relative to the c.g.
yh 0.0 −0.025 y hub position relative to the c.g.
zh −1.887 −1.80 z hub position relative to the c.g.
Ixx 6.94 6.878 roll moment of inertia
Iyy 14.48 19.051 pitch moment of inertia
Ixz 0.0 −0.034 product of inertia
Ibet s 0.02 0.0197 stab-bar moment of inertia
as 1.6 1.28 lift curve slope of stab-bar
k vel tail 1.1 1.68 corrective factor for velocity at tail

a Name of the parameters in MOSCA computer code.

2.6.2 Flight-test comparisons

Figures 2.3 and 2.4 illustrate selected frequency-response comparisons at hover and
cruise flight. The results for the full- and reduced-order linear models from MOSCA
are practically the same and difficult to distinguish in the figures. The agreement is
excellent especially if compared to the state-of-the-art linear models of Mettler et al.
[2002] identified with CIFER. The coherence plots are useful to interpret the flight tests:
a value smaller than 0.6 or a deep notch indicates a decrease in accuracy of flight tests
data or a strong effect of non linearities on the response.
Figures 2.5 and 2.6 show hover and cruise-flight time-history comparisons for dou-
blet inputs in all axes. The MOSCA reduced linear model has not been included since,
2.7. Conclusions on integrated modeling 35

as for frequency responses, it matches the MOSCA linear model without noticeable dif-
ferences. These results show clearly that the accuracy of the models justify their use for
high-bandwidth control design.

2.7 Conclusions on integrated modeling

This chapter has presented the MOSCA integrated modeling technique. The high-
fidelity of MOSCA models is suitable for both flight simulation and control design.
MOSCA was applied to the CMU robotic helicopter starting from the development of
a medium-complexity non-linear mathematical model. The same work-flow could be
equally applied to other helicopters, fixed-wing aircrafts, or other aerial vehicles sub-
stituting only the non-linear model. A case would be, for example, the tuning of higher
complexity helicopter models such as GENHEL models [Ballin and Dalang-Secrétan,
1991], or the development of high-fidelity models for aerial robots of unconventional
design.
The key results are

• MOSCA integrates synergistically first-principles and system-identification mod-


eling techniques. It delivers a very accurate non-linear model suitable for flight
simulations, and linear models adequate for control design. The linear models
compare favorably with high-order linear models identified with CIFER.

• MOSCA is particularly appealing for unmanned aerial vehicles (UAVs) of new


design. Contrarily to full-scale aircraft, UAVs suffer from the lack of an extended
body of knowledge and experience in modeling. MOSCA could be of help in de-
veloping accurate models for control design and simulation quickly and reliably.

• The non-linear model of the CMU R-50 used to test MOSCA has 30 states. Model
reduction was applied to the extracted linear models without decreasing the ac-
curacy. The results show that small-scale helicopter dynamics can be predicted
fairly well with a model which includes at least a first-order representation of the
rotor and stabilizer-bar dynamics.
2.7. Conclusions on integrated modeling 36

p/lon q/lon p/lat


Mag (dB)

50 50 50

0 0 0
0 1 0 1 0 1
10 10 10 10 10 10
Phase (deg)

−200 0 0

−400 −100 −100


−200 −200
−600
−300 −300
0 1 0 1 0 1
10 10 10 10 10 10
1
Coherence

1 1

0.8
0.9
0.5
0 1 0 1 0 1
10 10 10 10 10 10
q/lat w/col r/ped
40
Mag (dB)

50 50
20

0 −20 0
0 1 0 1 0 1
10 10 10 10 10 10
−200
Phase (deg)

0 0
−250
−200 −100
−300 −200
−400
−350 −300
0 1 0 1 0 1
10 10 10 10 10 10
1 1 1
Coherence

0.8
0.5
0.6 0.5
0
0 1 0 1 0 1
10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s) Frequency (rad/s)

Figure 2.3: Hover comparisons: flight tests (solid), MOSCA linear model (dash-dot),
MOSCA reduced linear model (dashed), CIFER linear model [Mettler et al., 2002] (dot-
ted).
2.7. Conclusions on integrated modeling 37

p/lon q/lon p/lat


50 50 50
Mag (dB)

0 0 0
0 1 0 1 0 1
10 10 10 10 10 10
Phase (deg)

−200 0 0

−100
−400 −200
−200
−600 −300 −400
0 1 0 1 0 1
10 10 10 10 10 10
1 1 1
Coherence

0.5 0.9 0.5

0 0.8 0
0 1 0 1 0 1
10 10 10 10 10 10
q/lat w/col r/ped
40
Mag (dB)

50 50
20

0 −20 0
0 1 0 1 0 1
10 10 10 10 10 10
0 100
Phase (deg)

0
0
−200 −500
−100
−400
−1000 −200
0 1 0 1 0 1
10 10 10 10 10 10
1 1 1
Coherence

0.5 0.5 0.5

0 0 0
0 1 0 1 0 1
10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s) Frequency (rad/s)

Figure 2.4: Forward flight (15 m/s) comparisons: flight tests (solid), MOSCA linear
model (dash-dot), MOSCA reduced linear model (dashed), CIFER linear model [Mettler
et al., 2002] (dotted).
2.7. Conclusions on integrated modeling 38

lon lat col ped


100 100 100 100
%

0 0 50 0

−100 −100 0 −100


5 5 5 5
u (m/s)

0 0 0 0

−5 −5 −5 −5
5 6 5 5
v (m/s)

4
0 2 0 0
0
−2
−5 −4 −5 −5
5 5 5 5
w (m/s)

0 0 0 0

−5 −5 −5 −5
50 50 50 50
p (deg/s)

0 0 0 0

−50 −50 −50 −50


50 50 50 50
q (deg/s)

0 0 0 0

−50 −50 −50 −50


50 50 50 100
r (deg/s)

0 0 0 0

−50 −50 −50 −100


40 40 40 40
φ (deg)

20 20 20 20
0 0 0 0
−20 −20 −20 −20
−40 −40 −40 −40
40 40 40 40
θ (deg)

20 20 20 20
0 0 0 0
−20 −20 −20 −20
−40 −40 −40 −40
40 40 40 0
ψ (deg)

20 20 20 −20
0 0 0 −40
−20 −20 −20 −60
−40 −40 −40 −80
0 5 0 2 4 6 0 2 4 6 0 5
Time (s) Time (s) Time (s) Time (s)

Figure 2.5: Time comparisons at hover among flight test (solid), MOSCA linear model
(dash-dot), CIFER linear model [Mettler et al., 2002] (dotted).
2.7. Conclusions on integrated modeling 39

lon lat col ped


100 100 100 100
%

0 0 50 0

−100 −100 0 −100


20 20 20 20
u (m/s)

15 15 15 15

10 10 10 10
2 2 2 5
v (m/s)

0 0 0 0
−2 −2 −2
−4 −4 −4 −5
−6 −6 −6
w (m/s)

2 2 2 2
0 0 0 0
−2 −2 −2 −2
−4 −4 −4 −4
−6 −6 −6 −6
50 50 50 50
p (deg/s)

0 0 0 0

−50 −50 −50 −50


50 50 50 50
q (deg/s)

0 0 0 0

−50 −50 −50 −50


50 50 50 50
r (deg/s)

0 0 0 0
−50
−50 −50 −50
40 40 40 40
φ (deg)

20 20 20 20
0 0 0 0
−20 −20 −20 −20
−40 −40 −40 −40
40 40 40 40
θ (deg)

20 20 20 20
0 0 0 0
−20 −20 −20 −20
−40 −40 −40 −40
40 40 40 40
ψ (deg)

20 20 20 20
0 0 0 0
−20 −20 −20 −20
−40 −40 −40 −40
0 5 0 2 4 6 0 2 4 6 0 5
Time (s) Time (s) Time (s) Time (s)

Figure 2.6: Time comparisons in forward flight (15 m/s) among flight test (solid),
MOSCA linear model (dash-dot), CIFER linear model [Mettler et al., 2002] (dotted).
Chapter 3
Robust hover control

3.1 H∞ loop shaping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3.2 Inner-loop design and analysis . . . . . . . . . . . . . . . . . . . . . . . 47

3.3 Outer-loop design and analysis . . . . . . . . . . . . . . . . . . . . . . 58

3.4 Controller implementation . . . . . . . . . . . . . . . . . . . . . . . . . 68

3.5 Flight tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

3.6 Loading analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

3.7 Conclusions on robust hover control . . . . . . . . . . . . . . . . . . . 76

Robotic helicopters are dynamic systems with characteristics that make them an in-
teresting problem for control system design. They are multiple-input/multiple-output
(MIMO), non-linear, non-minimum phase, highly coupled, and unstable. These prob-
lems have motivated the investigation of robust multivariable control design techniques,
and particularly H∞ loop shaping on the CMU R-50.
There are only a few cases in which applications of linear multivariable robust meth-
ods for high-bandwidth controller design have been actually implemented and flight
tested. The principal reason is that high-bandwidth controller design requires high-
fidelity modeling [Tischler, 1990]. This chapter shows how the use of MOSCA models

40
3.1. H∞ loop shaping 41

and H∞ loop shaping produces robust high-bandwidth controllers for hover and low-
speed flight.

3.1 H∞ loop shaping

H∞ loop shaping is a sensible and appealing procedure for designing robust controllers [Pa-
pageorgiou and Glover, 1999]. It is a combination of loop shaping [Doyle et al., 1992]
and robust stabilization and is described below after some useful preliminaries.

3.1.1 Classical loop shaping

Loop shaping is an approach to controller design. Given a SISO plant G (see Figure
3.1), it consists of shaping the loop transfer function L = GK, by choosing an appropri-
ate K, to meet desired closed-loop objectives such as reference tracking or disturbance
rejection. The relation between closed-loop objectives and L is evident considering the

r K G y

Figure 3.1: Block diagram of a closed loop system.

expressions of the transfer functions for the system in Figure 3.1:

L 1 L
y= r+ d− n (3.1)
1+L 1+L 1+L
3.1. H∞ loop shaping 42

To have the reference r following the output y we desire to have | L| large. A large loop
gain reduces also the effect of the disturbance d on y but increases the transmission of
noise n to y. In fact, to reduce the effect of n onto y we need | L| to be small. Require-
ments on | L| are conflicting, but, generally, reference tracking and disturbance rejection
(| L| large) are required at low frequency while noise rejection (| L| small) is required at
high frequency. In the crossover region, it is necessary to limit the slope of | L|, since the
robustness of the closed-loop system and the transient nominal performance increase
with a decrease in the slope of | L| at crossover [Bode, 1945]. Typically, the slope of | L| is
set at 20dB/decade. The design requires some trade-off.

3.1.2 Singular values

For a linear SISO system y = Gu, the gain at frequency ω is

|y(ω )| |G ( jω )u(ω )|
= = |G ( jω )| (3.2)
|u(ω )| |u(ω )|

that is, the gain is independent of the magnitude of the input |u(ω )|.
For a MIMO system y = Gu, where in general the input u and output y are vectors,
the gain at frequency ω can be defined, using the Euclidean 2-norm ||.||2 :

||y (ω )||2 ||G ( jω )u (ω )||2


= (3.3)
||u (ω )||2 ||u (ω )||2

The gain can be shown to be independent of the magnitude ||u (ω )||2 but depends on
the direction of u. Singular value decomposition (SVD) of G helps in determining the
maximum and the minimum values of the gain in Equation 3.3 as the direction of u is
varied. In fact, if σ̄ ( G ) and σ ( G) are the maximum and minimum singular values of G
then

||G ( jω )u(ω )||2 ||G ( jω )u (ω )||2


max = σ̄[G ( jω )], min = σ[G ( jω )] (3.4)
u ( ω )6 =0 ||u (ω )||2 u( ω )6 =0 ||u (ω )||2

These relations are valid except when G has more inputs than outputs, in which case
the minimum gain is zero.
3.1. H∞ loop shaping 43

Singular values allow the generalization of the practice of shaping transfer functions
of SISO systems for controller design to MIMO cases. In a MIMO problem, shaping
transfer functions means shaping their singular values. In this context, a very useful
metric is the ∞−norm of G, which is defined as

||G ||∞ = sup σ̄[G ( jω )] (3.5)


ω

||G ||∞ is then the maximum gain of G over all frequencies and all the input directions.

3.1.3 Coprime factor uncertainty

One of the main goals of feedback control is to preserve stability in the presence of
disturbances and plant uncertainties (robust stabilization). A general and powerful
description for uncertainties in a plant is the coprime factor uncertainty. A left coprime
factorization of a plant G(s) is
−1
G(s) = M̃ (s) Ñ (s) (3.6)

where M̃ (s) and Ñ (s) are stable coprime transfer functions.


M̃ (s) and Ñ (s) are left coprime if they satisfy the Bezout identity:

M̃X l + ÑY l = I (3.7)

where X l and Y l are stable transfer functions.


M̃ (s) and Ñ (s) are normalized left coprime if they satisfy, in addition to the Bezout
identity, the following normalization condition:

M̃ M̃ ∗ + Ñ Ñ ∗ = I, ∀ s ∈ R (3.8)
T T
where M̃ ∗ (s) = M̃ (−s) and Ñ ∗ (s) = Ñ (−s). Similar relations hold for a right co-
prime factorization of G (s) (see Zhou et al. [1996]).
Coprime factor uncertainty is obtained by introducing perturbations to a left (or
right) coprime factorization of G:

Gp = ( M̃ + ∆˜ M )−1 ( Ñ + ∆˜ N ) (3.9)
3.1. H∞ loop shaping 44

+ −
˜N
∆ ˜M

+ −1
u
Ñ + M̃ y

K∞

Figure 3.2: The coprime factor uncertainty.

where G p is the perturbed plant (see Figure 3.2).

3.1.4 Design procedure

The H∞ loop-shaping design procedure [McFarlane and Glover, 1992] is composed of


the following steps.

1. Let G denote a linear time-invariant model of the plant for which the controller
is to be designed. Shape the singular values of G with frequency-dependent
weights W 1 and W 2 according to closed-loop objectives. The weighted plant
Gs = W 2 GW 1 is depicted in Figure 3.3. What follows is a typical weighting

z2 w2 w1 z1

Gs = W 2 GW 1

K∞

Figure 3.3: The H∞ loop-shaping standard block diagram.

scheme. W 2 contains low-pass filters for noise rejection and lead-lag filters for
3.1. H∞ loop shaping 45

improving robustness. W 1 contains proportional plus integral (PI) filters. The


integrators are used to boost the low frequency gain, and thus improve output
decoupling, tracking, and disturbance rejection at both the plant input and out-
put. The proportional matrix gain is used to reduce the phase lag introduced by
the integrators around crossover (i.e., to reduce the slope of the singular values),
and to set the amount of actuator usage. The overall gains of W 1 and W 2 are used
to specify the desired loop bandwidth.

2. Maximize the inverse of the H∞ -norm of the transfer function matrix from dis-
turbances [w1 , w2 ]T to errors [z1 , z2 ]T over all stabilizing controllers K ∞ ; that is,
    −1   −1
w1 z1 I
h i
−1

max   →   = max   ( I − Gs K ∞ ) I Gs =ǫ
stab K ∞ w z2
stab K ∞ K
2 ∞
∞ ∞
(3.10)
The stability margin ǫ takes values in the interval [0, 1] and is a measure of robust-
ness and performance. A margin greater than 0.3 is considered good, based on
theory and practical experience. The stability margin can be seen also as an upper
bound to “tolerable” coprime factor perturbations of G s . In fact, given a perturbed
left coprime factorization G p of G s as in Equation 3.9, stability is guaranteed for

|| [∆˜ N ∆˜ M ] ||∞ < ǫ (3.11)

The theoretical basis for H∞ loop shaping is that K ∞ does not modify the desired
loop shape (G s ) significantly at low and high frequencies if the stability margin
is not too small [McFarlane and Glover, 1992]. Consider for example Figure 3.14.
The synthesized controller K ∞ achieves an ǫ = 0.36, is essentially a lead-lag filter
and indeed does not modify Gs significantly. Therefore, the designer can define
performance objectives by shaping the open-loop model G with weights W 1 and
W 2.

Recently, well established notions of robustness for single-input/single-output


systems have been related to the stability margin. It was proved in [Glover et al.,
3.1. H∞ loop shaping 46

2000] that the Nyquist or Nichols plot of any loop transfer function that results
from breaking the closed loop at one of the actuators or sensors lies outside a
region whose size is a function of the stability margin. (see Figure 3.4). The guar-

Guaranteed Nichols plot exclusion regions


8

2
Gain (dB)

ε = 0.3
0 ε = 0.35
ε = 0.4
−2

−4

−6

−8
−50 −40 −30 −20 −10 0 10 20 30 40 50
Phase (deg)

Figure 3.4: Guaranteed Nichols plot exclusion regions for ǫ = 0.3, 0.35, 0.4.

anteed single-loop gain and phase margins, denoted by GM and PM respectively,


are related to ǫ by:

1+ǫ
GM = ±20 log10 , and PM = 2 arcsin ǫ (3.12)
1−ǫ

Thus, a stability margin of 0.35 guarantees gain and phase margins of ±6.3dB
and 40.9 degrees respectively. In [Glover et al., 2000] the stability margin is also
related to a multi-input/multi-output robustness test popular in the aeronautical
industry.

3. Choose the position of the controller. There are three ways to implement the
H∞ loop-shaping controller: in the forward path, in the backward path, and in
the observer form [Sefton and Glover, 1990].
3.2. Inner-loop design and analysis 47

4. Check time simulations and frequency responses of the resulting closed-loop sys-
tem to verify robust performance. Iterations may be required.

The controller structure implemented and flight tested on the CMU R-50 for hover
and low-speed control is based on an inner-loop and outer-loops architecture: the inner
loop (Figure 3.5) provides stabilization and decoupling; the outer loops (Figure 3.6 and
3.7) are cascaded with the inner loop for velocity and position control. H∞ loop shaping
is used to design all the loops as illustrated in the next sections.
x
y
z
θr lon ψ
φr lat vx
vz K ∞ ( 0 )W 2 ( 0 ) W1 col R−50 vy
ψ̇r ped θ
φ
vz
k
s I ψ̇

K∞ W2

Figure 3.5: Inner loop.

3.2 Inner-loop design and analysis

The inner loop (Figure 3.5) is an attitude-command attitude-hold (ACAH) response


type controller. Reference inputs are pitch attitude (θ), roll attitude (φ), vertical velocity
(vz ), and yaw rate (ψ̇). The same four variables are used for feedback. The measurement
set does not include pitch rate (q) and roll rate (p) because the stabilizer bar acts already
3.2. Inner-loop design and analysis 48

φr
k θr
s

v xr x
θ
Kvx (0)W2vx (0) W1vx
y
v yr φ R−50
Kvy (0)W2vy (0) W1vy z
+
v zr Inner ψ
k Loop vx
s ψ̇r
vy

Kvy W2vy

Kv x W2vx

Figure 3.6: Outer loops for vx and vy control.

as a lagged-rate feedback system [Miller, 1950].

Given the absence of handling-quality requirements for small-scale helicopters, the


controller design process starts by pushing the bandwidth as high as possible, while
maintaining a good level of robustness. The compilation of a handling-quality standard
for robotic helicopters is still an open argument: there are many questions about the
kind of missions they will be asked to accomplish, and there is no human pilot on
board requesting and evaluating a set of specific dynamic characteristics. For flight
control problems with mostly qualitative closed-loop specifications (i.e., very few or
none of the specifications are quantified as would probably be the case for an unmanned
air vehicle), H∞ loop-shaping control can be used to establish quickly the achievable
performance. This is because H∞ control can systematically handle the multivariable
and uncertain nature of the aircraft.

In the H∞ loop-shaping framework it is easy to set the system bandwidth and thus,
the achievable performance would be determined by increasing the system bandwidth
3.2. Inner-loop design and analysis 49

y
xr vx R−50 x
z +
yr vy Inner y
Loop
zr vz + z
v x and vy
ψr ψ̇ Outer ψ
Kψ (0)W2ψ (0) W1ψ Loops

ψ̇r

Kψ W2ψ

Figure 3.7: Outer loops for position control.

until robustness as quantified by the stability margin (see Equation 3.11) decreases be-
low the desired level. Moreover, the use of a high-fidelity model in controller design
provides confidence in the achieved robustness, and makes it possible to represent ac-
curately the high-bandwidth and multivariable effects that H∞ control exploits.

To help in assessing the performance of the hover controller, the analysis uses some
specifications adopted in the full-scale military-helicopter community [ADS-33E, 2000,
MIL-F-9490D, 1975]. Although a higher level of performance (in particular agility and
maneuverability) is generally expected from small-scale helicopters, the CMU R-50 is
quite limited in achievable performance; given the original design aim of the vehicle
(crop dusting at limited velocities), together with the relevant payload added by the
CMU instrumentation (50% of the machine’s dry weight), and the low bandwidth of
the actuators (15 rad/s), the use of a set of full-scale helicopters’ specifications may be
justified. Nevertheless, the design tries to satisfy the requirements with a large margin
on the examined bounds.
3.2. Inner-loop design and analysis 50

The requirements for the inner-loop controller are as follows.1

Stability. In single-loop analysis, gain margin shall be ≥ 6 dB and phase margin ≥ 45


degrees to guarantee a reasonable level of robustness.

Bandwidth. Closed-loop bandwidth ωBW shall satisfy Level 1 Target Acquisition and
Tracking Specifications.2 For the θ, φ, and ψ channel, ADS-33E defines ωBW as
the frequency where the phase of the closed-loop response is Φ = −135 deg.
ωBW should be larger then 2.0, 2.5, and 3.5 rad/s for the θ, φ, and ψ channels
respectively.

For the vz channel, ADS-33E gives the desired speed of response in terms of
bounds on a time constant RTvz ≤ 5.0 and a time delay τvz ≤ 0.2 of a first-order
transfer function
eτvz s
TFIT =
RTvz s + 1
that is used to fit the step response of the vz channel. ADS-33E considers the fit
successful when the following cost coefficient is between 0.97 and 1.03:

||vzFIT − v̄z ||2


r2 = ,
||vz − v̄z ||2
where v̄z is the average of vz during the step.

φ θ ψ̇
Decoupling. Off-axis responses due to on-axis requests ( θr , φr , vzr ) shall comply with
Level 1 Aggressive agility specifications.

For the first two transfer functions, ADS-33E imposes the following bounds:
∆θpk ∆φpk
−0.25 ≤ ≤ 0.25, −0.25 ≤ ≤ 0.25,
∆φ4 ∆θ4

where the numerators are the differences between the values (peak within 4 s and
trim) of the off-axis variable caused by a step change of the on-axis variable, and
1 Details about specifications’ metrics are in ADS-33E [2000], and MIL-F-9490D [1975].
2 Level 1 and Target Acquisition and Tracking are the highest levels achievable for military rotorcraft in
the ADS-33E specifications.
3.2. Inner-loop design and analysis 51

the denominators are the differences between the values (at 4 s and at trim) of the
on-axis variable.

For the third transfer function3 , ADS-33E imposes bounds on ψ̇ following a step
request in vz :
ψ̇1 ψ̇3
vz (3) ≤ 0.65, −0.15 ≤ ≤ 0.2,

|vz (3)|
where ψ̇1 is the first peak of ψ̇ before 3 seconds or the value of ψ̇ at 1 second ψ̇ (1)
if no peak occurs, and

 ψ̇(3) − ψ̇1 : ψ̇1 > 0
ψ̇3 =
 −ψ̇(3) + ψ̇ : ψ̇1 < 0
1

The units for the correct estimation of the bounds are deg/s for ψ̇, and ft/s for vz .

Disturbance Rejection. Response of θ, φ, vz , and ψ̇ to pulse disturbances injected di-


rectly at the inputs of the plant shall return to within 10% of the peak value in less
than 10 seconds.

Some of these requirements are more meaningful with a pilot in the loop; they are con-
sidered here, since there are applications for unmanned helicopters, where the pilot will
still perform primarily the guidance and navigation functions. Examples are filmmak-
ing or some full-scale unmanned helicopters [Frost et al., 2000].
The diagonal weight W 1 has integral action on all the axes; W 1 contains also zeros
in the θ, φ, and vz axes to decrease roll-off rates at crossover. The complete expression
for W 1 is
s + zi1
 
ki1 s
 
 s + zi2 
 ki2 s 
W1 = 
 s + zi3

 (3.13)
 ki3 s

 
1
ki4 s

3 ADS-33E takes in consideration the transfer function r/vzr , where r is the yaw rate in the body frame.
At hover the difference between r and ψ̇ is minimal.
3.2. Inner-loop design and analysis 52

The diagonal W 2 contains second-order low-pass filters for sensor noise rejection on vz
and ψ̇. The complete expression for W 2 is
 
ko1
 
 
 ko2 
W2 =   (3.14)
 ωn2 
ko3
s +2ζωn + ωn2
 
 
ωn2
ko4
s +2ζωn + ωn2

The gains ki1−4 of W 1 and ko1−4 of W 2 are tuned to have crossover frequencies of the
open-loop system at 7 rad/s. The tuning follows the guidelines of Hyde [1995], and it
is done automatically. The procedure consists of ordering the inputs and outputs of the
plant such that the plant is as diagonally dominant as possible; the model with inputs
u and outputs y
   
δlon θ
   
 δlat  φ
   
u=
 ,
 y=
 ,
 (3.15)
 δcol   vz 
   
δped ψ̇

already satisfies this condition. Then, calculating the values at the desired crossover
frequencies for the singular values from each input to all outputs, and the singular
values from all inputs to each output, it is possible to tune the gains ki1−4 and ko1−4 .
Figures 3.8 and 3.9 show the singular values of the plant after the tuning.
Figure 3.10 and 3.11 show the singular values of W 1 and W 2 respectively. Figure 3.12
shows the singular values of the plant with and without the shaping. The H∞ optim-
ization gives ǫ = 0.362. Figure 3.13 shows the singular values of the controller and
Figure 3.14 shows the singular values of the shaped hover model with the H∞ loop-
shaping controller. The singular values of the closed-loop system

T = ( I − GW 1 K ∞ W 2 )−1 GW 1 K ∞ (0)W 2 (0)

are in Figure 3.15. K ∞ (0) and W 2 (0) represent K ∞ and W 2 evaluated at zero frequency,
3.2. Inner-loop design and analysis 53

δ to all outputs δ to all outputs


lon lat
100 100

50 50
Magnitude (dB)

Magnitude (dB)
0 0

−50 −50

−100 −2 −1 0 1 2
−100 −2 −1 0 1 2
10 10 10 10 10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)

δcol to all outputs δ to all outputs


ped
80 100

60

50
Magnitude (dB)

Magnitude (dB)

40

20
0
0

−20
−50

−40

−60 −2 −1 0 1 2
−100 −2 −1 0 1 2
10 10 10 10 10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)

Figure 3.8: Singular values of the shaped plant from each input to all outputs.

and form the constant prefilter needed to have steady-state gain from references to out-
puts equal to I.
The inner-loop controller contains also an anti-windup loop (Figure 3.5) based on a
classical scheme [Kothare et al., 1994]: a diagonal system with integrators and constant
gains feeds back, to the output of W 1 , the difference between the actual input to the
plant and the output from W 1 . There are two reasons for the presence of an anti-windup
loop. First, W 1 contains integrators that in case of actuators’ saturation continue to
integrate the error signals causing windup problems (overshoot and rapid degradation
of performance). Second, the anti-windup loop ensures a smooth transition (bumpless
3.2. Inner-loop design and analysis 54

All inputs to θ All inputs to φ


50 60

40
Magnitude (dB)

Magnitude (dB)
20
0
0

−20
−50
−40

−60

−100 −2 −1 0 1 2
−80 −2 −1 0 1 2
10 10 10 10 10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)

All inputs to v All inputs to dψ/dt


z
80 80

60 60
Magnitude (dB)

Magnitude (dB)
40 40

20 20

0 0

−20 −20

−40 −40

−60 −2 −1 0 1 2
−60 −2 −1 0 1 2
10 10 10 10 10 10 10 10 10 10
Frequency (rad/s) Frequency (rad/s)

Figure 3.9: Singular values of the shaped plant from all inputs to each output.

transfer) between human-pilot mode and computer mode conditioning the controller
outputs while the controller is off line.
Figure 3.16 shows Bode plots with gain and phase margins for single-loop analysis;
all the loops satisfy the robustness specifications.
Table 3.1 compares bandwidth ωBW to the ADS-33E Level 1 specification for Target
Acquisition and Tracking introduced on page 49. The ADS-33E does not specify ωBW
for the ψ̇ channel, but the calculated bandwidth may be considered satisfactory. For the
vz channel, Table 3.1(b) reports the evaluation of response speed as specified on page 50.
Figure 3.17 shows two non-linear step responses of the closed loop; both on-axis
3.2. Inner-loop design and analysis 55

σ(W1)
60

40
Magnitude (dB)

20

−20

−40 −2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

Figure 3.10: Singular values of W 1 .

σ(W2)
5

0
Magnitude (dB)

−5

−10

−15 −2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

Figure 3.11: Singular values of W 2 .


3.2. Inner-loop design and analysis 56

σ(G) (dashed), σ(W2 G W1) (solid)


80

60

Magnitude (dB) 40

20

−20

−40

−60

−80

−2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

Figure 3.12: Singular values of G and W 2 GW 1 .

and off-axis responses needed for the evaluation of decoupling performance (defined
on page 50) are present. Table 3.2 shows the calculated values and bounds. The de-
ψ̇
coupling specification for the v zr transfer function requests the presence of a heading
(ψ) hold controller; since the ψ controller is on the outer loop, this specification is not
tested.

The last specification is disturbance rejection. Figure 3.18 shows the non-linear nor-
malized (respect to the peak values) responses for each of the 4 axes when sharp square
pulses are applied directly at correspondent plant’s input; they all clear the 10% bound
well before the 10 seconds required. The θ and the φ responses in Figure 3.18 show os-
cillatory responses. The oscillatory nature of the attitude responses of the R-50 is due to
the lightly-damped coupling mode between the rotor system and the fuselage evident
in the frequency responses of the angular rates to cyclic control inputs (Figure 2.3); the
stabilizer bar is responsible for the low damping of this mode [Heffley, 1979].

Similar behavior is present in the closed-loop responses to step references (see Fig-
3.2. Inner-loop design and analysis 57

σ(K )

10

Magnitude (dB) 5

−5

−10

−15

−20 −2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

Figure 3.13: Singular values of the H∞ controller.

σ(W G W ) (dashed), σ(W K W G) (solid)


2 1 1 ∞ 2
100

50
Magnitude (dB)

−50

−100

−150 −2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

Figure 3.14: Singular values of the shaped plant with and without the H∞ controller.
3.2. Inner-loop design and analysis 58

σ(T)

Magnitude (dB) −20

−40

−60

−80

−100
−2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

Figure 3.15: Singular values of the closed-loop system.

ure 3.17). In [Mettler et al., 2000], the authors proposed the use of notch filters at the
input of the plant to compensate the effects of the lightly-damped modes (i.e., to allevi-
ate the reduction in gain margin caused by the resonances, as evident in Figure 3.16(b),
and to avoid exciting the lightly-damped modes with fast attitude reference changes).
This choice, besides reducing the robustness of the system with the addition of
phase delay at crossover, does not improve the disturbance rejection problem when
the notch filters are added to a high-bandwidth controller; in this case, the effect of dis-
turbances injected at the actuators’ input could be in fact, even worse. The disturbances
may still excite the lightly-damped mode (they enter the plant without being filtered
by the notch filters), and any counteracting effect of the high-bandwidth feedback con-
troller at the frequency of the under-damped mode is largely reduced by the presence
of the notch filters in front of the actuators. If the light oscillatory responses to step
reference inputs are of concern, the notch filters may be safely introduced as prefilters
at the input of the closed-inner loop. A more drastic but better solution would consist
of removing the stabilizer bar [Mettler et al., 2000].
3.2. Inner-loop design and analysis 59

GM = 8.12 dB (at 8.47 rad/s) GM = 7.61 dB (at 13.34 rad/s)


40

Magnitude (dB)
Magnitude (dB)

40
20
20
0
0
−20
−20
−40
0 1 0 1
10 10 10 10

PM = 45.55 deg (at 2.95 rad/s) PM = 46.81 deg (at 3.85 rad/s)
0

Phase (deg)
Phase (deg)

−100 −100

−200 −200

−300 −300

−400 0 1 0 1
10 10 10 10
Frequency (rad/s) Frequency (rad/s)
(a) θ channel. (b) φ channel.

GM = 11.06 dB (at 15.78 rad/s) GM = 9.59 dB (at 8 rad/s)


40
20
Magnitude (dB)
Magnitude (dB)

20
0

0
−20

−20
−40
0 1 0 1
10 10 10 10
PM = 54.32 deg (at 4.53 rad/s) PM = 68.22 deg (at 2.23 rad/s)
−100
−100
Phase (deg)

−150
Phase (deg)

−200

−200 −300

−250 −400
0 1 0 1
10 10 10 10
Frequency (rad/s) Frequency (rad/s)
(c) vz channel. (d) ψ̇ channel.

Figure 3.16: Bode plots for broken single loop with gain and phase margins.
3.3. Outer-loop design and analysis 60

Table 3.1: Bandwidth evaluations with ADS-33E.


(a) θ, φ, and ψ̇ channels.

Channel ωBW (rad/s) ωBW (ADS-33E)

θ 5.6 ≥ 2.5
φ 9.08 ≥ 2.5
ψ̇ 6.4 -

(b) vz channel.

Channel RTvz (s) RTvz (ADS-33E) τvz (s) τvz (ADS-33E)

vz 0.22 ≤ 5.0 0.04 ≤ 0.2

Table 3.2: Inter-axis coupling evaluations with ADS-33E.

∆θpk ∆θpk ∆φpk ∆φpk


∆φ4 ∆φ4 (ADS-33E) ∆θ4 ∆θ4 (ADS-33E)

0.08 ∈ ±0.25 -0.08 ∈ ±0.25

3.3 Outer-loop design and analysis

The inner-loop closure makes available a high-bandwidth, robust, and decoupled sys-
tem with four inputs (θr , φr , vzr , ψ̇r ) and four outputs (θ, φ, vz , ψ̇). The decoupling jus-
tifies the use of cascaded single-input/single-output (SISO) outer loops to allow the
helicopter to fly a specified trajectory. Before proceeding with the actual design of
the outer-loop controller, the closed inner-loop system is approximated using balanced
model truncation [Zhou et al., 1996] to reduce the order of the outer loops.
The outer-loop structure uses and controls variables in a reference frame known as
heading-referenced inertial frame. This frame has its origin fixed at the helicopter’s
center of gravity and when ψ = 0 the x, y, and z axes are parallel to the earth frame; if
3.3. Outer-loop design and analysis 61

θ (solid), φ (dash−dot)
5

Attitude (deg)
−5

−10

−15

−20

−25
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s)
(a) θr step.
φ (solid), θ (dash−dot)
5

0
Attitude (deg)

−5

−10

−15

−20

−25
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s)
(b) φr step.

Figure 3.17: Non-linear step responses for inter-axis coupling evaluation.

θ = a1 , φ = a2 , and ψ = a3 , where a1 , a2 , and a3 are three given angles, the heading-


referenced inertial frame rotates by a3 around the z axis of the earth frame but maintains
its z axis parallel to the z axis of the earth frame (i.e., it is unaffected by θ and φ angular
motions of the helicopter). In this frame, vx and vy are sometimes referred to as along-
heading and cross-heading velocities. In the remainder of this thesis, the control-system
variables for velocities and positions vx , vy , vz , x, y, and z are all assumed to be in the
3.3. Outer-loop design and analysis 62

1 θ
φ
v
z
dψ/dt
0.5

Amplitude
0

−0.5

−1
0 2 4 6 8 10
Time (s)

Figure 3.18: Non-linear normalized responses to square pulse disturbances.

heading-referenced inertial frame.


The first two outer loops (Figure 3.6) are closed on the θ and φ channels of the inner
loop. They allow control of vx and vy . For the vx and vy channels, ADS-33E specifies
only the rise time of their closed-loop response. The other general requirement that is
added is the achievement of a good level of robustness measured by gain and phase
margins as in MIL-F-9490D.
The crossover frequencies are set at 0.8 rad/s for both loops. The W1 weights for
each loop have a proportional-plus-integral structure; the W2 weights contain the same
second-order low-pass filters of the inner loop for noise rejection. Figure 3.19 shows
vx vy
the singular value of the transfer functions θr and φr before and after the shaping. The
H∞ optimization gives ǫ = 0.374 for the vx controller and ǫ = 0.398 for the vy controller.
Figure 3.20 shows the singular values of the two controllers and Figure 3.21 shows the
singular values of the closed-loop systems

Tv x = (1 − Gv x /θ W1v x Kv x W2v x )−1 Gv x /θ W1v x Kv x (0)W2v x (0),

Tvy = (1 − Gvy /φ W1vy Kvy W2vy )−1 Gvy /φ W1vy Kvy (0)W2vy (0).
3.3. Outer-loop design and analysis 63

v /θ : σ(G) (dashed), σ(W G W ) (solid)


x r 2 1

50

Magnitude (dB)
0

−50

−100

−2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

v /φ : σ(G) (dashed), σ(W G W ) (solid)


y r 2 1

50
Magnitude (dB)

−50

−100

−2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

vx vy
Figure 3.19: Singular values of θr and φr with and without shaping weights.

Notice that the stability margins obtained with these controllers refer to the open-loop
systems obtained by breaking the vx and vy loops at the θ and φ reference inputs of the
closed inner loop (see Figure 3.6). Table 3.3 reports, instead, gain and phase margins
obtained by breaking the loops at the plant inputs when the vx and vy loops are closed.
3.3. Outer-loop design and analysis 64

σ(K): vx channel (solid), vy channel (dashed)


5

Magnitude (dB)
−5

−10

−15

−20 −2 −1 0 1 2
10 10 10 10 10
Frequency (rad/s)

Figure 3.20: Singular values of Kv x and Kvy .

σ(T): vx channel (solid), vy channel (dashed)


10

0
Magnitude (dB)

−10

−20

−30

−40

−50

−60 −2 −1 0 1
10 10 10 10
Frequency (rad/s)

Figure 3.21: Singular values of Tv x and Tvy .

The phase margins for the loops broken at the longitudinal and lateral cyclic inputs
of the plant do not satisfy the requested phase margins bounds (PM ≥ 45 degrees). A
3.3. Outer-loop design and analysis 65

Table 3.3: Gain and phase margins with broken-loop analysis for velocity outer loops.

Loop Gain Margin (dB) Phase Margin (deg)

vx 8.23 @ 8.39 rad/s 38.07 @ 2.74 rad/s


vy 7.64 @ 13.2 rad/s 41.69 @ 3.63 rad/s
vz 11.22 @ 16.3 rad/s 55.11 @ 4.53 rad/s
ψ̇ 9.64 @ 8.04 rad/s 68.20 @ 2.22 rad/s

possible remedy would be to increase the robustness of the inner-loop design modifying
W 1 and W 2 (for example adding lead compensation to decrease the phase delay), or to
reduce the bandwidth of the vx and vy loops reducing the gains in the respective W1 s.
Classical anti-windup loops are also present to counteract wind up of the integrators
in the W1 weights caused by the limiters on the requested attitude angles. For the vx and
vy channels, ADS-33E gives bounds on their rise time, which is defined as the time RT
in seconds when the step response reaches 63.2% of the steady-state value. The bounds
are

2.5 ≤ RTv x ≤ 5.0, 2.5 ≤ RTvy ≤ 5.0.

Figure 3.22(a) and 3.22(b) show step responses for the two loops. The calculated rise
times are RTv x = 2.64 s and RTvy = 2.5 s, which satisfy the specifications. Figure 3.22(c)
shows the step response for vz used in the inner-loop section for RTvz and τvz evalua-
tion; during this test, the vx and vy loops are closed, but simulations confirm that the
evaluation of RTvz and τvz is practically unchanged when leaving the vx and vy loops
open. With the closure of the vx and vy outer loops, the main inputs to the systems
become vxr , vyr , vzr , and ψ̇r (see Figure 3.6); these four channels are utilized by the next
four independent SISO outer loops to control x, y, z, and ψ. The controller architec-
ture also provides the capability to inject θr and φr for feed-forward control and turn
coordination.
3.3. Outer-loop design and analysis 66

Flight Test (solid), Simulation (dashed) Flight Test (solid), Simulation (dashed)
6 6

5 5

4 4
v (m/s)

v (m/s)
3 3

2 2
x

y
1 1

0 0

−1 −1
0 2 4 6 8 10 0 2 4 6 8 10
Time (s) Time (s)
(a) v x step. (b) vy step.
Flight Test (solid), Simulation (dashed) Flight Test (solid), Simulation (dashed)
0.5 25

0 20

−0.5 15
vz (m/s)

x (m)

−1 10

−1.5 5

−2 0

−2.5 −5
0 1 2 3 4 5 6 0 5 10 15
Time (s) Time (s)
(c) vz step. (d) 20 m x step.
Flight Test (solid), Simulation (dashed) Flight Test (solid), Simulation (dashed)
25 100

20 80

15
60
ψ (deg)
y (m)

10
40

5
20

0
0
−5
0 5 10 15 0 2 4 6 8 10
Time (s) Time (s)
(e) 20 m y step. (f) 90-degree ψ step.

Figure 3.22: Flight tests and MOSCA linear model vx , vy , vz , x, y and ψ step responses.
3.3. Outer-loop design and analysis 67

The design of the four position outer loops is analogous to the design of the vx and
vy loops; x, y, z, and ψ loops are closed respectively on vx , vy , vz , and ψ̇ (Figure 3.7).
Again, before closing the four outer loops, model reduction is performed on all the
x (s)
transfer functions (e.g., v xr (s)
). The crossover frequency is set at 0.6 rad/s for the x and
y loops, and at 0.8 rad/s for the z and ψ loops. None of the W1 s contain integral action
since each loop has already a free integrator4 . To decrease the roll-off rate at crossover
the W1 weights for the x and y loops contain lead compensation. The W2 weights are
simply unity gains. The H∞ optimization gave the following values of ǫ: ǫ = 0.59 for
the x loop, ǫ = 0.56 for the y loop, ǫ = 0.52 for the z loop, and ǫ = 0.54 for the ψ loop.

Figures 3.22(d-f) show step responses for the x, y, and ψ loops. For the ψ channel
the ADS-33E requires ωBW ≥ 3.5 rad/s. With the synthesized controller structure (ψ
controlled in the outer loop) it is nearly impossible to satisfy this specification with-
out a decrease in robustness of the design. The achieved ωBW is 2.5 rad/s. Since the
Yamaha R-50 is already equipped with an elementary yaw rate feedback system, a sen-
sible choice would be to eliminate the inner-loop control on ψ̇ and control directly ψ
instead. This would allow to have ωBW for ψ to be as high as the one obtained for ψ̇
(see Table 3.1) without decreasing the stability margins. The wide-envelope controller
presented in the next chapter is designed based on this observation.

The ψ response in Figure 3.23 is used to check the compliance with the specifications
in ADS-33E for a position-hold controller; ADS-33E requires that, for Aggressive agility,
a 360-degree yaw turn shall be completed in less than 10 s, in steady wind of up to 18
m/s, and maintaining the helicopter position within a 3 m diameter circle at constant
altitude. Since the wind during flight tests was not at the level prescribed, a MOSCA
non-linear simulation is used. The controller cleared the position and time bounds with
a frontal steady wind of 10 m/s. This wind speed is close to an upper limit for the R-50;
at larger wind speeds (≥ 15 m/s), the R-50 lacks tail rotor power to execute the turn (i.e.,
the helicopter stops turning at ψ = 90 degrees). The controller architecture leaves the
option of directly injecting vxr , vyr , vzr , and ψ̇r for feed-forward control (see Figure 3.7);

4 Position is measured in the earth frame and then transformed in the heading-referenced inertial frame
3.4. Controller implementation 68

6
ψ

Position (m), ψ (rad)


5 x
y
4
z
3

−1
0 5 10 15 20 25
Time (s)

Figure 3.23: 360-degree turn with 10 m/s steady wind.

it is also possible to switch off any position outer loops in case they are not needed to
fly a particular maneuver (e.g., in straight forward flight the x loop is switched off since
the request is on vx ).

3.4 Controller implementation

The MATLAB µ-Analysis and Synthesis Toolbox [Balas et al., 1995] is used for controller
design and reduction. After the reduction, the controller is discretized with Tustin bi-
linear transformation; the sampling frequency is 100 Hz. Using a ground laptop it is
possible to update gains, weights, filters, and H∞ controllers, in flight without landing
the helicopter. After any modification, an ad hoc utility recomputes the new state-
space matrices (which can also have a different number of states if the structures of
the weights have changed) for all the loops and sends them to the on-board computer
through a wireless ethernet connection. Then, the operator or the safety pilot switches
in the new controller. More drastic modifications are also possible. In fact, the high-

such that x, y, z are the integrals of v x , vy , vz ; yaw angle ψ is by definition the integral of yaw rate ψ̇.
3.5. Flight tests 69

fidelity non-linear simulation software MOSCA is interfaced with MATLAB. This ar-
chitecture allows the user to (1) modify the helicopter configuration on the field, (2)
update the model, (3) recompute the controller, (4) test the controller with non-linear
simulations, and (5) fly test the new system. Possible reconfigurations of the vehicle
include, for example, different payloads or weight redistribution.
This aspect is of primary importance, since a general criticism of multivariable tech-
niques has always been the difficulty in modifying and tuning the elements of the con-
troller. Once the controller architecture is fixed, H∞ loop shaping with automated code
generation and high-fidelity simulation models allows the designer to utilize classical
loop-shaping experience for quick and reliable tuning of the original design. This tech-
nique was tested during the flight tests shown in the next section. The original outer-
loop crossover frequencies were too high and caused undesirable fast attitude changes
in presence of wind gusts. After switching in the baseline controller, the gains of the
outer-loops W1 s were reduced and the utility to rebuild the whole controller executed.
The new controller was flight tested again without having to land the helicopter.

3.5 Flight tests

The goals of the flight tests were (1) to check the overall performance and robustness of
the system, and (2) to gain confidence in flying a set of maneuvers limiting the tracking
error within acceptable bounds. The maneuvers flown were the following:

• Forward Coordinated Turn: The helicopter starts at hover. The operator commands
a 6 m/s step in vx . The helicopter flies on a straight line until the operator com-
mands a turn. All the loops are engaged except the x position loop since the re-
quest is on vx . The turn trajectory is a circle of 10 m radius (R). Once the operator
issues the turn command, the helicopter flies the turn while maintaining vx = 6
vx
m/s, vy = 0, constant altitude, and ψ̇ = R. The y loop is engaged to drive the
horizontal tracking error to zero.
3.5. Flight tests 70

• Backward Coordinated Turn: Similar to the forward coordinated turn with vx =


−5 m/s. The helicopter starts at hover. The operator commands a -5 m/s step
in vx . The helicopter flies on a straight line until the operator commands a turn.
All the loops are engaged except the x position loop since the request is on vx .
The turn trajectory is a circle of 10 m radius (R). Once the operator issues the turn
command, the helicopter flies the turn while maintaining vx = −5 m/s, vy = 0,
vx
constant altitude, and ψ̇ = R. The y loop is engaged to drive the horizontal track-
ing error to zero.

• Nose-out Pirouette: The helicopter starts at hover. The operator commands a -4


m/s step in vy . The helicopter flies sideways on a straight line until the operator
commands a turn. All the loops are engaged except the y position loop since
the request is on vy . The turn trajectory is a circle of 10 m radius (R). Once the
operator issues the turn command, the helicopter flies the turn while maintaining
vy
vy = −4 m/s, vx = 0, constant altitude, ψ̇ = R, and pointing the tail towards the
center of the circle. The x loop is engaged to drive the horizontal tracking error to
zero.

• Nose-in Pirouette: Similar to the nose-out pirouette with vy = 4 m/s and helicopter
nose towards the center of the circle. The helicopter starts at hover. The operator
commands a 4 m/s step in vy . The helicopter flies sideways on a straight line until
the operator commands a turn. All the loops are engaged except the y position
loop since the request is on vy . The turn trajectory is a circle of 10 m radius (R).
Once the operator issues the turn command, the helicopter flies the turn while
vy
maintaining vy = 4 m/s, vx = 0, constant altitude, ψ̇ = R, and pointing the nose
towards the center of the circle. The x loop is engaged to drive the horizontal
tracking error to zero.

To analyze the data correctly, the following comments are necessary. The controller
was designed for hover operations. Wind gusts of 4-7 m/s were present throughout the
3.6. Loading analysis 71

tests. The trajectories were not designed to be easy to follow, but to test the controller in
limiting cases. In fact, all the turn trajectories would require the helicopter to be able to
follow, exactly, step changes in attitude and yaw rate to achieve zero tracking error. The
operator issued the turn commands following the indications of the safety pilot. This
procedure gave a higher level of confidence to the safety pilot, since the maneuvers
were never flown before. The safety pilot relies only on his vision to estimate the flying
speed, and he asked to start all the turns before the helicopter reached a steady straight
flight. Nevertheless, the tracking performance was remarkable.
Figure 3.24 shows the results for the forward coordinated turn. The two dotted
vertical lines in Figures 3.24(c)-3.24(f) mark the instants when the operator commands
straight and turning flight. The horizontal error is less than 2 m, while the vertical error
is less than 0.3 m. The presence of a constant wind can be clearly seen in Figure 3.24(c):
vx oscillates with a period equal to the time needed to complete a 360-degree turn (the
completion of the 360-degree turns are marked by the horizontal lines in Figure 3.24(f))
and thus the constant wind acts as a sinusoidal forcing term.
Figure 3.25 shows the results for the nose-out pirouette. The two dotted lines have
the same meaning described above. The horizontal error is less than 1 m, while the
vertical error is less than 0.2 m.
Videos of the maneuvers in Figures 3.24 and 3.25, and of all the other described
maneuvers, can be found on line at http://www.roboticflight.org.

3.6 Loading analysis

Some of the robotic helicopter’s missions will require extra payload that was not taken
into consideration during the controller design. For example, a video camera will be
mounted on the helicopter for reconnaissance and filming missions, and some long
distance missions will need extra fuel tanks. One of the advantages of a high-fidelity
model with physical parameters is the possibility to analyze the effect of parameter
3.6. Loading analysis 72

Desired flight path (solid), flight test (dash−dot) Desired flight path (solid), flight test (dash−dot)

0 10

−5 5
y (m)

z (m)
−10 0

−15 −5

−20 −10

−5 0 5 10 15 20 −5 0 5 10 15 20
x (m) x (m)
(a) Horizontal flight path. (b) Vertical flight path.
θ (dash−dot), φ (solid), dψ/dt (dashed)
7 10
θ and φ (deg), dψ/dt (deg/s)
6
0
5
Velocity (m/s)

4
Turn vx −10

vy
3 −20
v
z
2
−30
1
−40
0

−1 −50
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Time (s) Time (s)
(c) v x , vy , and vz . (d) θ, φ, and ψ̇.
δlon (solid), δlat (dash−dot), and δcol (dashed), δped (dotted)
1
0
δlon−lat−ped [−0.5, 0.5], δcol [0, 1]

−5

0.5 −10
ψ (rad)

−15

−20
0
−25

−30

−0.5 −35
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Time (s) Time (s)
(e) Controller outputs. (f) ψ.

Figure 3.24: Forward turn: the two vertical lines delimit hover, straight, and turn flight.
3.6. Loading analysis 73

Desired flight path (solid), flight test (dash−dot) Desired flight path (solid), flight test (dash−dot)

0 10

−5 5
y (m)

z (m)
−10 0

−15 −5

−20 −10

0 5 10 15 20 0 5 10 15 20
x (m) x (m)
(a) Horizontal flight path. (b) Vertical flight path.
θ (dash−dot), φ (solid), dψ/dt (dashed)
1 15

θ and φ (deg), dψ/dt (deg/s) 10


0
5

0
Velocity (m/s)

−1 vx
v −5
y
−2 v −10
z
−15
−3
Turn −20

−25
−4
−30

−5 −35
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Time (s) Time (s)
(c) v x , vy , and vz . (d) θ, φ, and ψ̇.
δlon (solid), δlat (dash−dot), and δcol (dashed), δped (dotted)

0
[0, 1]

0.8
col

−5
[−0.5, 0.5], δ

0.6
ψ (rad)

−10
0.4
lon−lat−ped

0.2 −15

0
−20
δ

−0.2
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Time (sec) Time (s)
(e) Controller outputs. (f) ψ.

Figure 3.25: Nose-out pirouette: the vertical lines delimit hover, straight, and turn flight.
3.6. Loading analysis 74

changes on performance and stability.


This section presents the effect of extra payload on the robustness of the inner-loop
controller. The robustness of the inner-loop controller will be quantified using single-
loop gain and phase margins. The objective of the analysis is to generate loading plots
that will enable the helicopter’s operator to safely add extra mass to the helicopter.
The first step of the analysis procedure is to generate a linear time-invariant (LTI)
model of the helicopter for each loading configuration. Two point masses, m1 and m2 ,
were added to the helicopter at coordinates ( x, y, z) and ( x, −y, z) with respect to the
nominal center of gravity of the helicopter (x, y and z are positive forward, starboard
and downwards respectively). It is assumed that m1 , m2 ∈ [0, 5] kg, x ∈ [−0.2, 0.6] m
and y, z ∈ [0, 0.4] m, and a grid with 6912 configurations was selected. For each loading
configuration it was straightforward to calculate the new mass, center of gravity and
moments of inertia of the helicopter, and to generate an LTI model using MOSCA.
The second step of the analysis procedure is to compute the single-loop gain and
phase margins for each loading configuration. Instead of breaking the closed-loop sys-
tem at each actuator and sensor and calculating the gain and phase margins of the
resulting transfer function, lower bounds of the worst case single-loop gain and phase
margins were computed. The optimization used to calculate the lower bounds is com-
putationally efficient and does not typically result in conservative lower bounds for
robust closed-loop systems.
Let G ∆ denote the helicopter model at some loading configuration, W 1 and W 2 the
shaping weights, and K the inner-loop controller. The optimization solved is
  −1
I
h i
− 1 − 1

max D   ( I − W 2 G∆W 1 K ) I W 2 G∆W 1 D : = ǫµ (3.16)
D
K

where D is  
D2
D=  (3.17)
D1−1
3.6. Loading analysis 75

and D1 and D2 are restricted to be diagonal, stable, and minimum phase. Figure 3.26
shows the setup. The idea is that pre and post multiplying the shaped plant W 2 G ∆ W 1

z2 w2 w1 z1

D1 W 2 G∆W 1 D2

D1−1 K D2−1

Figure 3.26: Setup for the optimization in Equation 3.16.

by D1 and D2 respectively, and pre and post multiplying the controller K ∞ by D2−1 and
D1−1 respectively, does not alter the gain and phase margins. Thus, the optimization
over D in Equation 3.16 gives less restrictive guaranteed gain and phase margins. Note
that without D, Equation 3.16 is identical to Equation 3.10.

ε vs y and z at x = 0.02 m with m = 5 kg and m = 1.67 kg


µ 1 2

0.36

0.34

εµ 0.32

0.3

0.28
0

0.2 0.4
0.3
0.2
y (m) 0.1
0.4 0 z (m)

Figure 3.27: Loading plot for two fuel tanks of different mass at x = 0.02 m.

The lower bounds of the worst case single-loop gain and phase margins are ob-
3.7. Conclusions on robust hover control 76

tained by substituting ǫµ into Equation 3.12. The power of Equation 3.16 lies in the fact
that ǫµ is also related to a multi-input/multi-output robustness test. The reader should
refer to Papageorgiou and Hyde [2001] for a complete explanation of how ǫµ is related
to single-loop and multi-loop gain and phase margins. Note that optimization 3.16 is
straightforward to solve using commercial software as, for example, the µ-tools com-
mand mu [Balas et al., 1995].
Figure 3.27 illustrates a typical loading plot. It is evident from Figure 3.27 that
adding two extra fuel tanks of different mass at x = 0.02 m anywhere in the region
of the y-z plane considered does not decrease ǫµ significantly from its nominal value.
The worst case ǫµ is 0.274 which corresponds to lower bounds of 4.88 dB and 31.8 de-
grees.

3.7 Conclusions on robust hover control

The synergistic use of a high-fidelity simulation model and the H∞ loop-shaping design
method proved to be an effective strategy for the rapid and reliable development of a
high-bandwidth hover controller for a robotic helicopter. H∞ loop shaping enables the
engineer to tackle the problem in an intuitive and elegant fashion. The design space
can be explored quickly, and it is simple to understand what level of closed-loop per-
formance is achievable from the system. Using classical notions of bandwidth and loop
gain results in clear and intuitive choices on performance and robustness trade-off.
The synthesized high-bandwidth controller easily satisfied a set of handling-quality
specifications that are a standard in the full-scale military-rotorcraft community. The
implementation of the controller on the on-board computer was straightforward. Fur-
thermore, integrating all the tools used in the design, the whole process from controller
modification to flight test is automated, making it possible to tune the controller in
flight. The new controller was capable of flying several moderate-speed coordinated
maneuvers. The effect of typical extra-loading conditions on the robustness of the con-
3.7. Conclusions on robust hover control 77

troller is analyzed. The hover controller is the starting point for the development of a
controller for wide-envelope flight presented in the next chapter.
Chapter 4
Robust full-envelope flight control

4.1 Observer form for scheduling . . . . . . . . . . . . . . . . . . . . . . . 78

4.2 Controller structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

4.3 Inner-loop design for fixed operating points . . . . . . . . . . . . . . . 83

4.4 Scheduling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

4.5 Scheduled inner-loop evaluation . . . . . . . . . . . . . . . . . . . . . 93

4.6 Outer loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

4.7 Controller implementation . . . . . . . . . . . . . . . . . . . . . . . . . 109

4.8 Flight tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

Complex future missions in urban and remote environments will require robotic
helicopters to operate at the edge of their flight envelope. This chapter shows how to
use effectively MOSCA with the H∞ loop-shaping framework to produce robust full-
envelope flight controllers.

78
4.1. Observer form for scheduling 79

4.1 Observer form for scheduling

The helicopter dynamics vary widely across the flight envelope. A single linear con-
troller cannot cope with this highly non-linear behavior, and thus cannot deliver good
stability and performance levels across the full flight envelope. One technique to over-

+
r P(ρ) W 1 (ρ) G y
+
W 2 (ρ)

− +

H (ρ)
+ +
Bs ( ρ )
+
x̂s R
F (ρ)

As (ρ )

ŷs
C s (ρ)

State Observer
K (ρ)

Figure 4.1: Observer form.

come this problem consists of designing several linear controllers on a grid in the do-
main that covers the flight-envelope region of interest, and then switching among the
controllers as the vehicle flies across the different operating points. Another well-
known technique is gain scheduling.

Scheduling — in its most straightforward form — consists of interpolating the en-


4.2. Controller structure 80

tries of the state-space matrices of the controllers designed for different operating points.
This interpolation requires that the controller states maintain the same physical inter-
pretation and do not change in number for all the controllers that need to be scheduled.
One of the advantages of the H∞ loop-shaping approach is that the controller can be
written as an exact plant observer plus state feedback [Sefton and Glover, 1990]. This
means that the controller has a well defined structure that can be used for scheduling.

Figure 4.1 shows a H∞ loop-shaping controller implemented in the observer form.


The synthesized controller is enclosed in the dashed line. The overall structure is analo-
gous to the one already presented for the hover case (see Figure 3.5): W 1 (ρ) and W 2 (ρ)
are the weights used to shape the plant, and P(ρ) is a pre-filter. The difference is in
K (ρ), which is now implemented as a state observer (shaded region) for the shaped
plant, plus a full state feedback matrix F (ρ). ρ is the scheduling variable, which in
general can be a vector.

4.2 Controller structure

In an inner-loop/outer-loop structure, it is often necessary to schedule only the inner


loop. In fact, the scheduled, and thus non-linear, inner-loop controller cancels the non-
linearity of the plant if similar levels of closed inner-loop performance are requested
and achieved across the flight envelope. The obvious choice to produce a wide–flight-
envelope controller, would have been to schedule the inner loop of the hover controller
presented in Chapter 3, but an initial attempt in this direction presented several prob-
lems.

The main problem was the appearance of very low frequency right-half-plane (RHP)
transmission zeros at different operating points. RHP transmission zeros impose limi-
tations on the achievable bandwidth or robustness and make the controller design dif-
4.2. Controller structure 81

ficult. For example, linear models with inputs u and outputs y


   
δ θ
 lon   
 δlat  φ
   
u=
 ,
 y=
 ,
 (4.1)
 δcol   vz 
   
δped ψ̇

extracted at speeds vx < −3 m/s in straight level-flight trim have RHP real zeros at
frequencies between 0.002 and 0.01 rad/s. For the same set of inputs and outputs,
complex conjugate and real RHP zeros appear also in left coordinated-turn trims; for
example at a speed of 15 m/s and turn rate lower than zero (left turn) the RHP zeros
are at frequencies between 0.09 and 0.35 rad/s.
Other problems were those experienced in Section 3.3 during the design of the outer
loops for the hover controller: (1) a decrease in robustness at the plant input with the
closure of the velocity outer loops, and (2) difficulty in satisfying the bandwidth specifi-
cation for the ψ channel, with an outer loop on ψ̇, without violating the stability margins
at the plant input.
The remedy for this situation consists of modifying the structure of the inner loop as
follows: vx and vy are added to the output variables of the plant to eliminate the RHP
transmission zeros, and as suggested in Section 3.3, the ψ̇ signal is replaced by ψ. The
new outputs are

y = [θ v x φ vy vz ψ] T (4.2)

Since there are four inputs we can control only four outputs. Blending vx and θ in
a longitudinal variable (vθ ), and vy and φ in a lateral variable (vφ ) reduces the number
of output variables to four. The blending utilizes the attitude signals (θ and φ) at high
frequency and the velocity signals (vx and vy ) at low frequency. This is possible by
combining the variables as follows:

kθ s
vθ = θ + kvx vx (4.3)
s + pθ
4.2. Controller structure 82

kφ s
vφ = φ + k vy vy (4.4)
s + pφ

The parameters kθ , pθ , kv x , kφ , pφ , and kvy are chosen by the designer. For vθ , for example,
kθ is used to set the crossover frequency of vθ (i.e., θ), and kv x is used to set the crossover
frequency of vx (or the frequency up to where it is desired that the vx is larger than θ in
magnitude). kv x is also chosen such that the vx term does not affect gain and phase be-
havior of vθ at crossover [Hyde, 1995] (i.e., at crossover the attitude channel dominates
the dynamics). The high-pass filter on the attitude signal is needed to remove the atti-
tude trim value, which is different from zero in general, and thus allows the controller
to hold a constant velocity. The location of the high-pass filter pole pθ involves some
trade-offs: increasing its value increases robustness (i.e., introduces phase lead), but in-
creases overshoot; decreasing its value improves overshoot but increases settling time
due to the longer time needed to wash out the attitude trim value. Figure 4.2 illustrates
an example of the vθ blending for a linear model at 10 m/s forward speed.

vθ (solid), vx (dashed), θ (circles)


60
Magnitude (dB)

40
20
0
−20
−40
−60
180
90
Phase (deg)

0
−90
−180
−270
−360
−2 −1 0 1
10 10 10 10
Frequency (rad/s)

Figure 4.2: Blending of vx and θ outputs at 10 m/s forward speed.


4.2. Controller structure 83

Figure 4.3 shows the new inner-loop controller structure in the observer form (see
also Figure 4.1). The blending is included in the W 2 weight, which hence has six inputs
and four outputs (see Equation 4.6). Another option would be to use the single elements
of the blending in a diagonal W 2 with six inputs and six outputs. This would leave the
actual blending of the plant outputs to the controller.

x
y
z
v xr lon
θ
v yr lat
vx
v zr P W1 col R−50
φ
ψr K ped
vy
vz
ψ
k
s I

W2

Figure 4.3: Inner loop.

The new inner-loop structure presents several advantages:

• The overall controller structure is simpler, since it eliminates the need for velocity-
control outer loops.

• It eliminates the decrease in stability margin at the actuator inputs observed in


the hover design and caused by the presence of the velocity and yaw-angle outer
loops.

• It can be used by a human operator with no particular skills to fly the helicopter
controlling directly the velocity vector.
4.3. Inner-loop design for fixed operating points 84

• It makes it still possible to set independently the bandwidth for velocity and atti-
tude responses by adjusting the parameters in Equations 4.3-4.4.

Furthermore, it is possible to control directly θ, φ, and φ̇ designing appropriate pre-


filters on the reference signals.
The complete wide-envelope controller architecture implemented and flight tested
on the CMU R-50 is composed of the scheduled inner loop and three position outer
loops similar to the ones presented for the hover controller. The outer loops are not
scheduled since, as shown in the following sections, the gain-scheduled inner loop con-
troller (Figures 4.1 and 4.3) delivers a system with dynamic characteristics that do not
vary widely in the target flight envelope.

4.3 Inner-loop design for fixed operating points

The purpose of the scheduled inner-loop controller (Figures 4.1 and 4.3) is to provide
robust stabilization and decoupling over the helicopter’s flight envelope. The next sec-
tion shows how to choose the scheduling variables, the number of fixed operating-point
controllers, and the interpolation functions. Assume for now that the locations in the
flight envelope of the fixed operating-point controllers are known, and concentrate on
the common design procedures and characteristics of these fixed operating-point con-
trollers.
The inner-loop controller controls the 4 variables of Section 4.2 (see Figure 4.3): (1)
a blend of pitch attitude θ with forward velocity vx , (2) a blend of roll attitude φ with
lateral velocity vy , (3) vertical velocity vz , and (4) yaw angle ψ. In normal operations,
the references are vxr , vyr , vzr , and ψr , but it is possible to control also θ, φ, and ψ̇ using
appropriate pre-filters.
The requirements for the scheduled inner-loop controller, and therefore for the fixed
operating-point controllers, are similar to those used for the hover design (see page 49).
The differences are the following:
4.3. Inner-loop design for fixed operating points 85

Robust stability. Stability margin ǫ ≥ 0.3, which guarantees gain and phase margins
larger than ±5.4 dB and 34.9 deg respectively. The choice of ǫ ≥ 0.3 does not
relax the stability-margin requirements used for the hover design. An ǫ = 0.3
does not mean that gain and phase margins in broken single-loop analysis are
exactly ±5.4 dB and 34.9 deg. Rather, the margins are guaranteed to be greater
than or equal to ±5.4 dB and 34.9 deg respectively. In practice ǫ ≥ 0.3 will result
in designs which satisfy the usual specification that gain and phase margins shall
be ≥ 6 dB and ≥ 45 deg respectively.

Bandwidth. Speed of response for the vx and vy channels defined on page 64 becomes
an inner-loop specification.

Decoupling. For the decoupling between vertical response and yaw rate, ADS-33E
takes in consideration the transfer function r/vzr , where r is the yaw rate in body
frame. Since the inner-loop controller presented here controls ψ, the substitution
of r with ψ̇ gives a more meaningful evaluation of the achieved level of decou-
pling, especially during coordinated-turns.

Given the structure of the inner-loop controller, the bandwidth and decoupling require-
ments for the θ and φ channel are enforced on vθ and vφ . This choice is appropriate,
since in the cross-over–frequency region of vθ and vφ , θ and φ dominate the responses
(see Figure 4.2).

For scheduling, the structure of the controllers must be the same at each operat-
ing point. This is ensured by designing identical structures for the W 1 (ρ) and W 2 (ρ)
shaping weights where ρ indicates the dependence of the weights on the scheduling
variables. With this choice, the observer form, which observes the shaped plant (see
Figure 4.1) ensures that the synthesized controller has a fixed structure. The structures
of the W 1 (ρ) and W 2 (ρ) weights are chosen to be as follows for all of the operating
points. The diagonal W 1 (ρ) structure contains proportional plus integral action on all
4.3. Inner-loop design for fixed operating points 86

the axes:
s + zi1
 
ki1 ( ρ) s
 
 s + zi2 
 ki2 ( ρ) s

W 1 ( ρ) = 
 s + zi3

 (4.5)
 ki3 ( ρ) s

 
s + zi4
ki4 ( ρ) s

The W 2 (ρ) structure contains second-order low-pass filters for sensor noise rejection on
all channels and the blending filters of Equations 4.3 and 4.4.

ωn2
 
ko1 ( ρ)
 s +2ζωn + ωn2 
 ωn2 
ko2 ( ρ)
s +2ζωn + ωn2
 
W 2 ( ρ) =  
 ωn2 
ko3 ( ρ)
s +2ζωn + ωn2
 
 
ωn2
ko4 ( ρ)
s +2ζωn + ωn2
  (4.6)
s
k θ (ρ) s+ p k v x (ρ)
 θ 
 s 
 k φ (ρ) s+ p k vy (ρ) 
 φ 
 
 1 
 
1

The gains ki1−4 (ρ) of W 1 and ko1−4 (ρ) of W 2 are tuned to have crossover frequencies of
the open-loop system at 5 rad/s for the blended channels and 4 rad/s for the vz and ψ
channels. This tuning is done automatically for every design point using the singular
values of the shaped plants as illustrated in Chapter 3.
After calculating the weights, the plant is shaped and the controller is calculated
and expressed in observer form. A reduced-order model (see Section 2.5) is used to
synthesize the controller. The choice of the pre-filter P(ρ) (Figures 4.1 and 4.3) depends
on its purpose. At any single operating point P is calculated as follows. If the transient
response of the closed loop is satisfactory, only the steady-state gain from references to
outputs must be enforced. To have this gain equal to I, P is

P = T ( 0) − 1 (4.7)
4.3. Inner-loop design for fixed operating points 87

where T (0) is the closed-loop transfer function matrix from the references to the vx , vy , vz ,
and ψ outputs evaluated at frequency ω = 0.
If the closed-loop responses need to be improved, dynamic pre-filters are designed.
Suppose it is desired that the closed-loop transfer function Tv x from vxr to vx matches
the transfer function of an ideal system To ; the SISO pre-filter Pv x is then calculated by
minimizing over Pv x the H∞ -norm of the transfer function from d to [e1 , e2 ]T in Fig-
ure 4.4. In Figure 4.4, We1 and We2 are weights to emphasize low-frequency matching

T0

-
d Pvx Tvx We1 e1

We2 e2

Figure 4.4: Setup for the design of dynamic pre-filters.

and reduce high-frequency content of the pre-filter respectively.


The same procedure can be applied to the other channels. P is then assembled as
 
Pv x
 
Pvy
 
 
P= 

 (4.8)
 Pvz 
 

Dynamic pre-filters should not be used to correct a bad controller design, since they
could result in poor robust performance [Papageorgiou, 1998]. The procedure works
well if the bandwidth of To is smaller than the closed-loop bandwidth [Papageorgiou,
1998]. A dynamic pre-filter calculated as in Figure 4.4 requires special care during
scheduling, since its states do not necessarily maintain the same physical meaning at
each operating point. In this case, scheduling can be done on the outputs; the technique
4.4. Scheduling 88

consists of simulating the pre-filters in parallel and scheduling their outputs. Another
possibility is to switch between pre-filters as the helicopter traverses the flight envelope.
Four anti-windup loops are finally added for each actuator as in the hover design
of Chapter 3. The gains of the anti-windup loops are the same for each operating point.
The next section shows how to choose the operating points for which this design pro-
cedure is applied.

4.4 Scheduling

Scheduling is necessary to ensure robust stability and performance across the flight
envelope. The target flight envelope tested in this thesis is represented by the following
limits:
−10 m/s ≤ vx ≤ 20 m/s
−2 m/s ≤ vy ≤ 2 m/s
(4.9)
−2 m/s ≤ vz ≤ 2 m/s
−30 deg/s ≤ ψ̇ ≤ 30 deg/s
There are three interrelated aspects in scheduling multiple linear controllers: (1) choos-
ing the interpolation functions, (2) determining the variables for the scheduling, and (3)
establishing the number of fixed operating-point controllers.
The simplest choice for the interpolation functions (the one that was implemented
for the flight tested controller), is linear interpolation. Supposing that the scheduling is
on one variable ρ and that the fixed operating points where the controllers need to be
designed are known, the matrices — for example H (ρ) — of the scheduled controller
K (ρ) (Figure 4.1) between two adjacent operating points i, j are calculated as

H (ν) = (1 − ν) H i + νH j (4.10)

where ν varies in the interval [0, 1] and is

ρ − ρi
ν= . (4.11)
ρ j − ρi
4.4. Scheduling 89

Linear interpolation can easily be applied to more scheduling variables. In case of two
variables ρa and ρb the expression for H (ρa , ρb ) in the square composed of the corners
ia , ja , ib , and jb is

H (νa , νb ) = (1 − νa )[(1 − νb ) H ia ,ib + νb H ia ,jb ] + νa [(1 − νb ) H ja ,ib + νb H ja ,jb ], (4.12)

where
ρa − ρ ia ρb − ρ ib
νa = , and νb = . (4.13)
ρ ja − ρia ρ jb − ρib
Having chosen the interpolation function it is now possible to check which of the
variables that define the flight envelope need to be scheduled and then determine the
number and positions of the fixed-point controllers across the scheduled-variable range.
This procedure starts from the hover controller (designed as illustrated in the previous
section), and uses the stability margin as a design tool. Recall that, given a plant G,
shaping weights W 1 and W 2 , and the H∞ loop-shaping controller K, the stability mar-
gin of their feedback connection is (see Equation 3.10)
  −1
I
h i
  ( I − W 2 GW 1 K )−1 I W 2 GW = ǫ.

b(W 2 GW 1 , K ) =
1 (4.14)
K

The scheduling procedure is divided in two parts. For the sake of clarity, suppose
that only vx and vy are used to define the flight envelope (i.e., the helicopter is only
allowed to move forward and laterally, with neither turning nor vertical flight). It is also
assumed that the fixed-point controllers are designed such that their stability margin is
greater than 0.3 at the respective operating points. The steps of the first part, which will
determine the variables to schedule, are the following.

1. Start at the hover operating point (Γ(0,0) in Figure 4.5), design a controller (W 1(0,0) ,
W 2(0,0) , and K (0,0) ), and choose a variable in the 2-dimensional domain (e.g., vx )

2. Extract linear models of the helicopter at the lower and upper bounds of the
chosen-variable range of variation, and at the conditions of the current operat-
ing points for the other variable (Γ(−10,0) and Γ(20,0) in Figure 4.5).
4.4. Scheduling 90

3. Check if
b(W 2(0,0) G (−10,0)W 1(0,0) , K (0,0) ) ≥ 0.3,
b(W 2(0,0) G (20,0) W 1(0,0) , K (0,0) ) ≥ 0.3.

If the conditions are not satisfied, vx needs to be scheduled. In this case, design
controllers at Γ(−10,0) and Γ(20,0) and check that

b(W 2(i,0) G(i,j) W 1(i,0) , K (i,0) ) ≥ 0.3

with

i = [−10, 0, 20] and j = [−2, 2].

If any of these conditions is not satisfied, vy also needs to be scheduled. Hence,


the first part of the scheduling procedure is concluded with the determination of
the variable to schedule.

4. In case that the first conditions in the previous step are satisfied, check that the
hover controller gives a stability margin greater than 0.3 at the remaining six
points on the boundary in Figure 4.5 (i.e., Γ(−10,−2) , Γ(−10,2) , Γ(0,−2) , Γ(0,2) , Γ(20,−2) ,
and Γ(20,2) ). If the condition is satisfied, neither vx nor vy need to be scheduled.
Otherwise, vy is the only variable that is scheduled.

Supposing that the first part of the scheduling procedure showed that only vx needs
to be scheduled, the steps of the second part, which will determine the position of the
fixed-point controllers across vx , are the following.

1. Divide the domain represented by vx and vy in a grid of sufficient resolution in the


direction of the scheduled variable. For example, choose ∆vx = 1 m/s (30 vx grid
points). ∆vy is chosen such that the grid points along vy are the boundary and
the vy = 0 points. For the case under consideration, ∆vy = 2 m/s. Extract linear
models at every grid point. In case ψ̇ is a scheduling variable, ∆ψ̇ = 5 deg/s is
sufficient.
4.4. Scheduling 91

Γ(−10,−2) Γ(0,−2) Γ(20,−2)

Γ(−10,0) Γ(0,0) Γ(20,0)


vx

Γ(−10,2) Γ(0,2) Γ(20,2)

vy

Figure 4.5: Boundaries of a flight-envelope slice for vz = 0 and ψ̇ = 0.

2. Generate controllers at Γ(l,0) and Γ(m,0) . The first time, start with l and m at the
lower and upper bound of the scheduling-variable range of variation (i.e., for vx ,
l = LB = −10 and m = UB = 20).

3. Schedule the controller between Γ(l,0) and Γ(m,0) as shown in Equations 4.12 and
4.13.

4. Check if

b(W̄ 2(l +kv x ∆v x , 0) G(l +kv x ∆v x , kvy ∆vy ) W̄ 1(l +kv x ∆v x , 0) , K̄ (l +kv x ∆v x , 0) ) ≥ 0.3
(m − l )
∀ kvx ∈ [1, 2, . . . , − 1], and ∀ kvy ∈ [−1, 0, 1], (4.15)
∆vx

where the bar over a system indicates that the system is scheduled. Equation 4.15
evaluates the stability margin of the scheduled controller at every point of a strip
(between l and m) of the grid defined in step 1.

5. If Equation 4.15 is not satisfied, use steps 2-4 with a bisection algorithm (moving l)
over the vx grid points to find the maximum m − l that satisfies Equation 4.15.
4.4. Scheduling 92

6. If Equation 4.15 is satisfied, register the current Γ(l,0) . Then substitute Γ(m,0) with
Γ(l,0) and Γ(l,0) with Γ(LB,0) and iterate steps 2-5 until Equation 4.15 has been sat-
isfied at all operating points between Γ(−10,0) and Γ(20,0) . The registered Γ(l,0) are
the operating points along vx needed for the scheduling together with Γ(LB,0) and
Γ(UB,0) .

7. If Equation 4.15 is not satisfied not even for l = m − 1, then vy needs to be sched-
uled.

The whole scheduling procedure can be extended to multiple variables, and can be com-
pletely automated. This was indeed done for the synthesis of the flight-tested controller
of the CMU R-50, and the results of the procedure are presented in the next section.
If the choice of the weighting functions is done properly, the use of the stability
margin for scheduling not only guarantees robustness throughout the desired flight
envelope, but guarantees also that the achieved loop shapes are reasonably close to
the desired ones (i.e., closed-loop objectives are most likely satisfied by the scheduled
controller). Scheduling with the use of gain and phase margins instead, does not allow
one to draw clear performance-related conclusions and necessitates many more time
simulations. Consider the following. Figure 4.7 shows the broken-loop gain and phase
margin at the inputs of the plant achieved by a hover controller with linear models
generated between vx = 0 m/s and vx = 20 m/s. The scheduling procedure carried
on using gain and phase margins, would conclude that the hover controller could be
used up to 20 m/s without any scheduling, since it satisfies the standard requirements
of 6 dB and 45 degrees for gain and phase margin respectively. Figure 4.6 shows the
stability margin for the same situation: above 6 m/s, ǫ drops below 0.3 indicating the
need for scheduling. Figure 4.8 shows some step responses at hover, 6 m/s, and 20 m/s.
Clearly, the level of performance is well predicted by ǫ: the hover controller at 6 m/s
still delivers good decoupling and the on-axis responses are close to the nominal ones;
at 20 m/s, instead, performance deteriorates considerably.
4.5. Scheduled inner-loop evaluation 93

14

lon
12 lat

GM (dB)
col
ped
10

lon
85 lat
col
PM (deg)

ped

65

45
0 5 10 15 20
vx (m/s)

Figure 4.6: GM and PM of the hover controller for vx ∈ [0, 20] with ∆vx = 1 m/s.

0.4
ε

0.3

0.2

0 5 10 15 20
vx (m/s)

Figure 4.7: ǫ of the hover controller for vx ∈ [0, 20] with ∆vx = 1 m/s.

4.5 Scheduled inner-loop evaluation

The application to the CMU R-50 of the scheduling procedure introduced in the pre-
vious section results in the following: to cover the flight envelope described in Equa-
tion 4.5, the controller needs to be scheduled on vx and ψ̇. There are 30 requested fixed–
4.5. Scheduled inner-loop evaluation 94

v step reference to v
z z
vz step reference to ψ
0.02
Amplitude (normalized)

Amplitude (normalized)
1 0.01

0
0.8
−0.01
@ 0 m/s
0.6
@ 6 m/s
−0.02
@ 20 m/s
0.4 @ 0 m/s
−0.03
@ 6 m/s
0.2 @ 20 m/s
−0.04

0 −0.05
0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s) Time (s)

vx step reference to vz ψ step reference to ψ


0.25
Amplitude (normalized)

Amplitude (normalized)
0.2 1
@ 0 m/s
0.15
@ 6 m/s 0.8
@ 20 m/s @ 0 m/s
0.1
0.6 @ 6 m/s
0.05 @ 20 m/s
0.4
0

−0.05 0.2

−0.1 0
0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s) Time (s)

Figure 4.8: Selected on-axis and off-axis step responses for evaluation of the hover con-
troller performance at vx = 0, 6, 20 m/s.

operating-point controllers at

vx = [−10, −6, 0, 6, 12, 20] m/s and ψ̇ = [−30, −15, 0, 15, 30] deg/s. (4.16)

Figures 4.9, 4.10, and 4.11 show the ǫ values achieved by the scheduled controller at
some slices of the flight-envelope. The values satisfy the requirement for robust stability
given on page 84.
Figures 4.12, 4.13, and 4.14 show the vθ , vφ , and ψ bandwidths achieved by the
scheduled controller at some sections of the flight-envelope. The values satisfy the
4.5. Scheduled inner-loop evaluation 95

0.38

0.36

0.34
ε

0.32

0.3

0.28

−20

0 20
15
dψ/dt (deg/s) 10
20 5
0
−5 vx (m/s)
−10

Figure 4.9: ǫ vs vx and ψ̇ for vy = 0 and vz = 0.

0.38

0.36

0.34
ε

0.32

0.3

0
15 20
vz (m/s) 5 10
−5 0
−2 −10
v (m/s)
x

Figure 4.10: ǫ vs vx and vz for vy = 0 and ψ̇ = 0.


4.5. Scheduled inner-loop evaluation 96

0.38

0.36

0.34
ε

0.32

0.3

2
1
0
vy (m/s)
−1 15 20
5 10
−2 −5 0
−10
vx (m/s)

Figure 4.11: ǫ vs vx and vy for vz = 0 and ψ̇ = 0.

bandwidth requirements given on page 49.

For reference and future comparisons with other controllers, Figures 4.15, 4.16, and
4.17 show the vθ , vφ , and ψ bandwidths evaluated as the frequency where the gain of
the closed-loop response is equal to -3 dB.

Figures 4.18, 4.19, and 4.20 show the performance of the vz channel with RTvz , τvz
and r2 . The values satisfy the bounds given on page 50.

Figures 4.21 and 4.22 show the rise time of the vx and vy channels. At several points
in the grid, both RTv x and RTvy violate the lower bound given on page 64. This is a
typical case were it is safe to use a dynamic prefilter (see Section 4.3) to slow down the
response. The same result can be obtained by decreasing kv x and kvy in the blending fil-
ters of Equations 4.3 and 4.4. In fast coordinated turns, there is also an increase in RTv x ,
which does not violate the bounds, but indicates that the vx response becomes sluggish.
The proper way to decrease RTv x is to increase kv x in the blending filter (i.e., increase the
vx crossover frequency), or to tune the zero in the first entry of W 1 (ρ) (Equation 4.5).
4.5. Scheduled inner-loop evaluation 97

Note that the zeros of W 1 (ρ) are chosen at hover and they are not scheduled.
Figures 4.23, 4.24, 4.25, and 4.26 show the time t10% needed for the θ, φ, vz , and ψ
responses to return to less than 10% of their peak values following a pulse input in the
corresponding actuator. The values satisfy the bounds given on page 51.
Figures 4.27, 4.28, and 4.29 show the evaluation of the decoupling performance. The
ψ̇3
values obtained satisfy the decoupling requirements given on page 50, except for |vz (3)|
(Figure 4.29) that slightly exceeds the limit of 0.2 in steep/right coordinated turns.

8.5

7.5
ωBW

6.5
20 20
0 10
−20 0
dψ/dt (deg/s) −10 vx (m/s)

Figure 4.12: vθ channel: ωBW vs vx and ψ̇ for vy = 0 and vz = 0.

To understand why the controller delivers this excellent performance, it is worth-


while analyzing its structure. Sefton and Glover [1990] showed that a controller synthe-
sized with H∞ loop shaping can be written as a state observer for the shaped plant plus
state feedback (see Figure 4.1). This fact implies that, if the model used for designing
the controller is of high order (i.e., contains the rotor states), the observer estimates the
rotor states, which are then used for compensation by the state-feedback matrix F. It
is well known in the helicopter community [Hall and Bryson, 1973, Takahashi, 1994a,b,
Howitt et al., 2001] that rotor-state feedback has a great effect in increasing bandwidth,
4.5. Scheduled inner-loop evaluation 98

11

10.5
ωBW

10

9.5
20
20
0
10
dψ/dt (deg/s) −20 0
−10 vx (m/s)

Figure 4.13: vφ channel: ωBW vs vx and ψ̇ for vy = 0 and vz = 0.

8
ωBW

6
20
0 20
10
dψ/dt (deg/s) −20 0
−10 vx (m/s)

Figure 4.14: ψ channel: ωBW vs vx and ψ̇ for vy = 0 and vz = 0.


4.5. Scheduled inner-loop evaluation 99

4
ω
−3dB
3

20
20
0
10
dψ/dt (deg/s) −20 0
−10 vx (m/s)

Figure 4.15: vθ channel: ω−3dB vs vx and ψ̇ for vy = 0 and vz = 0.

4.5

ω 4
−3dB

3.5

2.5

20

0 20
10
dψ/dt (deg/s) −20 0
−10 vx (m/s)

Figure 4.16: vφ channel: ω−3dB vs vx and ψ̇ for vy = 0 and vz = 0.


4.5. Scheduled inner-loop evaluation 100

4.5

3.5
ω
−3dB
3

2.5

20

0 20
15
dψ/dt (deg/s) 10
−20 5
0
−5
−10 vx (m/s)

Figure 4.17: ψ channel: ω−3dB vs vx and ψ̇ for vy = 0 and vz = 0.

0.36
0.34
0.32
RT

0.3
0.28
0.26

20

0 20
10
dψ/dt (deg/s) −20 0
−10 v (m/s)
x

Figure 4.18: vz channel: RTvz vs vx and ψ̇ for vy = 0 and vz = 0.


4.5. Scheduled inner-loop evaluation 101

0.06

0.05

τ
0.04

0.03

20

0 20
15
10
dψ/dt (deg/s) −20 5
0
−5
−10 vx (m/s)

Figure 4.19: vz channel: τvz vs vx and ψ̇ for vy = 0 and vz = 0.

0.995

r2
0.99

0.985
20
0 20
10
dψ/dt (deg/s) −20 0
−10 vx (m/s)

Figure 4.20: vz channel: r2 vs vx and ψ̇ for vy = 0 and vz = 0.


4.5. Scheduled inner-loop evaluation 102

3.5

3
RT

2.5

1.5

20

0 20
10
dψ/dt (deg/s) −20 0
−10 v (m/s)
x

Figure 4.21: vx channel: RTv x vs vx and ψ̇ for vy = 0 and vz = 0.

2.5

2
RT

1.5

1
20
0 20
10
dψ/dt (deg/s) −20 0
−10 vx (m/s)

Figure 4.22: vy channel: RTvy vs vx and ψ̇ for vy = 0 and vz = 0.


4.5. Scheduled inner-loop evaluation 103

2.5
t10%
2

1.5

20

0 20
15
10
−20 5
dψ/dt (deg/s) 0
−5
−10 vx (m/s)

Figure 4.23: θ channel: t10% vs vx and ψ̇ for vy = 0 and vz = 0.

2.5
t10%

1.5
20
0 20
10 15
dψ/dt (deg/s) −20 0 5
−10 −5
vx (m/s)

Figure 4.24: φ channel: t10% vs φ and ψ̇ for vy = 0 and vz = 0.


4.5. Scheduled inner-loop evaluation 104

2.1

2
t10%

1.9

1.8
20
0 20
10 15
dψ/dt (deg/s) −20 0 5
−10 −5
vx (m/s)

Figure 4.25: vz channel: t10% vs vx and ψ̇ for vy = 0 and vz = 0.

2.2

2
t10%

1.8

1.6

20

0 20
15
10
dψ/dt (deg/s) −20 5
0
−5
−10 vx (m/s)

Figure 4.26: ψ channel: t10% vs vx and ψ̇ for vy = 0 and vz = 0.


4.5. Scheduled inner-loop evaluation 105

0.15

0.1

0.05

0
20
0 20
10
dψ/dt (deg/s) −20 0
−10 vx (m/s)

∆θpk
Figure 4.27: ∆φ4 vs vx and ψ̇ for vy = 0 and vz = 0.

0.15

0.1

0.05

0
20
0 20
10
dψ/dt (deg/s) −20 0
−10 vx (m/s)

∆φpk
Figure 4.28: ∆θ4 vs vx and ψ̇ for vy = 0 and vz = 0.
4.5. Scheduled inner-loop evaluation 106

0.3

0.25

0.2

0.15

0.1

0.05

0
20
0 20
10
dψ/dt (deg/s) −20 0
−10
vx (m/s)


ψ̇ ψ̇3
Figure 4.29: vz (13) ≃ |vz (3)| vs vx and ψ̇ for vy = 0 and vz = 0.

reducing cross coupling, improving robustness and disturbance rejection, and reducing
actuator usage. Hall and Bryson [1973] showed that high gain necessitates rotor-state
feedback to avoid instability of the coupled rotor-fuselage mode. Using a mathematical
model, they showed also that (1) low-quality rotor measurements improve marginally
the tracking performance with respect to accurate body measurements and rotor-state
estimation, and (2) accurate body measurements and rotor-state estimation result in bet-
ter disturbance rejection with respect to the case when rotor states are measured with
low quality. Howitt et al. [2001] have evaluated the performance of a H∞ loop shaping
controller on a model-scale rotor rig showing the benefit of rotor-state estimation and
feedback.

It was thus hypothesized that H∞ loop shaping had the potential of excellent per-
formance if used in combination with accurate and high-order models of the CMU R-50
as the ones delivered by MOSCA. The results shown have confirmed the hypothesis. It
should be emphasized that in the H∞ loop shaping framework, the feedback and esti-
4.5. Scheduled inner-loop evaluation 107

mation of rotor states do not require any additional design effort. In fact, the designer
follows the same exact procedure as if only rigid-body states were used for feedback.
The frequency shaping maintains its transparency allowing simple trade-offs between
robustness and performance, while the H∞ optimization exploits the high order dynam-
ics synthesizing the optimal observer and feedback matrix for decoupling and robust
stability.

The analysis of the state-feedback matrix makes it possible to gain physical insight
about the use that the controllers makes of high-order dynamics. For the controller de-
veloped in this thesis, the analysis (shown below) resulted in the following finding: the
improvement of stability and performance with respect to rigid-body-states-only feed-
back is caused mainly by the feedback of the stabilizer-bar states. That is, estimating
only the main-rotor states would not increase the performance as much as estimating
only the stabilizer-bar states. This result may be explained by the slower dynamics of
the stabilizer-bar with respect to the main rotor dynamics, resulting in a predominant
effect of the stabilizer-bar on the rigid-body dynamics. The analysis was performed on
the hover model. After generating the H∞ loop shaping controller, several performance
metrics were evaluated for the following cases.

Case 1 Both rotor and stabilizer-bar states are used for feedback (i.e., the matrix F is
unchanged). The entries in F representing the feedback of these states (a1 longitu-
dinal and b1 lateral flapping of the main rotor, and a1s longitudinal and b1s lateral
flapping of the stabilizer bar) to the longitudinal and lateral input δlon and δlat are

a1 b1 a1s b1s
δlon −1.3373 −0.6915 −2.9125 −0.2555 (4.17)

δlat 0.2206 −1.7680 −0.2906 −3.0390


4.5. Scheduled inner-loop evaluation 108

Case 2 Only stabilizer-bar states are used for feedback.

a1 b1 a1s b1s
δlon 0.0 0.0 −2.9125 −0.2555 (4.18)

δlat 0.0 0.0 −0.2906 −3.0390

Case 3 Only rotor states are used for feedback.

a1 b1 a1s b1s
δlon −1.3373 −0.6915 0.0 0.0 (4.19)

δlat 0.2206 −1.7680 0.0 0.0

Case 4 Neither stabilizer-bar states nor rotor states are used for feedback.

a1 b1 a1s b1s
δlon 0.0 0.0 0.0 0.0 (4.20)

δlat 0.0 0.0 0.0 0.0

Table 4.1 shows the results. ǫ captures well the deterioration of the overall perfor-
mance. Case 2 (stabilizer-bar–state feedback only) satisfies all the specifications. Phase
margins and the ADS-33E bandwidths have opposite tendencies and increase while the
other performance metrics deteriorate and violate the bounds for Case 3 and 4. The
coupling of φ due to θ request varies only marginally. The reason for the poor stabil-
ity margin of Case 4 is due to the high bandwidth of the design; a lower bandwidth
(i.e., a decrease of the attitude gains) would restore acceptable values for ǫ and the gain
margins.
This analysis suggests also a possible way to optimize a H∞ loop-shaping controller
to tailor better its performance to multiple and conflicting specifications, which may not
be captured exhaustively by ǫ. Tischler et al. [2001] compared several controller design
techniques (including H∞ loop shaping) for the design of an aircraft lateral/directional
controller using the CONDUIT software package. CONDUIT optimizes the controllers
4.5. Scheduled inner-loop evaluation 109

Table 4.1: Comparisons among rotor- plus stabilizer-bar–state feedback, stabilizer-bar–


state feedback, rotor-state feedback, and rigid-body–state feedback.

ǫ GM (vθ ) GM (vφ ) PM (vθ ) PM (vφ )


Case 1 0.41 9.2 10.2 51.8 55.9
Case 2 0.33 6.9 6.9 54.8 59.3
Case 3 0.21 4.3 4.1 63.6 63.6
Case 4 0.09 2.6 1.6 66.8 66.6

(a) Stability margins.

∆θpk ∆φpk
ωBW a (vθ ) ω−3dB b (vθ ) ωBW a (vφ ) ω−3dB b (vφ ) ∆φ4 ∆θ4

Case 1 6.9 4.7 10.3 5.0 0.06 0.05


Case 2 7.2 3.8 10.8 3.7 0.06 0.06
Case 3 7.5 2.2 11.3 2.0 0.17 0.06
Case 4 7.9 2.1 11.8 1.2 0.22 0.07

(b) Bandwidth and decoupling.


a Bandwidth evaluated at the -135 deg point (ADS-33E bandwidth).
b Bandwidth evaluated at the -3 dB point.

to satisfy a set of user-selected specifications. For the H∞ loop shaping case, the opti-
mization is run optimizing a fixed structure of the shaping weights and recalculating
the controller at each step of the optimization. It was noted that optimizing the weights
did not map well into the design requirements chosen in the example case. This re-
sult may be expected, since every time the weights are changed, the H∞ optimization
recalculates an optimal controller to maximize the stability margin ǫ. If ǫ does not cap-
ture comprehensively the set of specified requirements, then the optimization may not
be successful. A better approach would be to optimize only the gains in the F matrix
leaving both the weights and the observer part unchanged. This would allow one to
4.6. Outer loops 110

trade-off the optimal value of ǫ obtained after the first synthesis with other specifica-
tions. Including also the weights in this optimization would allow further tuning of the
design at the expense of an increase in computational cost.

4.6 Outer loops

To allow the helicopter to track predetermined trajectories, its position over time must
be constrained. This is achieved with three outer loops on the vx , vy , and vz channels to
control x, y, and z respectively. These outer loops are not scheduled, since the scheduled
inner loop delivers a system with dynamic characteristics that do not vary widely in the
target flight envelope. The design is done using the closed inner loop at hover, and it is
identical to the one presented in Section 3.3.

4.7 Controller implementation

The scheduled-controller design uses MATLAB µ-Analysis and Synthesis Toolbox [Balas
et al., 1995]. The resulting controller is discretized with a zero order hold (ZOH) with
sampling frequency of 100 Hz. The ZOH discretization is necessary because the phys-
ical interpretation of the states of the discretized controllers must remain the same for
the scheduling. The design procedure is highly automated for changes that do not in-
volve the controller structure. The designer sets the crossover frequencies of all the
channels and a special-developed MATLAB routine generates

• the locations of the fixed operating points needed for scheduling;

• the weights W 1 and W 2 and the controllers K at each determined operating point;

• the evaluation of all the requirements;

• the outer-loop weights and controllers.

If the user judges the performance to be satisfactory, the routine continues generating
4.8. Flight tests 111

• the discretized versions of all the weights and controllers;

• the FORTRAN code that contains the controllers.

The automatically-produced FORTRAN code for the controller is exactly the same for
simulation on the non-linear model and flight tests. Before actual flight tests, both
manned and unmanned flight simulations make it possible to test further the controller
performance especially during maneuvering flight.

4.8 Flight tests

The CMU R-50 equipped with the presented controller flew ascending and descending,
forward, and backward turns, and steps at different speeds. The R-50 flew up to a speed
of 21.8 m/s (78.5 km/h) forward and 12 m/s (43.2 km/h) backward.

2
∆ v (m/s)

1.5

1
x

0.5

0
20 6
15
10 4
5 2
0 Time (s)
vx (m/s) −5 0

Figure 4.30: vx steps: non-linear simulation (solid), flight tests (dashed).

Figures 4.30, 4.31, and 4.32 show the results for step inputs in vx , vz , and ψ respec-
tively at different speeds. It is evident that the closed-loop dynamic characteristics are
4.8. Flight tests 112

−0.5
v (m/s)

−1

−1.5
z

−2

−2.5
0

5 20
10
Time (s) 10 0
−10 v (m/s)
x

Figure 4.31: vz steps: non-linear simulation (solid), flight tests (dashed).

15

10
ψ (deg)

20 10
10 5
0
v (m/s) −10 0 Time (s)
x

Figure 4.32: ψ steps: non-linear simulation (solid), flight tests (dashed).


4.8. Flight tests 113

Desired flight path (dashed), flight test (solid) Desired flight path (dashed), flight test (solid)
30
80

70 20
60

50 10
x (m)

z (m)
40
0
30

20
−10
10

0 −20
−20 −10 0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70 80
y (m) x (m)
(a) Horizontal flight path. (b) Vertical flight path.
20

θ (deg)
vz (m/s) vy (m/s) v (m/s)

10

5 0
x

0
−20
5
φ (deg)

30
20
0 10
0
−5 −10
dψ/dt (deg/s)

4
30
2 20
0 10
−2 0
−4 −10
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Time (s) Time (s)
(c) v x , vy , and vz . (d) θ, φ, and ψ̇.

δlon, δlat, δped [−0.5, 0.5], δcol [0, 1]


0.5
30
lon

0
δ

−0.5 25
0.5
20
δlat

0
ψ (rad)

−0.5 15
1
col

0.5 10
δ

0
0.5 5
ped

0 0
δ

−0.5
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Time (s) Time (sec)
(e) Controller outputs. (f) ψ.

Figure 4.33: Helicoidal turn at 10 m/s.


4.8. Flight tests 114

Desired flight path (dashed), flight test (solid) Desired flight path (dashed), flight test (solid)
160 50

140 40

120 30

100 20
x (m)

z (m)
80 10

60 0

40 −10

20 −20

0 −30
−20 0 20 40 60 80 100 120 0 50 100 150
y (m) x (m)
(a) Horizontal flight path. (b) Vertical flight path.
20

θ (deg)
vx (m/s)

15
10
0
5
0
−20
10
φ (deg)
vy (m/s)

30
20
0 10
0
−10 −10
dψ/dt (deg/s)

5
30
v (m/s)

20
0 10
0
z

−10
−5
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (s) Time (s)
(c) v x , vy , and vz . (d) θ, φ, and ψ̇.

δ ,δ ,δ [−0.5, 0.5], δ [0, 1]


lon lat ped col
0.5
30
δlon

−0.5 25
0.5
20
lat

0
ψ (rad)
δ

−0.5 15
1
col

0.5 10
δ

0
0.5 5
δped

0 0
−0.5
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (s) Time (sec)
(e) Controller outputs. (f) ψ.

Figure 4.34: Helicoidal turn at 15 m/s.


4.8. Flight tests 115

Desired flight path (dashed), flight test (solid) Desired flight path (dashed), flight test (solid)
0 30

−10
20
−20

−30
10
x (m)

z (m)
−40

−50 0

−60

−70 −10

−80
−20
−60 −50 −40 −30 −20 −10 0 10 20 −80 −70 −60 −50 −40 −30 −20 −10 0
y (m) x (m)
(a) Horizontal flight path. (b) Vertical flight path.
20

θ (deg)
vy (m/s) v (m/s)

−5 0
x

−10
−20
φ (deg)

5 10
0
0 −10
−20
−5 −30
dψ/dt (deg/s)

4
30
vz (m/s)

2 20
0 10
−2 0
−10
−4
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Time (s) Time (s)
(c) v x , vy , and vz . (d) θ, φ, and ψ̇.

δ ,δ ,δ [−0.5, 0.5], δ [0, 1]


lon lat ped col
0.5
30
δlon

−0.5 25
0.5
20
δlat

0
ψ (rad)

−0.5 15
1
col

0.5 10
δ

0
0.5 5
δped

0 0
−0.5
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90
Time (s) Time (sec)
(e) Controller outputs. (f) ψ.

Figure 4.35: Helicoidal turn at -10 m/s.


4.8. Flight tests 116

almost unchanged while traversing the speed envelope. The figures report also the re-
sults from non-linear simulations obtained with the MOSCA model augmented with
the flight-tested controller. The simulations match closely the flight-test responses, con-
firming the quality of the model across the flight envelope.
Figures 4.33, 4.34, and 4.35 show the flight of helicoidal turns at speeds of 10, 15,
and -10 m/s respectively. The turns are composed of the following trim trajectories:
(1) hover, (2) straight flight at the target speed, (3) constant yaw-rate turn, (4) climbing
flight at constant vertical velocity and constant yaw-rate, (5) constant yaw-rate turn, (6)
descending flight at constant vertical velocity and constant yaw-rate, (7) straight flight
at the target speed, (8) hover. Polynomial interpolation of requested vx , vz , and ψ̇ is
used to connect the trim trajectories and thus ensure smooth transition between any
two consecutive trims. The max requested yaw-rate is based on the max allowed bank
angle that was set at 25 deg for all the turns. Using

tan φmax
ψ̇max = g ,
vx

where g is gravity’s acceleration, gives requested yaw-rate of 26.2 deg/s for the 10 and
-10 m/s turns, and 17.5 deg/s for the 15 m/s turn. The maximum requested vertical
velocity during climbing and descending phase is 1.5 m/s. During the flights, the po-
sition outer loops are selectively closed on the inner loop to limit deviation from the
desired path.
Chapter 5
Conclusions

5.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

5.2 Suggestions for future work . . . . . . . . . . . . . . . . . . . . . . . . 117

5.1 Contributions

The main contribution of this thesis is the development of a modeling and control
framework that greatly reduces the time, cost, and both human and physical resources
needed to design high-performance control systems for robotic helicopters. In develop-
ing the framework this work achieved the following.

• The development of a modeling technique (MOSCA) that produces high-fidelity,


non-linear, real-time models suitable for flight simulation and controller design.
MOSCA needs few flight tests and ground measurements to deliver models with
high accuracy throughout the vehicle’s flight envelope.

• The first application of H∞ loop shaping to an autonomous vehicle. This tech-


nique was used to design, in a short amount of time, a complete controller for

117
5.2. Suggestions for future work 118

hover that performed successfully showing coordinated flight away from the de-
sign operating point.

• The development of an automated and systematic scheduling procedure using the


H∞ loop shaping stability margin. Given the target flight envelope, the procedure
determines which variables need to be scheduled and the minimum number of
fixed operating-point controllers required for the interpolation. It was shown that
using ǫ as a design tool results not only in robust stability, but also in a closed-loop
system that delivers excellent performance.

• The determination, with an analysis done using the observer plus state-feedback
form of the H∞ loop shaping controller, that in a helicopter with a stabilizer bar
the feedback of estimated stabilizer-bar states is more important than the feedback
of main-rotor states.

• The first successful flight of a scheduled H∞ loop shaping controller on a heli-


copter (manned or unmanned).

5.2 Suggestions for future work

There are several future research activities that could expand and capitalize on the re-
sults of this thesis.

• The application of MOSCA to different robotic aerial vehicles.

• The development and/or use of a larger set of specifications during controller


design. In particular the compilation of handling qualities for unmanned aerial
vehicles would better guide the design choices and would constitute a common
tool to evaluate different controllers synthesized with different design techniques.

• The investigation of the direct optimization of the state feedback matrix F to sat-
isfy conflicting specifications not well captured by the optimization of ǫ.
5.2. Suggestions for future work 119

• The use of model-based predictive control as an optimal reference generator (the


Reference Governor of Bemporad [1997]) for the closed inner-loop system pre-
sented in this thesis.
Bibliography

ADS-33E. Aeronautical Design Standard Performance Specification, Handling Qualities Re-


quirements for Military Rotorcraft. United States Army Aviation and Missile Command,
2000.

Amidi O., Kanade T., and Miller J.R. Vision-based autonomous helicopter research at
Carnegie Mellon Robotics Institute 1991-1997. In Heli Japan ’98, Paper No. T7-3. 1998.

Balas G.J., Doyle J.C., Glover K., Packard A., and Smith R. µ-Analysis and Synthesis
Toolbox User’s Guide. The MathWorks, Inc., second edition, 1995.

Ballin M.G. and Dalang-Secrétan M. Validation of the dynamic response of a blade-


element UH-60 simulation model in hovering flight. Journal of the American Helicopter
Society, 36(4):77–88, 1991.

Bates D.G., Gatley S.L., Postlethwaite I., and Berry A.J. Design and piloted simulation
of a robust integrated flight and propulsion controller. Journal of Guidance, Control,
and Dynamics, 23(2):269–277, 2000.

Bemporad A. Reference Governors: On-Line Set-Point Optimization Techniques for Con-


straint Fulfillment. Ph.D. thesis, Università di Firenze, 1997.

120
Bibliography 121

Bode H.W. Network Analysis and Feedback Amplifiers. D. Van Nostrand Company, Inc.,
1945.

Bruce P.D., Silva J.E.F., and Kellett M.G. Maximum likelihood identification of a rotary-
wing RPV simulation model from flight-test data. In Atmospheric Flight Mechanics
Conference, AIAA-98-4157. Boston, MA, 1998.

Doyle J.C. Analysis of feedback systems with structured uncertainties. IEE Proceedings,
129(6):242–250, 1982.

Doyle J.C., Francis B.A., and Tannenbaum A.R. Feedback Control Theory, chapter 7 and
8, pages 105–155. Macmillan Publishing Company, 1992.

Eshow M.M., Orlandi D., Bonaita G., and Barbieri S. Results of an A109 simulation val-
idation and handling qualities study. In Fourteenth European Rotorcraft Forum, Paper
No. 64. Milano, Italy, 1988.

Frost C.R., Tischler M.B., and Bielefield M. Design and test of flight control laws for
the Kaman BURRO unmanned aerial vehicle. In AIAA Atmospheric Flight Mechanics
Conference, AIAA 2000-4205. 2000.

Gavrilets V., Mettler B., and Feron E. Nonlinear model for a small-size acrobatic heli-
copter. In AIAA Guidance, Navigation, and Control Conference and Exhibit, AIAA 2001-
4333. 2001.

Glover K., Vinnicombe G., and Papageorgiou G. Guaranteed multi-loop stability mar-
gins and the gap metric. In Proceedings of the 39th IEEE Conference on Decision and
Control, pages 4084–4085. 2000.

Hall W.E. and Bryson A.E. Inclusion of rotor dynamics in controller design for heli-
copters. Journal of Aircraft, 10(4):200–206, 1973.

Heffley R.K. A compilation and analysis of helicopter handling qualities data. Technical
Report No 1087-2, Systems Technology, Inc., 1979.
Bibliography 122

Hoffmann F., Koo T., and Shakernia O. Evolutionary design of a helicopter autopilot.
In 3rd On-line World Conference on Soft Computing (WSC3). 1998.

Hong S.W. and Curtiss H.C. An analytic modeling and system identification study of
rotor/fuselage dynamics at hover. In Piloting Vertical Flight Aircraft: A Conference on
Flying Qualities and Human Factors, NASA Conference Publication 3220, pages 265–
286. 1993.

Hovakimyan N., Kim N., Calise A., Prasad J.V.R., and Corban E. Adaptive output
feedback for high-bandwidth control of an unmanned helicopter. In Conference on
Guidance, Navigation and Control, AIAA 2001-4181. 2001.

Howitt J., Howell S.E., McCallum A.T., and Brinson P.R. Experimental evaluation of
flight control system designs exploiting rotor state feedback. In Proceedings of the 57th
Forum of the American Helicopter Society. Washington D.C., 2001.

Hyde R.A. H∞ Aerospace Control Design - A VSTOL Flight Application. Advances in


Industrial Control Series. Springer-Verlag, 1995.

Johnson W. Helicopter Theory. Dover, 1994.

Keller J.D. An investigation of helicopter dynamic coupling using an analytical model.


Journal of the American Helicopter Society, 41(4):322–330, 1996.

Kim F.D., Celi R., and Tischler M.B. High-order state space simulation models of heli-
copter flight mechanics. Journal of the American Helicopter Society, 38(2):16–27, 1993.

Kim S.K. and Tilbury D.M. Mathematical modeling and experimental identification of
a model helicopter. In AIAA Modeling and Simulation Technologies Conference, AIAA-
98-4357. 1998.

Koo T., Shim D., Shakernia O., Sinopoli B., Ma Y., Ho F., and Sastry S. Hierarchical hy-
brid system design on Berkeley UAV. In The International Aerial Robotics Competition.
1998. http://citeseer.nj.nec.com/koo98hierarchical.html.
Bibliography 123

Kothare M.V., Campo P.J., Morari M., and Nett C.N. A unified framework for the study
of anti-windup designs. Automatica, 30(12):1869–1883, 1994.

La Civita M., Messner W.C., and Kanade T. Modeling of small-scale helicopters with
integrated first-principles and system-identification techniques. In Proceedings of the
58th Forum of the American Helicopter Society, volume 2, pages 2505–2516. Montreal,
Canada, 2002a.

La Civita M., Papageorgiou G., Messner W.C., and Kanade T. Design and flight testing
of a high-bandwidth H∞ loop shaping controller for a robotic helicopter. In Pro-
ceedings of the AIAA Guidance, Navigation, and Control Conference, AIAA-2002-4836.
Monterey, CA, 2002b.

La Civita M., Papageorgiou G., Messner W.C., and Kanade T. Design and flight testing
of a gain-scheduled H∞ loop shaping controller for wide-envelope flight of a robotic
helicopter. In American Control Conference. Denver, CO, 2003. (Submitted).

McFarlane D. and Glover K. A loop shaping design procedure using H∞ synthesis.


IEEE Transactions on Automatic Control, 37(6):759–769, 1992.

Mettler B.F., Tischler M.B., and Kanade T. System identification modeling of a


model-scale helicopter for the development of high-performance helicopter-based
unmanned aerial vehicles. Journal of the American Helicopter Society, 47(1):50–63, 2002.

Mettler B.F., Tischler M.B., Kanade T., and Messner W.C. Attitude control optimization
for a small-scale unmanned helicopter. In AIAA Guidance, Navigation, and Control
Conference. 2000.

MIL-F-9490D. General Specification for Flight Control Systems - Design, Installation and Test
of Piloted Aircraft. United States Air Force, 1975.

Miller R.H. A method for improving the inherent stability and control characteristics of
helicopters. Journal of Aeronautical Sciences, 17(6):363–374, 1950.
Bibliography 124

Morris J.C., van Nieuwstadt M., and Bendotti P. Identification and stabiliza-
tion of a model helicopter in hover. In American Control Conference. 1994.
http://citeseer.nj.nec.com/morris94identification.html.

Papageorgiou G. Robust Control System Design: H∞ Loop Shaping and Aerospace Applica-
tions. Ph.D. thesis, University of Cambridge, 1998.

Papageorgiou G. and Glover K. H∞ loop shaping: Why is it a sensible procedure for


designing robust flight controllers? In Proceedings of the AIAA Guidance, Navigation,
and Control Conference, AIAA 99-4272/3. 1999.

Papageorgiou G. and Glover K. Taking robust LPV control into flight on the VAAC
Harrier. In Proceedings of the 39th IEEE Conference on Decision and Control, pages 4558–
4564. 2000.

Papageorgiou G. and Hyde R.A. Analysing the stability of NDI-based flight controllers
with LPV methods. In Proceedings of the AIAA Guidance, Navigation, and Control Con-
ference, AIAA 01-4039. 2001.

Pitt D.M. and Peters D.A. Theoretical prediction of dynamic-inflow derivatives. Vertica,
5:21–34, 1981.

Prasad J., Calise A., Pei Y., and Corban J.E. Adaptive nonlinear controller synthesis and
flight test evaluation on an unmanned helicopter. In Proceedings of the IEEE Conference
on Control Applications. 1999.

Rowan T. Functional Stability Analysis of Numerical Algorithms. Ph.D. thesis, Department


of Computer Sciences, University of Texas at Austin, 1990.

Schrage D.P., Peters D.A., Prasad J.V.R., Stumpf W.F., and He C.J. Helicopter stability
and control modeling improvements and verification on two helicopters. In Four-
teenth European Rotorcraft Forum, Paper No. 77. Milano, Italy, 1988.
Bibliography 125

Sefton J.A. and Glover K. Pole/zero cancellations in the general H∞ problem with
reference to a two block design. Systems & Control Letters, 14(4):295–306, 1990.

Shim D.H., Kim H.J., and Sastry S. Control system design for rotorcraft-based un-
manned aerial vehicles using time-domain system identification. In IEEE International
Conference on Control Applications, CCA-808. Anchorage, AK, 2000.

Smerlas A., Postlethwaite I., Walker D.J., Strange M.E., Howitt J., Horton R.I., Gubbels
A.W., and Baillie S.W. Design and flight testing of an H∞ controller for the NRC Bell
205 experimental fly-by-wire helicopter. In Proceedings of the AIAA Guidance, Naviga-
tion, and Control Conference, AIAA 98-4300. 1998.

Sprague K., Gavrilets V., Dugail D., Mettler B., Martinos I., and Feron E. Design and
applications of an avionics system for a miniature acrobatic helicopter. In Digital
Avionics Systems Conference. 2001.

Stengel R.F. and Ray L.R. Stochastic robustness of linear time-invariant control systems.
IEE Transactions on Automatic Control, 36(1):82–87, 1991.

Takahashi M.D. H∞ helicopter flight control law design with and without rotor state
feedback. Journal of Guidance, Control, and Dynamics, 17(6):1245–1251, 1994a.

Takahashi M.D. Rotor-state feedback in the design of flight control laws for a hovering
helicopter. Journal of the American Helicopter Society, 39(1):50–62, 1994b.

Talbot P.D., Tinling B.E., Decker W.A., and Chen R.T.N. A mathematical model of a
single main rotor helicopter for piloted simulation. Technical Memorandum 84281,
NASA, 1982.

Tischler M.B. System identification requirements for high-bandwidth rotorcraft flight


control system design. Journal of Guidance, Control, and Dynamics, 1990.
Bibliography 126

Tischler M.B. and Cauffman M.G. Frequency-response method for rotorcraft system
identification: Flight applications to BO 105 coupled rotor/fuselage dynamics. Jour-
nal of the American Helicopter Society, 37(3):3–17, 1992.

Tischler M.B., Lee J.A., and Colbourne J.D. Optimization and comparisons of alterna-
tive flight control system design methods using a common set of handling-qualities
criteria. In Proceedings of the AIAA Guidance, Navigation, and Control Conference, AIAA-
2001-4266. Montreal, Canada, 2001.

Tischler M.B. and Tomashofski C.A. Flight test identification of SH-2G flapped-rotor
helicopter flight mechanics models. Journal of the American Helicopter Society, 47(1):18–
32, 2002.

Walker D.J. and Postlethwaite I. Advanced helicopter flight control using two-degree-
of-freedom H∞ optimization. Journal of Guidance, Control, and Dynamics, 19(2):461–
468, 1996.

Weilenmann M.F., Christen U., and Geering H.P. Robust helicopter position control at
hover. In American Control Conference, pages 2491–2495. Baltimore, 1994.

Weilenmann M.F. and Geering H.P. A test bench for rotorcraft hover control. Journal of
Guidance, Control, and Dynamics, 17:729–736, 1994.

Zhao X. and Curtiss H.C. A linearized model of helicopter dynamics including corre-
lation with flight test. In Second International Conference on Rotorcraft Basic Research,
ZHAO-1–ZHAO-14. University of Maryland, MD, 1988.

Zhou K., Doyle J.C., and Glover K. Robust and Optimal Control. Prentice Hall, first
edition, 1996.

Vous aimerez peut-être aussi