robust and adaptive attitude control of spacecraft using

105
UNLV Theses, Dissertations, Professional Papers, and Capstones 12-1-2013 Robust and Adaptive Attitude Control Of Spacecraft Using Solar Robust and Adaptive Attitude Control Of Spacecraft Using Solar Radiation Pressure Radiation Pressure Lakshmi Srinivasan University of Nevada, Las Vegas Follow this and additional works at: https://digitalscholarship.unlv.edu/thesesdissertations Part of the Controls and Control Theory Commons Repository Citation Repository Citation Srinivasan, Lakshmi, "Robust and Adaptive Attitude Control Of Spacecraft Using Solar Radiation Pressure" (2013). UNLV Theses, Dissertations, Professional Papers, and Capstones. 2031. http://dx.doi.org/10.34917/5363949 This Thesis is protected by copyright and/or related rights. It has been brought to you by Digital Scholarship@UNLV with permission from the rights-holder(s). You are free to use this Thesis in any way that is permitted by the copyright and related rights legislation that applies to your use. For other uses you need to obtain permission from the rights-holder(s) directly, unless additional rights are indicated by a Creative Commons license in the record and/ or on the work itself. This Thesis has been accepted for inclusion in UNLV Theses, Dissertations, Professional Papers, and Capstones by an authorized administrator of Digital Scholarship@UNLV. For more information, please contact [email protected].

Upload: others

Post on 05-Jun-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Robust and Adaptive Attitude Control Of Spacecraft Using

UNLV Theses, Dissertations, Professional Papers, and Capstones

12-1-2013

Robust and Adaptive Attitude Control Of Spacecraft Using Solar Robust and Adaptive Attitude Control Of Spacecraft Using Solar

Radiation Pressure Radiation Pressure

Lakshmi Srinivasan University of Nevada, Las Vegas

Follow this and additional works at: https://digitalscholarship.unlv.edu/thesesdissertations

Part of the Controls and Control Theory Commons

Repository Citation Repository Citation Srinivasan, Lakshmi, "Robust and Adaptive Attitude Control Of Spacecraft Using Solar Radiation Pressure" (2013). UNLV Theses, Dissertations, Professional Papers, and Capstones. 2031. http://dx.doi.org/10.34917/5363949

This Thesis is protected by copyright and/or related rights. It has been brought to you by Digital Scholarship@UNLV with permission from the rights-holder(s). You are free to use this Thesis in any way that is permitted by the copyright and related rights legislation that applies to your use. For other uses you need to obtain permission from the rights-holder(s) directly, unless additional rights are indicated by a Creative Commons license in the record and/or on the work itself. This Thesis has been accepted for inclusion in UNLV Theses, Dissertations, Professional Papers, and Capstones by an authorized administrator of Digital Scholarship@UNLV. For more information, please contact [email protected].

Page 2: Robust and Adaptive Attitude Control Of Spacecraft Using

ROBUST AND ADAPTIVE ATTITUDE CONTROL OF SPACECRAFT USING

SOLAR RADIATION PRESSURE

by

Lakshmi Srinivasan

Bachelor of Technology in Electrical Engineering

SASTRA University, TamilNadu, India

2008

A thesis submitted in partial fulfillment of

the requirements for the

Master of Science in Engineering – Electrical Engineering

Department of Electrical and Computer Engineering

Howard R. Hughes College of Engineering

The Graduate College

University of Nevada, Las Vegas

December 2013

Page 3: Robust and Adaptive Attitude Control Of Spacecraft Using

ii

THE GRADUATE COLLEGE

We recommend the thesis prepared under our supervision by

Lakshmi Srinivasan

entitled

Robust and Adaptive Attitude Control of Spacecraft Using Solar Radiation Pressure

is approved in partial fulfillment of the requirements for the degree of

Master of Science in Electrical Engineering

Department of Electrical Engineering and Computer Engineering

Sahjendra N. Singh, Ph.D., Committee Chair

EbrahimSaberinia,Ph.D.,Committee Member

VenkatesanMuthukumar, Ph.D., Committee Member

WoosoonYim, Ph.D., Graduate College Representative

Kathryn HausbeckKorgan, Ph.D., Interim Dean of the Graduate College

December 2013

Page 4: Robust and Adaptive Attitude Control Of Spacecraft Using

Abstract

Adaptive Attitude Control of Spacecraft Using SRP

by

Lakshmi SrinivasanDr. Sahjendra N. Singh, Examination Committee Chair

Professor of Electrical and Computer Engineering DepartmentUniversity of Nevada, Las Vegas

Satellites in orbit are expected to maintain a preset attitude either pointing

towards Earth (in case of satellites for weather) or pointing towards space for the

purpose of research and exploration. The satellite as a system though is extremely

nonlinear and the system parameters are not easily available. The goal of this

thesis is to develop robust and adaptive control laws that can be used to control

the attitude of satellites in elliptic orbits. The attitude of the satellite is controlled

by the use of Solar Radiation Pressure (SRP) on the solar panels of the satellite.

The SRP is basically a mechanical pressure caused by the photons impinging on the

solar panels. By deflecting the solar panels the area that is impinged by photons is

varied and therefore the torque on the satellite is also varied. This torque is used to

control the attitude of the satellite which is expected to be maintained at a preset

orientation. In this work, different methods will be used to control the satellite

under different conditions involving state and output feedback. For state feedback

all the states of the system are assumed to be available. In this case the states would

involve the pitch angle, angular velocity and acceleration of the satellite. However

the information on these states may not be easily available. Output feedback is when

iii

Page 5: Robust and Adaptive Attitude Control Of Spacecraft Using

the output alone of the system is used and only an estimation of other states is used.

Simulation is used to project the results of the different types of controllers.

iv

Page 6: Robust and Adaptive Attitude Control Of Spacecraft Using

Acknowledgements

I would like to take this opportunity to express my sincere gratitude to my advisor

Dr. Sahjendra N. Singh, whose advice, guidance and constant encouragement helped

me with this work. Throughout my academic career here at UNLV, he has supported

and advised me despite my shortcomings.

I would like to also thank Dr. KeumW. Lee for his patient teaching in MATLAB

programming and also for his support in my research work.

I wish to express my very warm thanks to Dr. Woosoon Yim, Dr. Ebrahim

Saberinia, Dr. Yahia Baghzouz and Dr. Venkatesan Muthukumar for monitoring

my research work and taking effort in reading my thesis and providing me with

valuable comments.

I am grateful to the Department of Electrical and Computer Engineering at the

University of Nevada, Las Vegas for providing me with not only an excellent work

environment but also a Teaching Assistantship to help me with my studies.

Finally a special thanks to my husband, my in-laws, my parents and my brother

whose constant support and encouragement enabled me to complete this work.

Lakshmi Srinivasan

University of Nevada, Las Vegas

December 2013

v

Page 7: Robust and Adaptive Attitude Control Of Spacecraft Using

Table of Contents

Abstract iii

Acknowledgements v

Table of Contents vi

List of Figures viii

Chapter 1 Introduction 1

1.1 Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2 Thesis Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

Chapter 2 Mathematical Model of Satellite Dynamics 7

Chapter 3 Robust Control of Spacecraft Using Finite Time Control 12

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

3.2 Mathematical Formulation of the Control Problem . . . . . . . . . . 12

3.3 Robust Control System . . . . . . . . . . . . . . . . . . . . . . . . . 14

3.3.1 A Nominal Nonlinear Control Law . . . . . . . . . . . . . . 15

3.3.2 Discontinuous Control Signal ur . . . . . . . . . . . . . . . . 17

3.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

Chapter 4 Adaptive Output Feedback Control of Spacecraft Using

High Gain Estimator 36

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

vi

Page 8: Robust and Adaptive Attitude Control Of Spacecraft Using

4.2 Mathematical Formulation of the Control Problem . . . . . . . . . . 36

4.3 Feedback Linearizing Control Law . . . . . . . . . . . . . . . . . . . 38

4.4 Adaptive Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

4.5 High Gain Estimator . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4.6 Simulations results . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

4.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

Chapter 5 Robust Finite Time Satellite Attitude Control Using Out-

put Feedback 57

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

5.2 Problem Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 57

5.3 Robust Control System . . . . . . . . . . . . . . . . . . . . . . . . . 59

5.4 Universal Output Feedback SISO Controller-Estimator . . . . . . . 60

5.5 Simulations results . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

5.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Chapter 6 Conclusion 82

APPENDIX 84

REFERENCES 86

Vita 92

vii

Page 9: Robust and Adaptive Attitude Control Of Spacecraft Using

List of Figures

1.1 Force on a reflector resulting from reflection photon flux [1] . . . . . . . 3

2.1 Orbital and satellite co-ordinate systems . . . . . . . . . . . . . . . . . 8

3.1 C = 5, i = 23.5, e = 0.2, φ0 = 45o; α∗ = 100o(a) response for K = 1, (b)

response for K = −1, (c) response for K = 0.5 . . . . . . . . . . . . . . 22

3.2 C = 5, i = 23.5, e = 0.2, φ0 = 45o;α∗ = 0o(a) response for K = 1, (b)

response for K = −1, (c) response for K = 0.5 . . . . . . . . . . . . . . 24

3.3 K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 1000; (a) response for C = 10,

(b) response for C = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.4 K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for C = 10,

(b) response for C = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

3.5 C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 100o; (a) response for φ0 = 90o,

(b) response for φ0 = 135o . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.6 C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 0o; response for φ0 = 135o . . . 28

3.7 C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 100o; (a) response for

sinusoidal disturbance, (b) random disturbance (c) pulse type disturbance 29

3.8 C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for

sinusoidal disturbance, (b) random disturbance (c) pulse type disturbance 30

3.9 C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 100o; (a) response for e = 0.05,

(b) response for e = 0.1, (c) response for e = 0.4 . . . . . . . . . . . . . 31

3.10 C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 0o; (a) response for e = 0.05,

(b) response for e = 0.1, (c) response for e = 0.4 . . . . . . . . . . . . . 32

viii

Page 10: Robust and Adaptive Attitude Control Of Spacecraft Using

3.11 C = 5, K = 0.5, φ0 = 45o, α∗ = 100o; (a) response for i = 30o, e = 0.1,

(b) response for i = 10o, e = 0.2 . . . . . . . . . . . . . . . . . . . . . . 33

3.12 C = 5, K = 0.5, φ0 = 45o, α∗ = 0o; (a) response for i = 30o, e = 0.1, (b)

response for i = 10o, e = 0.2 . . . . . . . . . . . . . . . . . . . . . . . . 34

4.1 C = 5, i = 23.5, e = 0.2, φ0 = 45o; α∗ = 100o(a) response for K = 1, (b)

response for K = −1, (c) response for K = 0.5 . . . . . . . . . . . . . . 43

4.2 C = 5, i = 23.5, e = 0.2, φ0 = 45o;α∗ = 0o(a) response for K = 1, (b)

response for K = −1, (c) response for K = 0.5 . . . . . . . . . . . . . . 45

4.3 K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 1000; (a) response for C = 10,

(b) response for C = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

4.4 K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for C = 10,

(b) response for C = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.5 C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 100o; (a) response for φ0 = 90o,

(b) response for φ0 = 135o . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.6 C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 0o; (a) response for φ0 = 90o,

(b) response for φ0 = 135o . . . . . . . . . . . . . . . . . . . . . . . . . 49

4.7 C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 100o; (a) response for

sinusoidal disturbance, (b) random disturbance (c) pulse type disturbance 50

4.8 C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for

sinusoidal disturbance, (b) random disturbance (c) pulse type disturbance 51

4.9 C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 100o; (a) response for e = 0.05,

(b) response for e = 0.4 . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4.10 C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 0o; (a) response for e = 0.05,

(b) response for e = 0.4 . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4.11 C = 5, K = 0.5, φ0 = 45o, α∗ = 100o; (a) response for i = 30o, e = 0.1,

(b) response for i = 10o, e = 0.2 . . . . . . . . . . . . . . . . . . . . . . 54

4.12 C = 5, K = 0.5, φ0 = 45o, α∗ = 0o; (a) response for i = 30o, e = 0.1, (b)

response for i = 10o, e = 0.2 . . . . . . . . . . . . . . . . . . . . . . . . 55

ix

Page 11: Robust and Adaptive Attitude Control Of Spacecraft Using

5.1 C = 5, i = 23.5, e = 0.2, φ0 = 45o; α∗ = 100o(a) response for K = 1, (b)

response for K = 0.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5.2 Estimates of α, α and α for α∗ = 100o(a) response for K = 1, (b)

response for K = 0.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

5.3 C = 5, i = 23.5, e = 0.2, φ0 = 45o;α∗ = 0o(a) response for K = 1, (b)

response for K = −1, (c) response for K = 0.5 . . . . . . . . . . . . . . 64

5.4 Estimates of α, α and α for α∗ = 0o(a) response for K = 1, (b) response

for K = 0.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

5.5 K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 1000; (a) response for C = 10,

(b) response for C = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

5.6 Estimates of α, α and α for α∗ = 100o(a) response for C = 10, (b)

response for C = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

5.7 K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for C = 10,

(b) response for C = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

5.8 Estimates of α, α and α for α∗ = 0o(a) response for C = 10, (b) response

for C = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

5.9 C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 100o; (a) response for φ0 = 90o,

(b) response for φ0 = 135o . . . . . . . . . . . . . . . . . . . . . . . . . 69

5.10 Estimates of α, α and α for α∗ = 100o(a) response for φ0 = 90o, (b)

response for φ0 = 135o . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

5.11 C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 0o; (a) response for φ0 = 90o,

(b) response for φ0 = 135o . . . . . . . . . . . . . . . . . . . . . . . . . 71

5.12 Estimates of α, α and α for α∗ = 0o(a) response for φ0 = 90o, (b)

response for φ0 = 135o . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

5.13 C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 100o; (a) response for

sinusoidal disturbance, (b) random disturbance (c) pulse type disturbance 72

5.14 Estimates of α, α and α for α∗ = 100o(a) response for sinusoidal distur-

bance, (b) random disturbance (c) pulse type disturbance . . . . . . . . 73

x

Page 12: Robust and Adaptive Attitude Control Of Spacecraft Using

5.15 C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for

sinusoidal disturbance, (b) random disturbance (c) pulse type disturbance 74

5.16 Estimates of α, α and α for α∗ = 0o(a) response for sinusoidal distur-

bance, (b) random disturbance (c) pulse type disturbance . . . . . . . . 74

5.17 C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 100o; (a) response for e = 0.05,

(b) response for e = 0.4 . . . . . . . . . . . . . . . . . . . . . . . . . . 75

5.18 Estimates of α, α and α for α∗ = 100o (a) response for e = 0.05, (b)

response for e = 0.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

5.19 C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 0o; (a) response for e = 0.05,

(b) response for e = 0.4 . . . . . . . . . . . . . . . . . . . . . . . . . . 77

5.20 Estimates of α, α and α for α∗ = 0o (a) response for e = 0.05, (b)

response for e = 0.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

5.21 C = 5, K = 0.5, φ0 = 45o, α∗ = 100o; (a) response for i = 30o, e = 0.1,

(b) response for i = 10o, e = 0.2 . . . . . . . . . . . . . . . . . . . . . . 78

5.22 Estimates of α, α and α for α∗ = 100o (a) response for i = 30o, e = 0.1,

(b) response for i = 10o, e = 0.2 . . . . . . . . . . . . . . . . . . . . . . 79

5.23 C = 5, K = 0.5, φ0 = 45o, α∗ = 100o; (a) response for i = 30o, e = 0.1,

(b) response for i = 10o, e = 0.2 . . . . . . . . . . . . . . . . . . . . . . 80

5.24 Estimates of α, α and α for α∗ = 0o (a) response for i = 30o, e = 0.1,

(b) response for i = 10o, e = 0.2 . . . . . . . . . . . . . . . . . . . . . . 80

xi

Page 13: Robust and Adaptive Attitude Control Of Spacecraft Using

Chapter 1

Introduction

Spacecraft attitude is the orientation of the spacecraft with respect to an external

frame of reference. All spacecraft are required to point to a particular direction.

While most of the satellites are pointed towards the Earth, there are also many that

are oriented towards the sun or other areas of interest or even change their attitude

in time to point towards another area of interest. Often, one part of the satellite

is required to point towards the Earth while the solar panels are required to point

towards the sun. In order to maintain the spacecraft attitude accurately, control

systems form an integral part of the spacecraft’s design.

There are many forces that act on a satellite in space that affect both its orbit

and attitude. Even though orbital dynamics and attitude dynamics are translational

and rotational, respectively, in nature, they are mutually coupled and have great

capacity to affect one another. Gravity is the most prominent force that affects

the dynamics of a satellite. However there are also other forces that affect the

spacecraft and while not as big in magnitude as the force of gravity due to Earth and

other celestial bodies, they are also important since they can affect the spacecraft

dynamics greatly. One of such forces is Solar Radiation Pressure (SRP).

Photons, quantization of all electromagnetic radiation, though considered to be

zero-rest mass particles, possess the properties of energy and momentum and so

exhibit the property of mass at the speed of light. When photons impinge on an

1

Page 14: Robust and Adaptive Attitude Control Of Spacecraft Using

object they exert a physical pressure due to the momentum they possess. Radiation

pressure has long been the object of study and was predicted to exist by Maxwell,

when investigating into the nature of light. It was soon proved to exist and the

idea of spaceflight using photons from the sun and other stars for propulsion was

investigated. SRP is the physical pressure experienced by the spacecraft when

impinged by solar radiation. This physical force has enough magnitude to affect

the attitude of a spacecraft and as such can be channeled to control the same.

Control surfaces are specially mounted on satellites that can be deflected to control

the torque required to maintain the orientation due to radiation that impinges them.

Indeed SRP is used as a viable means of regulating spacecraft dynamics, especially

for long life missions.

SRP is used in the following way. Since the momentum possessed by each in-

dividual photon is infinitesimal, the solar flaps must be able to intercept a large

number of photons for any effective pressure. This means that the solar flaps must

have a large surface area and also must be perfectly reflective since the pressure

experienced by the flaps will be reduced in case of absorption of light by the sur-

face. If perfect reflectivity is ensured, the momentum transferred to the flaps can be

almost double the momentum transferred by the incident photons. At the distance

of Earth from the sun, i.e 1 Astronomical Unit, only 9 newtons of force is available

per square kilometre of area. Since the force experienced is inversely proportional

to the square of the distance, the force is much reduced the farther away from the

sun the reflectors are. Fig.(1.1) shows the force experienced by the solar flaps. This

force then produces a moment about the centre of gravity of the satellite over the

distance of the flap from the centre of gravity. This causes change in orientation of

the satellite.

Many designs are available for the control system that governs the deflection of

the solar panels of the satellite so that the required attitude is maintained. The

spacecraft as a system has many nonlinearities and uncertainties associated with it.

The nature of the system demands a robust adaptive control system for an effective

2

Page 15: Robust and Adaptive Attitude Control Of Spacecraft Using

Figure 1.1: Force on a reflector resulting from reflection photon flux [1]

regulation of its attitude.

1.1 Literature review

Various control schemes have been put forward by the research community for ro-

bust and effective control of spacecraft orientation. In particular, several works

concentrate on using SRP for the same. Much analysis had been done on the effect

of the dynamics of celestial bodies [2,3]. The effects of SRP on small bodies are

discussed in [4]. Various configurations have been proposed for the utilization of

SRP. Some of them are the trailing cone system [5], weathervane type sail surfaces

[6], reflector-collector system [7], corner mirror arrays [8], solar paddles [9], grated

solar sails [10] and mirror like surfaces [11-26]. These configurations have been sug-

gested for sun-pointing satellites and gravity oriented satellites. Studies were also

made on spinning [8-15] and non-spinning satellites [16-26]. Satellite attitude con-

trol has been achieved by rotating the control surfaces about satellite body-fixed

axes [11-23] or translatory motions of single or multiple control surfaces [24-26]. A

few missions like the Mariner IV [27] employed solar vanes for passive sun pointing

3

Page 16: Robust and Adaptive Attitude Control Of Spacecraft Using

attitude and geo-stationary communication satellite OTS-2 mission of the European

Space Agency [28] used solar flaps for satellite attitude control.

The SRP control torque can thus be utilized to stabilize librational dynamics

of a satellite with a desired level of accuracy. An analysis of SRP-induced coupled

liberations of gravity stabilized geostationary communication satellites has been

presented in [31]. Several control systems have been designed in the past for single

and three-axis control using SRP. A time optimal solar pressure controller for pitch

angle has been designed [30,31].A control law for attitude control has been designed

based on Floquet theory in [32]. Also a control system using SRP for attaining ar-

bitrary inertially fixed orientation of axis symmetric spacecraft has been developed

in [19]. SRP control of spacecraft with uncertain dynamics in circular and elliptic

orbits using variable structure control (VSC) systems have been designed [33,34].

In VSC system, however, control law is discontinuous and usually requires high-gain

feedback because the switching gains are computed based on the estimated bounds

of uncertain functions. A nonlinear SRP adaptive sliding mode controller has been

designed for obtaining improved performance for satellites orbiting in circular or-

bits[36]. An adaptive backstepping design method has been used for the control

of satellite using the SRP [36]. The structure of the adaptive control signal here

is based on the certainty-equivalence principle [37]. Based on the immersion and

invariance theory, a noncertainty-equivalent adaptive solar control law has been de-

veloped for the pitch attitude control of spacecraft in circular orbits [38]. A new

control law based on the attractive manifold design method, which overcomes cer-

tain limitations of the immersion and invariance method, has been proposed in [39]

for the control of satellites in elliptic orbits using SRP.Recently SRP control law for

attitude control was achieved using L1 adaptive control in [40].

4

Page 17: Robust and Adaptive Attitude Control Of Spacecraft Using

1.2 Thesis Outline

Satellites and other spacecraft have been the source of information for space re-

search as well as research and communication on Earth. The effectiveness of the

information critically depends on the maintenance of the orientation of the space-

craft. Many schemes have been put forward for a robust and accurate control of

satellite attitude.

This thesis contributes in designing adaptive laws using various closed-loop con-

trol design methods. The implementation of control systems for maintenance of

spacecraft attitude in spite of uncertainties in parameters and presence of exter-

nal disturbance inputs is considered. The scope of this research work covers the

design of adaptive laws for the maintenance of spacecraft attitude using different

approaches. This thesis presents work on robust adaptive control using finite time

control (chapter 3) and output feedback alone (chapter 4,5), which has so far not

been explored. The mathematical model of the satellite in orbit is presented in

Chapter 2.

In chapter 3, a finite-time control law is derived and is used to control the pitch

angle of the spacecraft. The controller is fed with the pitch angle, pitch angular

velocity and pitch angle acceleration. These states are assumed to be available to

the controller and the trajectory of the pitch angle is produced based on those states.

The simulation results are analyzed for the effectiveness of the controller.

In chapter 4, a nonlinear adaptive controller that achieves feedback linearization

is developed with a high gain estimator. This controller is an ouptput feedback

controller which requires only the pitch angle of the system and uses the estimated

values of the angular velocity and acceleration. The chosen estimator is a high gain

observer which converges very quickly with the actual states of the system.

In chapter 5, the same controller as used in Chapter 3 is used but with an output

feedback system using another higher order sliding mode observer for estimating the

higher order derivatives of the pitch angle. The observer attains state estimation in

5

Page 18: Robust and Adaptive Attitude Control Of Spacecraft Using

finite time.

Simulation results are presented in each chapter and each controller is tested

with changes in parameters, eccentricity of the orbit, uncertainties and presence of

disturbances to check for robustness and performance.

6

Page 19: Robust and Adaptive Attitude Control Of Spacecraft Using

Chapter 2

Mathematical Model of Satellite

Dynamics

The mathematical model of the satellite used in chapters 3, 4, and 5 is presented in

this chapter. Though the derived control law is different for chapters 3 and 4, the

model of the satellite remains the same.

The mathematical model can be described as follows:

Fig. 2.1 shows a satellite with its center of mass S rotating in an elliptic orbit

about the Earth’s center O. The chosen inertial (XY Z), rotating orbital ( X0Y0Z0)

and body-fixed ( Xb, Yb, Zb) coordinate systems are also shown in the figure. (The

axes Z, Z0 and Zb normal to the orbital plane are not shown in the figure.) The

satellite is equipped with two highly reflective, lightweight solar control surfaces P1

and P2 for the purpose of control. These flaps are mounted along the Xb axis of the

satellite. The solar aspect angle is denoted by φ, and ω and θ are the arguments of

perigee and true anomaly, respectively. The pitch angle α is equal to λ + θ, where

λ is the angle between the body-fixed axis Xb and the local vertical axis X0.

For simplicity, only the pitch attitude control of the spacecraft using control

torque generated by the solar flaps is considered. The second-order differential

equation describing the pitch attitude dynamics is given by [44]

Izd2α

dt2=Mg +Ms +Md (2.1)

7

Page 20: Robust and Adaptive Attitude Control Of Spacecraft Using

q

w

fi

R

1P

2P

1d

2d

R

q

a

l 0X

0Y

bY

bX

Figure 2.1: Orbital and satellite co-ordinate systems

8

Page 21: Robust and Adaptive Attitude Control Of Spacecraft Using

where Ix, Iy and Iz are the moments of inertia of the satellite about the body-fixed

axes (Xb, Yb, Zb), Mg is the gravitational torque, Ms is the solar torque and Md(t)

denotes the external time-varying disturbance torque. Apparently, it is assumed

here that the roll and yaw angles of the satellite are controlled by means of additional

solar flaps and actuators so that its axis Zb remains normal to the orbit. Of course,

here the perturbations in orbit are ignored. The gravity gradient torque Mg acting

on the spacecraft is given by

Mg = − 3µ

R3(θ)(Ix − Iy) sinλ cosλ (2.2)

where R(θ) is the distance of the satellite center of mass from the Earth’s center.

The control torque produced by the solar flaps is a nonlinear function of the

rotation angles δ1 and δ2 of the two flaps, measured from the axis Xb. Because

the solar radiation forces on these control surfaces are directed along the surface

normals, only the rotation of the satellite about the axis normal to the orbital plane

is produced by the solar radiation pressure. The net solar torque Ms produced by

the control surfaces is given by [15]

Ms = C ′

sσs(φ)[sin2(α + βs(φ) + δ1)∆1 cos δ1 − sin2(α + βs(φ) + δ2)∆2 cos δ2]

.= C ′

sσsψ(α, βs, δ) (2.3)

where ∆i = sgn(sin(α + βs + δi)), i = 1, 2 and δ = (δ1, δ2)T and the nonlinear

function ψ is defined in Eq. (2.3). The functions σs and βs are

σs(φ) = 1− sin2 φ sin2 i

βs(φ) = ω − tan−1(tanφ cos(i)) (2.4)

The solar aspect angle φ varies from 0 to 2π radians in a year; and therefore, it is

a slowly varying function of θ.

The parameter C ′

s is given by

C ′

s = 2ρspAsl (2.5)

9

Page 22: Robust and Adaptive Attitude Control Of Spacecraft Using

where As is the surface area of the solar flap exposed to impinging photons, p is the

nominal SRP constant, ρs is the fraction of impinging photons specularly reflected,

and l is the distance between the center of pressure on the solar flap and the system

center of mass.

The radial distance R(θ) of the satellite from the center of the Earth and the

orbital angular velocity are given by [34,44]

R(θ) =a(1− e2)

1 + e cos θ=

µ1/3(1− e2)

Ω2/3(1 + e cos(θ))

dt=

µa(1− e2)

R2(2.6)

where e is the orbit eccentricity, a denotes the semi-major axis of the orbit, and

the mean orbital rate is Ω = (µ/a3)1/2. For the derivation of the control law as a

function of the true anomaly, θ, is preferred. Therefore, instead of the time t, here

θ is treated as an independent variable. (For simplicity in notation, the derivatives

of functions with respect to θ will be denoted by overdots.) Note that

dt= α

dt

d2α

dt2= α

(

dt

)2

+ αd2θ

dt2(2.7)

Using Eqs. (2.6) and (2.7) in Eq. (2.1), it can be shown that the pitch dynamics

can be written as [34]

(1 + e cos θ)α = −1.5K sin 2(α− θ) + 2eα sin θ + CsσsMsn(α, θ, βs(φ), δ) +Mdn(θ)

(2.8)

where

K =Ix − IyIz

;Cs =C ′

s

IzΩ2

Msn =

(

1− e2

1 + e cos θ

)3

ψ

Mdn =Md(1− e2)3

(1 + e cos θ)3IzΩ2(2.9)

10

Page 23: Robust and Adaptive Attitude Control Of Spacecraft Using

Solving for α, Eq. (2.8) gives

α = f0(α, α, θ) + Csv(α, θ, δ) (2.10)

where the nonlinear functions f0 and v are

f0(α, α, θ) = (1 + e cos θ)−1[2e sin θα− 1.5K sin 2(α− θ) +Mdn]

v(α, θ, δ) = (1 + e cos θ)−1σsMsn (2.11)

Note that the disturbance input Md is included in the nonlinear function f0. (The

argument θ denotes the dependence of f0 on Md.)

11

Page 24: Robust and Adaptive Attitude Control Of Spacecraft Using

Chapter 3

Robust Control of Spacecraft

Using Finite Time Control

3.1 Introduction

In this chapter, a state feedback control system is designed for robust finite time

control of a satellite. For the synthesis of this controller, one requires feedback of

pitch angle, pitch rate and pitch angular acceleration. These states are assumed

to be available and are used in calculating the required control input. The control

system includes a nominal nonlinear finite time continuous tracking control law and

a higher order sliding (discontinuous) control law for the compensation of uncer-

tainties in the spacecraft model. The nominal controller is designed based on the

notion of geometric homogeneity of vector fields. It is shown that in the closed loop

system, including the nominal and sliding mode control law, finite time control of

the complete state vector associated with the pitch error dynamics to the origin is

accomplished.

3.2 Mathematical Formulation of the Control Problem

For the development of the control law, the satellite orbiting in an elliptic orbit

derived in the previous chapter is considered. The pitch dynamics described in

12

Page 25: Robust and Adaptive Attitude Control Of Spacecraft Using

Chapter 2 are given by

α = (1 + e cos θ)−1[2e sin θα− 1.5K sin 2(α− θ)] + (1 + e cos θ−1)[CsσsMsn +Mdn]

(3.1)

which can be written as

α = f0(α, α, θ) + Csv(α, θ, δ) (3.2)

where the nonlinear functions f0 and v are

f0(α, α, θ) = (1 + e cos θ)−1[2e sin θα− 1.5K sin 2(α− θ) +Mdn]

v(α, θ, δ) = (1 + e cos θ)−1σs +Mdn (3.3)

Note that the disturbance input Mdn is included in the nonlinear function f0.

(The argument θ denotes the dependence of f0 on Mdn For the design of the con-

troller, it is assumed that the nonlinear function f0 as well as the solar parameter

Cs are not known.

The Eq. (3.2) describing the rotational motion of the satellite is nonaffine-

in-control variables δ1 and δ2. Therefore, it will be convenient to introduce the

derivatives of δ1 and δ2 as control inputs. Differentiating Eq. (3.2) with respect to

θ yields

d3α

dθ3= fa(α, α, θ), θ, δ +B(α, θ, δ)δ (3.4)

where

fa(α, α, θ) =∂f0(α, α, θ)

∂αα +

∂f0(α, α, θ)

∂αα +

∂f0(α, α, θ)

∂θ+ Cs

[

∂v

∂αα+

∂v

∂θ

]

B(α, θ, δ) = Cs∂v

∂δ

Note that for obtaining the expression for fa, α given in Eq. (3.2) has been

substituted for α. (Often the argument of functions are suppressed for simplicity in

13

Page 26: Robust and Adaptive Attitude Control Of Spacecraft Using

notation.) It is assumed here that the nonlinear function fa and the matrix B are

not known precisely. Let Ω1 ⊂ R4 be a region defined by

Ω1 = (α, θ, δ) : rankB(α, θ, δ) = 1

In this study, the motion of the satellite evolving in the region Ω1 will be of

interest. Note that outside the region Ω1, the input signal δ has no effect on the

pitch dynamics in Eq. (3.4).

Suppose that αr is a given reference pitch angle trajectory. The objective is to

design a robust control law such that pitch angle α converges to the reference tra-

jectory αr in a finite time, despite the presence of disturbance input and parameter

uncertainties in fa and B. (Asymptotic convergence of the tracking error is not of

interest here.)

3.3 Robust Control System

In this section, the design of a robust control system for the control of the pitch

angle is considered. Defined signals ξi, i = 1, 2, 3, as

ξ1 = α− αr; ξ2 = α− αr; ξ3 = α− αr

where α− αr is the tracking error. Then pitch dynamics given in Eq. (3.4) can

be written in a state variable form as

ξi = ξi+1; i = 1, 2

ξ3 = f(ξ, θ, δ) +B(ξ, θ, δ)u (3.5)

where ξ = (ξ1, ξ2, ξ3)T ∈ R3 and

u = δ; f(ξ, θ, δ) = fa − (d3αr/dθ3) (3.6)

In Eq. (3.6), it is assumed that the nonlinear function f and B are not known,

but some nominal values of these values of these are known. Let f ∗ and B∗ be the

14

Page 27: Robust and Adaptive Attitude Control Of Spacecraft Using

nominal values, and ∆f and ∆B be the uncertain portion of f and B, respectively.

Then the system EQ. (3.4) takes the form

ξi = ξi+1; i = 1, 2

ξ3 = f ∗(ξ, θ, δ) + ∆f(ξ, θ, δ) + (B∗(ξ, θ, δ) + ∆B(ξ, θ, δ)u (3.7)

It is assumed that B∗ has rank one in the region of interest Ω1. The objective

is to design a control law u which steers the state vector ξ to the origin despite

uncertainties in the system.

In view of Eq. (3.7), first for canceling the known function f ∗, one selects the

control signal as

u = (B∗)T ((B∗(B∗)T ))−1[−f ∗(ξ, θ, δ) + un + ur] (3.8)

where un is designed for the stabilization of the nominal system and then ur

is designed for annihilating the effect of uncertain function. (The superscript T

denotes matrix transposition.) Note that in the region Ω1, B∗(B∗)T is invertible.

Substituting Eq. (3.8) in Eq. (3.7) gives

ξ1 = ξi+1; i = 1, 2

ξ3 = ∆f(ξ, θ, δ) + un + ur +∆B(B∗)T ((B∗(B∗)T ))−1(−f ∗(ξ, θ, δ) + un + ur) (3.9)

Now the design of the nominal control signal un is considered.

3.3.1 A Nominal Nonlinear Control Law

For the satellite model without uncertainties (i. e., ∆f = 0, ∆B=0), Eq. (3.10)

yields

ξi = ξi+1; i = 1, 2

15

Page 28: Robust and Adaptive Attitude Control Of Spacecraft Using

ξ3 = un + ur (3.10)

For the class of systems described by a chain of integrators, Bhat and Bernstein

[42] have developed a nonlinear stabilizing control law. The control law un renders

the closed-loop system without uncertainties homogeneous, and achieves finite-time

stability. In this subsection, a nominal control law un is designed based on the

results of [42].

To this end, the notion of finite-time stability of a system is introduced [43].

Definition 1: Consider a system x = f(x) with f(0) = 0 where f : D → Rn

and D ⊂ Rn is an open neighborhood of the origin. The origin x = 0 is said to

be a finite-time-stable equilibrium if there exists an open neighborhood N ⊂ D of

the origin and a function Ts : N → [0,∞), called the settling time, such that the

following statements hold:

(i) Finite-time Convergence: For every nonzero initial state x0 ∈ N , the solution

x(t) remains in N for all t ∈ [0, Ts(x0)), and x(t) tends to zero, as t tends to Ts(x0).

(ii) Lyapunov stability: For every open neighborhood Uǫ of 0, there exists an

open set Uδ containing 0 such that the trajectory beginning in Uδ remains in Uǫ,

∀t ∈ [0, Ts(x0)).

The origin is said to be a globally finite-time-stable equilibrium if it is a finite-time

stable equilibrium with D = N = Rn. Now the definition of a homogeneous vector

field is introduced.

Definition 2: A vector field f(x1, ..., xn) ∈ Rn is said to be homogeneous of

degree µ ∈ R with dilation (ν1, ..., νn) ∈ ((0,∞))n if fi(pν1x1, ..., p

νnxn) = pµ+νifi(x),

for i = 1, ..., n, ∀x 6= 0, and ∀p > 0, where x = (x1, ..., xn)T and fi is the ith compo-

nent of f .

For the nominal attitude tracking error dynamics Eq. (3.10), according to [42],

16

Page 29: Robust and Adaptive Attitude Control Of Spacecraft Using

the finite-time stabilizing control signal un is chosen as

un(ξ) = −k1|ξ1|ν1sgn(ξ1)− k2|ξ2|ν2sgn(ξ2)− k3|ξ3|ν3sgn(ξ3) (3.11)

where the feedback gains ki > 0, (i = 1, 2, 3), are such that

Π(λ) = λ3 + k3λ2 + k2λ

2 + k1 (3.12)

is a Hurwitz polynomial, and the exponents νi, (i = 1, 2, 3), are chosen to satisfy

νi−1 =νiνi+1

2νi+1 − νi; i = 2, 3 (3.13)

with ν4 = 1, and ν3 = ν ∈ (1 − ǫ, 1) and ǫ ∈ (0, 1). Note that un is a continuous

function of ξ. Substituting the control law un Eq. (3.11) in Eq. (3.10) with ur = 0

gives

ξi = ξi+1, i = 1, 2

ξ3 = un = −k1|ξ1|ν1sgn(ξ1)− k2|ξ2|ν2sgn(ξ2)− k3|ξ3|ν3sgn(ξ3) (3.14)

It is easily verified that the system Eq. (3.14) is homogeneous of negative degree

µ = (ν − 1)/ν, with dilation (ν−11 , ν−1

2 , ν−13 ). It has been proven in [42] that there

exists ǫ ∈ (0, 1) such that, for every ν ∈ (1− ǫ, 1), the origin ξ = 0 of Eq. (3.14) is

finite-time stable. In fact, for the system Eq. (3.15), there exists a positive definite,

radially unbounded Lyapunov function which is used for establishing stability.

Of course, in the presence of uncertainties, this nominal control law un cannot

guarantee stability. Now for the robustness in the control system with respect to

∆f and ∆B, a discontinuous control signal ur is designed.

3.3.2 Discontinuous Control Signal ur

The design of the discontinuous control signal ur is based on a higher order sliding

mode design procedure given in [44]. The higher order sliding mode scheme provides

a control law which accomplishes finite-time stabilization of ξ1 and its first two

derivatives ξ1 and ξ1. For this purpose, it is essential to make certain assumptions

on the uncertain function ∆f and the unknown nonlinear input row vector ∆B.

17

Page 30: Robust and Adaptive Attitude Control Of Spacecraft Using

Assumption 1: There exists a function γ1(ξ) and a γ0 ∈ [0, 1) such that the

following inequalities hold:

|∆f(ξ, θ, δ) + ∆B(B∗)T ((B∗(B∗)T ))−1(un − f ∗(ξ, θ, δ)| ≤ γ1(ξ) (3.15)

|∆B(B∗)T ((B∗(B∗)T ))−1| ≤ γ0 < 1 (3.16)

Although the first inequality does not pose any restriction on the uncertain function

f , Eq. (3.16) limits the uncertain parameter ∆B. It is pointed out that the bound γ1

depends only on ξ. The reason for this is that θ and δi are the arguments of sinusoids

( sin(.) and cos(.)) in the uncertain function ∆f and f ∗; and the magnitudes of these

sinusoids cannot exceed one.

For the design, similar to [44], a sliding surface s(ξ, ξa) = 0 is chosen as

s(ξ, ξa) = ξ3 + ξa (3.17)

where the auxiliary signal ξa satisfies

ξa = −un (3.18)

It is noted that ξa is the integral of the nominal input un. It will be seen later that

introduction of the auxiliary signal ξa in the function s achieves simplification in

the derivative of s by cancelling the nominal input signal un. The derivative of s

along the solution of Eq. (3.9) can be written as

s = ∆f + ur +∆B(B∗)T ((B∗(B∗)T ))−1(−f ∗ + un + ur) (3.19)

For the derivation of the signal ur, consider a Lyapunov function

V (s) = s2/2 (3.20)

Differentiating V and using Eq. (3.19) gives

V = s[∆f + ur +∆B(B∗)T ((B∗(B∗)T ))−1(−f ∗ + un + ur)] (3.21)

18

Page 31: Robust and Adaptive Attitude Control Of Spacecraft Using

Using inequality Eq. (3.15) in Eq. (3.21) gives

V ≤ s(1+∆B(B∗)T ((B∗(B∗)T ))−1)ur+|s|.|∆f+∆B(B∗)T ((B∗(B∗)T ))−1(−f ∗(ξ, θ, δ)+un)|

≤ s(1 + ∆B(B∗)T ((B∗(B∗)T ))−1)ur + γ1(ξ)|s| (3.22)

For making the derivative of V negative, one selects ur as

ur = −G(ξ)sgn(s) (3.23)

where the gain G > 0. Substituting Eq. (3.23) in Eq. (3.22) gives

V ≤ −G(ξ)|s| − sG(ξ)∆B(B∗)T ((B∗(B∗)T ))−1sign(s) + γ1|s|

≤ −G(ξ)|s|+G(ξ)|s|.|∆B(B∗)T ((B∗(B∗)T ))−1|+ γ1(ξ)|s|

≤ −G(ξ)(1− γ0)|s|+ γ1(ξ)|s| (3.24)

In view of Eq. (3.24), one selects the gain G such that

G ≥ (1− γ0)−1[γ1(ξ) + η] (3.25)

where η > 0. For such a choice of G, Eq. (3.24) yields

V ≤ −η|s| (3.26)

Using Eq. (3.19), Eq. (3.26) gives

V ≤ −η√2√V (3.27)

This implies that s converges to zero in a finite time and the settling time satisfies

Ts(s(0)) ≤√2V (s(0))1/2(η)−1

. It is interesting to note that the subsequent motion of the system is such that s(t)

remains zero for all t ≥ Ts. On the sliding manifold, the equivalent control ueq is

determine by setting s = 0 in Eq. (3.19). Therefore ueq satisfies

[1+∆B(B∗)T ((B∗(B∗)T ))−1]ueq+∆f+∆B(B∗)T ((B∗(B∗)T ))−1(un−f ∗) = 0 (3.28)

19

Page 32: Robust and Adaptive Attitude Control Of Spacecraft Using

Substituting Eq. (3.28) in Eq. (3.9) gives

ξk = ξk+1, k = 1, 2

ξ3 = un (3.29)

Thus in closed-loop system, after a finite time, the trajectory evolves according to

system Eq. (3.14). Of course, it is already shown that the origin ξ = 0 of the system

Eq. (3.14) is finite-time stable. That is, the state vector ξ(t) converges to zero in a

finite time.

Substituting un and ur in Eq. (3.8) gives the input u = δ, which can be integrated

to obtain δ. It is pointed out that in the region of interest Ω1 the control law is well

defined. It easily follows that in the region Ω1 in which B(α, θ, δ) has rank one is

defined by(

∂v

∂δ1

)2

+

(

∂v

∂δ2

)2

6= 0 (3.30)

. Because for circular and elliptic orbits 0 ≤ e < 1, it follows from the expression

for Msn in Eq. (3.1) that in the complement Ωc1 of the region Ω1, either

σs(φ) = 1− sin2 φ sin2 i

or[

∂ψ

∂δ1

]2

+

[

∂ψ

∂δ2

]2

= 0

Of course, σs(φ) must not be zero to obtain a nonzero control torque.

Remark 1: This is unlike the sliding mode solar attitude control schemes with

linear sliding variable used in [33-35] in which although the sliding surface is reached

in a finite time, the converges of the state vector to the origin is accomplished only

as θ tends to infinity. Here the nominal control law together with the discontinuous

control law accomplishes convergence of the tracking error ξ1(t) and its first two

derivatives (ξ1, ξ1) to zero in a finite time. The motion of the system Eq. (3.9) and

Eq. (3.18) (with state variables (ξ1, ξ2, ξ3, ξa)T ∈ Ωe ⊂ R4) on the set defined by

ξ1 = 0, ξ1 = 0, ξi = 0

20

Page 33: Robust and Adaptive Attitude Control Of Spacecraft Using

is termed a third-order sliding mode.

Simulation Results This section presents the results of digital simulation. The

complete closed-loop system including the satellite model Eq. (3.2) with and without

external disturbance moment, and the control law Eqs. (3.7), (3.10) and (3.22) is

simulated for a set of values of K, Cs, eccentricity e, orbit inclination i and solar

aspect angle φ for target angles α = 0o, 100o. The solar aspect angle φ is a slowly

varying function. The function φ given by

φ(θ) = φ0 + (∂φ/∂θ)(θ − θ0)

is used here for computation, where φ0 = φ(θ0). The inclination of the orbital plane

of the geosynchronous satellite is i = 23.5o. The semi-major axis is a = 42, 241 km

and Iz is 500 kg.m2. The initial conditions of the spacecraft are chosen as θo = 0,

α(θo) = 100o, 0o and α(θo) = 0. The initial values of the flap deflections are δ1(θo) =

0o and δ2(θo) = 0o. For a limited perturbation in B, it is assumed to be of the form

B = (C∗

s + ∆Cs)∂v∂δ, where the vector function (∂v/∂δ) is known, C∗

s is the known

nominal value, and ∆Cs is the uncertain portion of Cs. The nonlinear nominal

function f ∗(ξ, θ, δ is assumed to be zero for simplicity in implementation; that is,

f(ξ, θ, δ) = ∆f . Apparently, such a choice of ∆f represents a large uncertainty

in the model. The reference pitch angle trajectory is generated by a fifth-order

reference generator given by

d5αr

dθ5= −p4

d4αr

dθ4− p3

d3αr

dθ3− p2

d2αr

dθ2− p1

dαr

dθ− p0(αr − α∗) (3.31)

where α∗ is the target pitch angle. The parameters in Eq. (3.31) are such that the

poles of the reference generator are [−2,−2,−2,−2,−2]. The initial conditions are

αr(0) = 0 for Case Ia, α = 100o for Case Ib and djαr(0)/dθj = 0, j = 1, 2...4. The

feedback gains in the nominal signal un are k1 = 29512, k2 = 2874, and k3 = 93.

Therefore, the roots of Eq. (3.15) are [-28, -31, -34]. The switching gain in Eq.

(3.24) is selected as G = 0.55. Here a constant gain G is used for simplicity in

implementation. These controller parameters have been selected by observing the

simulated responses.

21

Page 34: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−1

0

1

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−1

0

1

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5

10x 10

−5

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

δ2

δ2

δ2

δ1

δ1

δ1

(a) (b) (c)

Figure 3.1: C = 5, i = 23.5, e = 0.2, φ0 = 45o; α∗ = 100o(a) response for K = 1,(b) response for K = −1, (c) response for K = 0.5

The signum function in the control law is replaced by a saturation function

sat(s). Suppose that 0 < ǫsat ≪ 1. The function sat(s) is then defined as

sat(s) =

1ǫsat

s |s| ≤ ǫsat

1 s > ǫsat

−1 s < −ǫsat

Here ǫsat is taken as 3.5× 10−5

Case Ia. Robust Attitude Control, effect of K: K = 1, −1 and 0.5,

Cs = 5, e = 0.2, i = 23.5, φ0 = 45o, α∗ = 100o and disturbance input Md = 0

It is desired to control the pitch angle to a target value 100o. For this purpose,

one sets α∗ = 100o in Eq. (3.31). It is assumed that the actual value of the solar

parameter is Cs = 5, but C∗

s is 7. The disturbance input Md is assumed to be

zero. The eccentricity of the orbit is 0.2. The initial value of the solar aspect angle

is assumed to be φ0 = 45o. The closed-loop responses for the spacecraft model

Eq. (3.2) with δ in Eq. (3.1) for the values of K = 1, K = −1, K = 0.5 are

22

Page 35: Robust and Adaptive Attitude Control Of Spacecraft Using

obtained. It is pointed out that K = 1 refers to a dumbbell satellite aligned with

the local vertical while K = −1 refers to a dumbbell satellite aligned with the local

horizontal. Of course, K = −1 refers to an unstable gravity gradient configuration.

Note that because f ∗ = 0, the control law is not a function of the K-dependent

nonlinear function in the pitch dynamics. The selected responses for K = 1 (in

the left column) K = −1 (in the middle column) and K = 0.5 (in right column)

are shown in Fig. 3.1. It is observed that the pitch angle is smoothly regulated to

100o for all values of K in less than one orbit time. The maximum values of the

control surface deflections are about (25, 22) (deg) for K = 1, (27, 28) (deg), for

K = −1 and (17,13) (deg) for K = 0.5. It is observed that for K = −1 (unstable

gravity gradient configuration) larger control magnitude is required. In the steady-

state, the waveforms of the control plate deflections are oscillatory, and δ1 and δ2

are almost 180o out of phase. These oscillations in the solar flap deflections are

required to counter the periodic gravity gradient torqueMg so that the attitude can

be maintained at the target value α∗ = 100o. The SRP-dependent control signal

Csv (appearing in the α-dynamics Eq. (3.2)) has been computed using the signal

v from Eq. (3.3). Fig. (3.1) shows the waveform of the control signal Csv. Its

maximum value remains within 2(rad−1). Because δi, i = 1, 2, exhibit oscillations,

the control signal Csv has also oscillatory pattern.

Case Ib. Robust Attitude Control, effect of K: K = 1,−1,0.5 Cs = 5,

e = 0.2, i = 23.5, φ0 = 45o, α∗ = 0o and disturbance input Md = 0

It is desired to control the pitch angle to a target value 0o. For this purpose,

one sets α∗ = 0o in Eq. (3.31). It is assumed that the actual value of the solar

parameter is Cs = 5, but C∗

s is 7. The disturbance input Md is assumed to be

zero. The eccentricity of the orbit is 0.2. The initial value of the solar aspect angle

is assumed to be φ0 = 45o. The closed-loop responses for the spacecraft model

Eq. (3.2) with δ in Eq. (3.1) for the values of K = 1, K = −1, K = 0.5 are

obtained. It is pointed out that K = 1 refers to a dumbbell satellite aligned with

the local vertical while K = −1 refers to a dumbbell satellite aligned with the local

23

Page 36: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

4

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ2

δ1δ

1

(a) (b) (c)

δ2

δ2

δ1

Figure 3.2: C = 5, i = 23.5, e = 0.2, φ0 = 45o;α∗ = 0o(a) response for K = 1, (b)response for K = −1, (c) response for K = 0.5

horizontal. Of course, K = −1 refers to an unstable gravity gradient configuration.

Note that because f ∗ = 0, the control law is not a function of the K-dependent

nonlinear function in the pitch dynamics. The selected responses for K = 1 (in

the left column) K = −1 (in the middle column) and K = 0.5 (in right column)

are shown in Fig. 3.2. It is observed that the pitch angle is smoothly regulated

to 0o for all values of K in less than one orbit time. The maximum values of the

control surface deflections are about (27, 28) (deg) for K = 1, (33, 21) (deg), for

K = −1 and (24, 23) (deg) for K = 0.5. It is observed that for K = −1 (unstable

gravity gradient configuration) larger control magnitude is required. In the steady-

state, the waveforms of the control plate deflections are oscillatory, and δ1 and δ2

are almost 180o out of phase. These oscillations in the solar flap deflections are

required to counter the periodic gravity gradient torqueMg so that the attitude can

be maintained at the target value α∗ = 0o. The SRP-dependent control signal Csv

(appearing in the α-dynamics Eq. (3.2)) has been computed using the signal v from

24

Page 37: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−5

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−1

0

1

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

(a) (b)

δ2 δ

2

δ1

δ1

Figure 3.3: K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 1000; (a) response forC = 10, (b) response for C = 4

Eq. (3.3). Fig. (3.2) shows the waveform of the control signal Csv. Its maximum

value remains within 2(rad−1). Because δi, i = 1, 2, exhibit oscillations, the control

signal Csv has also oscillatory pattern.

Case IIa. Robust attitude control, effect of Cs: K = 0.5, Cs = 10 and

4, e = 0.2, i = 23.5o, φ0 = 45o, α∗ = 100o and Md = 0

Now the effect of uncertainty in the control input gain (solar parameter) Cs

on the performance of the controller is examined. For this purpose, simulation is

done for the satellite model with actual values of Cs = 10 and Cs = 4. But the

nominal value C∗

s of Cs for the controller design is assumed to be 7. As such one

has uncertainty of −30% and +75% in Cs, respectively. The value of K is 0.5,

e is 0.2 and φ0 = 45 (deg). The controller parameters of Case Ia are retained.

Selected responses are plotted in Fig. 3.3 for Cs = 10 (left column) and Cs = 4

(right column), respectively. One observes the convergence of the pitch angle to

the target value smoothly. The maximum values of the control surface deflections

25

Page 38: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−1

0

1

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−10

0

10

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−4

−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ2

δ1

δ2

δ1

(a) (b)

Figure 3.4: K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for C = 10,(b) response for C = 4

are about(12, 8) (deg) for Cs = 10 and (18, 16) (deg) for Cs = 4. As expected the

choice of larger Cs gives smaller solar flap deflection because control input matrix

has larger magnitude. The solar flap deflections are periodic in the steady-state

similar to Case Ia. Similar to Fig. 3.1, oscillatory pattern of the control signal Csv

is observed in Fig. 3.3.

Case IIb. Robust attitude control, effect of Cs: K = 0.5, Cs = 10 and

4, e = 0.2, i = 23.5o, φ0 = 45o, α∗ = 0o and Md = 0

Now the effect of uncertainty in the control input gain (solar parameter) Cs

on the performance of the controller is examined. For this purpose, simulation is

done for the satellite model with actual values of Cs = 10 and Cs = 4. But the

nominal value C∗

s of Cs for the controller design is assumed to be 7. As such one

has uncertainty of −30% and +75% in Cs, respectively. The value of K is 0.5,

e is 0.2 and φ0 = 45 (deg). The controller parameters of Case Ib are retained.

Selected responses are plotted in Fig. 3.3 for Cs = 10 (left column) and Cs = 4

26

Page 39: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−200

−100

0

100

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

(a) (b)

δ2

δ1δ

2

δ1

Figure 3.5: C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 100o; (a) response for φ0 = 90o,(b) response for φ0 = 135o

(right column), respectively. One observes the convergence of the pitch angle to

the target value smoothly. The maximum values of the control surface deflections

are about(15, 14) (deg) for Cs = 10 and (27, 27) (deg) for Cs = 4. As expected the

choice of larger Cs gives smaller solar flap deflection because control input matrix

has larger magnitude. The solar flap deflections are periodic in the steady-state

similar to Case Ib. Similar to Fig. 3.2, oscillatory pattern of the control signal Csv

is observed in Fig. 3.4.

Case IIIa. Robust attitude control, effect of φ0: K = 0.5, Cs = 5,

e = 0.2, φ0 = 90o and 135o, α∗ = 100o and Md = 0

Now the performance of the SRP control system for different known values of

the solar aspect angle φ is examined. Simulation is done using the initial value

φ0 = 90 (deg) or φ0 = 135 (deg) of the solar aspect angle. The value of K and Cs

are 0.5 and 5, respectively and the eccentricity e is 0.2. The responses are shown in

Fig. 3.5 for φ0 = 90o (left column) and for φ0 = 135o (right column), respectively.

27

Page 40: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 20

50

100

Orbits

α [d

eg]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ2

δ1

Figure 3.6: C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 0o; response for φ0 = 135o

Despite different initial values of φ0, smooth control of the pitch angle is observed.

The maximum values of the control surface deflections are about (31, 30) (deg) for

φ0 = 90o and (110, 45) (deg) for φ0 = 135o. But the response time is about one orbit

time for both values of φ0. Fig. 3.5 shows the oscillatory waveform of the control

signal Csv.

Case IIIb. Robust attitude control, effect of φ0: K = 0.5, Cs = 5,

e = 0.2, φ0 = 90o and 135o, α∗ = 0o and Md = 0

Now the performance of the SRP control system for different known values of

the solar aspect angle φ is examined. Simulation is done using the initial value

φ0 = 135 (deg) of the solar aspect angle. The value of K and Cs are 0.5 and 5,

respectively and the eccentricity e is 0.2. The responses are shown in Fig. 3.6 for

φ0 = 135o. Despite different initial value of φ0, smooth control of the pitch angle is

observed. The maximum values of the control surface deflections are about (15, 15)

(deg) for φ0 = 135o. The response time is about one orbit time for both values of

φ0. Fig. 3.6 shows the oscillatory waveform of the control signal Csv.

28

Page 41: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbitsα−

α r [deg

]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−6

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−10

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

1

2x 10

−6

Orbits

Dis

turb

ance

Md

δ2 δ

2

δ1

δ1

δ1

(a) (b) (c)

Figure 3.7: C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 100o; (a) response forsinusoidal disturbance, (b) random disturbance (c) pulse type disturbance

Case IVa. Robust attitude control despite sinusoidal, random and

pulse disturbance input Md: K = 0.5, Cs = 5, e = 0.2, i = 23.5o, φ0 = 45o,

α∗ = 100o

Simulation is done to examine the performance of the adaptive controller in the

presence (i) sinusoidal, (ii) random and (iii) pulse type disturbance inputs, shown in

the left, center, and right column in Fig 3.7, respectively. The random disturbance

is generated by passing a white noise with unit variance through a transfer function

F (s) = 5× 10−10/(s+ 5). The controller of Case Ia is retained. Selected responses

are shown in Fig. 3.7. It is observed that the controller achieves the regulation

of the pitch angle to the target value in the presence of each disturbance input.

In the steady-state, it is observed that flap deflection is a periodic function in the

presence of sinusoidal disturbance (Fig. 3.7, left column). The maximum value

of control surface deflection is about (16, 13) (deg). The control signal Csv also

exhibits periodic oscillations. The maximum values of the control surface deflections

29

Page 42: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbitsα−

α r [deg

]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−6

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−10

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

1

2x 10

−6

Orbits

Dis

turb

ance

Md

(a) (c)(b)

δ2δ

2

δ1

δ1

δ1

Figure 3.8: C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response forsinusoidal disturbance, (b) random disturbance (c) pulse type disturbance

are about (16, 13) (deg) for random (center column) and (16, 13) for the pulse-type

disturbance (right column), respectively. The response time remains of similar order

as in the previous cases.

Case IVb. Robust attitude control despite sinusoidal, random and

pulse disturbance input Md: K = 0.5, Cs = 5, e = 0.2, i = 23.5o, φ0 = 45o,

α∗ = 0o

Simulation is done to examine the performance of the adaptive controller in the

presence (i) sinusoidal, (ii) random and (iii) pulse type disturbance inputs, shown in

the left, center, and right column in Fig 3.8, respectively. The random disturbance

is generated by passing a white noise with unit variance through a transfer function

F (s) = 5× 10−10/(s+ 5). The controller of Case Ib is retained. Selected responses

are shown in Fig. 3.8. It is observed that the controller achieves the regulation

of the pitch angle to the target value in the presence of each disturbance input.

In the steady-state, it is observed that flap deflection is a periodic function in the

30

Page 43: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−5

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−5

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

δ1

δ2

δ1

δ2δ

2

δ1

(a) (b) (c)

Figure 3.9: C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 100o; (a) response for e = 0.05,(b) response for e = 0.1, (c) response for e = 0.4

presence of sinusoidal disturbance (Fig. 3.8, left column). The maximum value

of control surface deflection is about (28, 23) (deg). The control signal Csv also

exhibits periodic oscillations. The maximum values of the control surface deflections

are about (28, 23) (deg) for random (center column) and (28, 23) for the pulse-type

disturbance (right column), respectively. The response time remains of similar order

as in the previous cases.

Case Va. Robust attitude control, effect of eccentricity: K = 0.5,

Cs = 5, e = 0.05 and 0.4, i = 23.5o, φ0 = 45o, α∗ = 100o disturbance Md = 0

Now the performance of the controller for different values of the eccentricity of

the orbit is examined. Simulation is done for e = 0.05, e = 0.1 and e = 0.4. Note

that similar to Case (I-IV)a, it is assumed that e and i are known. The controller

parameters of Case Ia are retained. The responses are shown in Fig. 3.9(a) for

e = 0.05, Fig. 3.9(b) for e = 0.1 and Fig. 3.9(c) for e = 0.4. Fig. 3.9 shows

smooth regulation of the pitch angle to 100 (deg). It is observed that compared to

31

Page 44: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

δ1

δ1 δ

1

δ2

δ2δ

2

(a) (b) (c)

Figure 3.10: C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 0o; (a) response for e = 0.05,(b) response for e = 0.1, (c) response for e = 0.4

the responses for e = 0.05, larger eccentricity of the orbit causes presence of higher

harmonic terms in the solar flap rotation angle waveforms. The response time is

almost of similar order as in Cases Ia and IIa. The maximum values of the control

surface deflections are about (16, 12) (deg) for e = 0.05, (16, 12) (deg) for e = 0.1

and (32, 25) (deg) for e = 0.4, respectively. It is observed that the peaks in δi and

the control signal Csv are larger for the control of spacecraft in orbits of higher

eccentricity.

Case Vb. Robust attitude control, effect of eccentricity: K = 0.5,

Cs = 5, e = 0.05 and 0.4, i = 23.5o, φ0 = 45o, α∗ = 100o disturbance Md = 0

Now the performance of the controller for different values of the eccentricity of

the orbit is examined. Simulation is done for e = 0.05, e = 0.1 and e = 0.4. Note

that similar to Case (I-IV)b, it is assumed that e and i are known. The controller

parameters of Case Ia are retained. The responses are shown in Fig. 3.10(a) for

e = 0.05, Fig. 3.10(b) for e = 0.1 and Fig. 3.10(c) for e = 0.4. Fig. 3.10 shows

smooth regulation of the pitch angle to 0 (deg). It is observed that compared to

32

Page 45: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5

10x 10

−5

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5

10x 10

−5

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

(a) (b)

δ2 δ

2

δ1δ

1

Figure 3.11: C = 5, K = 0.5, φ0 = 45o, α∗ = 100o; (a) response for i = 30o, e = 0.1,(b) response for i = 10o, e = 0.2

the responses for e = 0.05, larger eccentricity of the orbit causes presence of higher

harmonic terms in the solar flap rotation angle waveforms. The response time is

almost of similar order as in Cases Ib and IIb. The maximum values of the control

surface deflections are about (26, 24) (deg) for e = 0.05, (25, 23) (deg) for e = 0.1

and (30, 28) (deg) for e = 0.4. It is observed that the peaks in δi and the control

signal Csv are larger for the control of spacecraft in orbits of higher eccentricity.

Case VIa. Robust attitude control, effect of e and i : K = 0.5, Cs = 5,

α∗ = 100o, φ0 = 45o, Md = 0

Simulation is done for two sets of values (e, i) = (0.1, 30o) and (0.2, 10o). Of

course, it is assumed that these values of (e, i) are known to the designer. The

closed-loop responses for (e, i) = (0.1, 30o) and (e, i) = (0.2, 10o) are shown in

Fig. 3.11(a) (left column) and Fig. 3.11(b) (right column), respectively. It is seen

that smooth regulation of the pitch angle to 1000 is accomplished for each (e, i).

The maximum values of the control surface deflections are about (17, 13) (deg) for

(e, i) = (0.1, 30o) and (15, 12) (deg) for (e, i) = (0.2, 10o), respectively. It is seen

33

Page 46: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ2

δ1 δ

1

δ2

(a) (b)

Figure 3.12: C = 5, K = 0.5, φ0 = 45o, α∗ = 0o; (a) response for i = 30o, e = 0.1,(b) response for i = 10o, e = 0.2

that the waveform of the control signal Csv is periodic in the steady-state.

Case VIb. Robust attitude control, effect of e and i : K = 0.5, Cs = 5,

α∗ = 0o, φ0 = 45o, Md = 0

Simulation is done for two sets of values (e, i) = (0.1, 30o) and (0.2, 10o). Of

course, it is assumed that these values of (e, i) are known to the designer. The closed-

loop responses for (e, i) = (0.1, 30o) and (e, i) = (0.2, 10o) are shown in Fig. 3.12(a)

(left column) and Fig. 3.12(b) (right column), respectively. It is seen that smooth

regulation of the pitch angle to 00 is accomplished for each (e, i). The maximum

values of the control surface deflections are about (26, 24) (deg) for (e, i) = (0.1, 30o)

and (22, 21) (deg) for (e, i) = (0.2, 10o), respectively. It is seen that the waveform

of the control signal Csv is periodic in the steady-state.

Extensive simulation has been performed for several values of the solar aspect

angle, the eccentricity e of the orbit, the orbit inclination i, and the model param-

eters K and Cs. These results show that the designed control law accomplishes

34

Page 47: Robust and Adaptive Attitude Control Of Spacecraft Using

robust regulation of the pitch angle trajectory, even in the presence of disturbance

input.

3.4 Conclusions

The design of a robust finite-time control system for the pitch angle control of

spacecraft with uncertain dynamics in elliptic orbits using SRP was considered.

The parameters of the nonaffine-in-control spacecraft model were assumed to be

unknown, and external disturbance input was assumed to be acting on the satel-

lite. A robust control law was designed for the tracking of reference pitch angle

trajectory in a finite time. The control system included a nominal nonlinear con-

trol designed for the finite time control of the satellite model without uncertainties.

Then a third-order sliding mode scheme was used to design a discontinuous control

signal for robustification. It was shown that the composite control system includ-

ing the nominal and discontinuous signals, the pitch angle tracking and its first two

derivatives converged to zero in a finite time. In the closed-loop system precise pitch

attitude control was accomplished, despite uncertainties and disturbance input.

35

Page 48: Robust and Adaptive Attitude Control Of Spacecraft Using

Chapter 4

Adaptive Output Feedback

Control of Spacecraft Using High

Gain Estimator

4.1 Introduction

In the previous chapter, a state variable feedback control system was designed for

the robust finite-time control of a satellite. For the synthesis of this controller one

requires feedback of the pitch angle, pitch rate and pitch angular acceleration. In

a practical case, it is of interest to design a controller which requires fewer sen-

sors for measurement. In this chapter, a new nonlinear adaptive controller is de-

signed.Interestingly, this new controller is implemented using only feedback of the

pitch angle and unlike the controller of the previous chapter, the measurement of

the first and second derivatives of the pitch angle are avoided.

4.2 Mathematical Formulation of the Control Problem

For the development of the control law, the satellite orbiting in an elliptic orbit

derived in Chapter 2 is considered. The dynamics of the satellite is described by

36

Page 49: Robust and Adaptive Attitude Control Of Spacecraft Using

α = (1+e cos θ)−1[−1.5K sin 2(α−θ)+2eα sin θ+CsσsMsn(α, θ, βs(φ), δ)+Mdn(θ)]

(4.1)

Differentiating Eq. 4.1

...α =

[

3e sin θ

1 + e cos θα

]

−K[

3 cos[2(α− θ)].(α− 1)

1 + e cos θ

]

+Cs(1− e2)3

(1 + e cos θ)4Q1d+

[

∂Q

∂δ1,∂Q

∂δ2

]

u

(4.2)

≡ f0(α, α, α, θ) +K.f1(α, α, θ) + Cf2(α, θ, β0, δ1, δ2, α, σ) + CsBsu (4.3)

where

Q ≡ σM

Q1d =

3e sin θ

1 + e cos θQ+

∂Q

∂αα +

∂Q

∂β0

∂β0∂φ

∂φ

∂θ+∂Q

∂σ

∂σ

∂φ

∂φ

∂θ

f0 =3e sin θ

1 + e cos θα +

2e cos θ

1 + e cos θα, f1 = −

[

3 cos[2(α− θ)] · (α′ − 1)

1 + e cos θ

]

u = δ ∈ R2, and Bs =(1−e2)3

(1+e cos θ)4Q1d + [ ∂Q

∂δ1, ∂Q∂δ2

]

Defing a state vectorx = (α, α, α)T ∈ Rn, one writes (3) in a state variable form

as

x =

x2

x3

fa(x, θ, δ)

+G(x1, θ, δ)u (4.4)

, f(x, θ, δ) +G(x1, θ, δ)u

y = h(x) = x1 = α

where y is the controlled output variable,

f(x) =

x2

x3

fa(x, θ, δ)

, G(x1, θ, δ) =

0

0

CsBs(x1, θ, δ)

(4.5)

37

Page 50: Robust and Adaptive Attitude Control Of Spacecraft Using

where

fa(x, θ, δ) = f0(θ, α, α) +Kf1(α, α, θ) + Csf2(θ, α, α, β0, σ, δ)

If f and Bs are known, one can choose a control law. It is assumed that the

function fa(x, θ, δ) is not known and moreover the solar parameter Cs is unknown.

The objective is to design a control law such that the pitch angle tracks any given

reference pitch angle trajectory αr(t) despite the uncertainties in the nonlinear func-

tion fa(x, θ, δ) and the parameter Cs. Moreover we are interested in synthesizing

the control law using only the pitch angle and the solar plate deflections δ1 and δ2.

4.3 Feedback Linearizing Control Law

In this subsection, a feedback linearizing control law is developed assuming that

the system (Eq.4) is completely known. For the development of the control law,

the pitch angle α is differentiated successively till the control input δ appears. Of

course, it follows from (Eq.4), that

...α = fa(x, θ, δ) + CsBs(x, θ, δ)u (4.6)

We are interested in the region Ω of the state space in which rankB(x, θ, δ) = 1.

Under the assumption that fa and CsBs are known, a feedback linearizing control

law is chosen as

δ = C−1s BT

s (BsBTs )

−1[αr − fa − p3α− p2α− p1α− p0xs] (4.7)

xs = αr

where α = α−αr is the tracking error. Control law Eq. (4.8) was integral feedback

of the tracking error. Substituting the control law Eq. (4.8) in Eq. (4.7) yields

α = −p3 ¨α− p2 ˙α− p1α− p0xs (4.8)

Differentiating Eq. (4.9) gives

...α + p3

...α + p2 ¨α + p1 ˙α + p0α = 0 (4.9)

38

Page 51: Robust and Adaptive Attitude Control Of Spacecraft Using

The design parameters pi, i = 0, 1, 2, 3 are chosen such that the characteristic poly-

nomial Π(s)

Π(s) = s4 + p3s3 + p2s

2 + p1s+ p0 (4.10)

is Hurwitz for such a choice of feedback gain pi, it follows from Eq. (4.10) that α

converges to zero as t→ ∞ and the pitch angle α asymptotically tracks the reference

trajectory αr(t). Of course, the control law Eq. (4.8) cannot be implemented

because the nonlinear function fa(x, θ, δ) and the solar parameter Cs are unknown.

Moreover, for the synthesis of Eq. (4.8), α and α are required which are not available

for feedback. In the next section, an adaptive control law is derived.

4.4 Adaptive Law

For the derivation of control law, the unknown nonlinear function fa(x, θ, δ) and the

unknown parameter C are decomposed as

fa = f ∗

a +∆fa

Cs = C∗

s +∆Cs

(4.11)

where functions with ’*’ are nominal values and unknown part is denoted with

∆(.) Then one can write Eq.(4.7) as

d3α

dθ3= −...

α r + f ∗

a +∆fa + (C∗

s +∆Cs)Bs(x, θ, δ)u (4.12)

Because the vector function Bs is known, consider a new input

va = Bs(x, θ, δ)u, va ∈ R (4.13)

Define a lumped nonlinear function η ∈ R as

η = ∆fa +∆Csva (4.14)

Note unknown function η includes all the unknown functions. Then one can write

Eq.(4.13) asd3α

dθ3= − ...

αr + f ∗

a + η + C∗

s va (4.15)

39

Page 52: Robust and Adaptive Attitude Control Of Spacecraft Using

Let z1, z2, z3 and z4 be the estimates of α, ˙α, ¨α and η respectively and suppose

that these estimated values are available. Then, in view of Eq. (4.16), using the

estimates of derivatives of α and η, one chooses a feedback linearizing control law

as

va = C∗−1s [

...α − f ∗

a − z4 − p3z3 − p2z2 + p1z1 + p0xs] (4.16)

Of course, α and xs are known. Using Eq. (4.7) for va, now δ = ucan be obtained

as

u = BTs (BsB

Ts )

−1va (4.17)

Note that if the estimation errors (α − z1), ( ˙α − z2), ( ¨α − z3) and (η − z4) are

zero, then the control law Eq. (4.17) and Eq. (4.18) reduces to the exact feedback

linearizing control law Eq. (4.8) for the deterministic system.

4.5 High Gain Estimator

In this subsection, the design of an estimator is considered. The structure of the

estimator is based on the results of [47] The nonlinear function η is treated as a

state variable. Differentiating η gives

η =d

dθ[∆fa +∆Csva] (4.18)

Let fη , η Note that u can be substituted using Eq. (4.18). Note that the nonlinear

function fη is not known. For the derivation of the estimator consider a set of

equations

d

dt

α

˙α

¨α

η

=

˙α

¨α

f ∗

a − ...α r + η + C∗

(4.19)

For constructing the estimates (z1, z2, z3, z4) of (α,˙α, α, η) a high gain estimator

is considered. The advantage of this observer is that the estimation error converges

40

Page 53: Robust and Adaptive Attitude Control Of Spacecraft Using

to zero in a short period. In view of (Eq.20), the observer is selected as

z1 = z2 + ǫ−1d1(α− z1)

z2 = z3 + ǫ−2d2(α− z1)

z3 = −...α r + f ∗ +Bδ + z4 + ǫ−3d3(α− z1)

z4 = ǫ−4d4(α− z1)

(4.20)

where di, i = 1, 2, 3, 4 are positive numbers and ǫ > 0 is a small parameter. Note

that in the observer, the measured error signal (α−z1) is feedback. The parameters

di are selected so that

s4 + d1s3 + d2s

2 + d3s+ d4 = 0 (4.21)

is stable.

Defining e1 = α− z1, e2 = ˙α− z2, e3 =¨α− z3 and e4 = η− z4 = η. Subtracting

(Eq.22) from (Eq.21), one obtains the dynamics of the estimation error

e1 = e2 + ǫ−1d1e1

e2 = e3 + ǫ−2d2e1

e3 = e4 + ǫ−3d3e1

e4 = ǫ−4d4e1

(4.22)

Note that Eq. (4.24) describing the state estimation error dynamics is solely used

for the purpose of analysis and it is not used for the control law synthesis. Introduce

a change of variables as (i = 1, 2, 3, 4)

ξi = eiǫi−4 (4.23)

Using the definition of ξi (Eq.25) gives

ǫξ = A0ξ + (0, 0, 1)T ǫfη (4.24)

41

Page 54: Robust and Adaptive Attitude Control Of Spacecraft Using

where ξ = (ξ1, ξ2, ξ3, ξ4)T ∈ R4 and

A0 =

−d1 1 0 0

−d1 0 1 0

−d1 0 0 1

−d1 0 0 0

(4.25)

(Eq.26) is in a singularly perturbed form. It has been shown in [48] that for

sufficiently small ǫ, the error ξi converges to zero in a short time, because A0 is a

Hurwitz matrix. For convergence analysis, one may follow the steps in the derivation

of [49] and therefore it is not repeated here.

4.6 Simulations results

This section presents the results of digital simulation. The complete closed-loop

system including the satellite model Eq. (4.2) with and without external disturbance

moment, and the control law is simulated for a set of values of K, Cs, eccentricity

e, orbit inclination i and solar aspect angle φ. The solar aspect angle φ is a slowly

varying function. The function φ given by

φ(θ) = φ0 + (∂φ/∂θ)(θ − θ0)

is used here for computation, where φ0 = φ(θ0). The inclination of the orbital plane

of the geosynchronous satellite is i = 23.5o. The semi-major axis is a = 42, 241

km and Iz is 500 kg.m2. The initial conditions of the spacecraft are chosen as

θo = 0, α(θo) = 100o, 0o and α(θo) = 0. The initial values of the flap deflections

are δ1(θo) = 0o and δ2(θo) = 0o. For a limited perturbation in Bs, it is assumed

to be of the form Bs = (C∗

s + ∆Cs)∂v∂δ, where the vector function (∂v/∂δ) where

v(α, θ, δ) = (1 + e cos θ)−1σsMsnis known, C∗

s is the known nominal value, and ∆Cs

is the uncertain portion of Cs. The nonlinear nominal function f ∗(ξ, θ, δ is assumed

to be zero for simplicity in implementation; that is, f(ξ, θ, δ) = ∆f . Apparently,

42

Page 55: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

δ1

δ2

δ1δ

2

δ1

δ2

(a) (c)(b)

Figure 4.1: C = 5, i = 23.5, e = 0.2, φ0 = 45o; α∗ = 100o(a) response for K = 1,(b) response for K = −1, (c) response for K = 0.5

such a choice of ∆f represents a large uncertainty in the model. The reference pitch

angle trajectory is generated by a fifth-order reference generator given by

d4αr

dθ5= −p3

d3αr

dθ3− p2

d2αr

dθ2− p1

dαr

dθ− p0(αr − α∗) (4.26)

where α∗ is the target pitch angle. The parameters in Eq. (39) are such that the

poles of the reference generator are [−2.5 − 3 − 4 − 3.5]. The initial conditions

are αr(0) = 0o, 100o and djαr(0)/dθj = 0, j = 1, 2...4. The feedback gains in the

estimator are d1 = 7, d2 = 17.75,, d3 = 19.25 and d4 = 7.5. Therefore, the roots

of Eq. (4.22) are [-1.5 -2.5 -2 -1] and ǫ = 0.005. These controller parameters have

been selected by observing the simulated responses.

Case Ia. Robust Attitude Control, effect of K: K = 1 and −1, Cs = 5,

e = 0.2, i = 23.5, φ0 = 45o, α∗ = 100o and disturbance input Md = 0

It is desired to control the pitch angle to a target value 100o. For this purpose,

one sets target angle α∗ = 100o. It is assumed that the actual value of the solar

parameter is Cs = 5, but C∗

s is 7. The disturbance input Md is assumed to be

43

Page 56: Robust and Adaptive Attitude Control Of Spacecraft Using

zero. The eccentricity of the orbit is 0.2. The initial value of the solar aspect angle

is assumed to be φ0 = 45o. The closed-loop responses for the spacecraft model

Eq. (4.2) with δ in Eq. (4.1) for the values of K = 1, K = −1, K = 0.5 are

obtained. It is pointed out that K = 1 refers to a dumbbell satellite aligned with

the local vertical while K = −1 refers to a dumbbell satellite aligned with the local

horizontal. Of course, K = −1 refers to an unstable gravity gradient configuration.

Note that because f ∗ = 0, the control law is not a function of the K-dependent

nonlinear function in the pitch dynamics. The selected responses for K = 1 (in

the left column) K = −1 (in the middle column) and K = 0.5 (in right column)

are shown in Fig. 4.1. It is observed that the pitch angle is smoothly regulated to

100o for all values of K in less than one orbit time. The maximum values of the

control surface deflections are about (28, 25) (deg) for K = 1, (45, 32) (deg), for

K = −1 and (18, 17) (deg) for K = 0.5. It is observed that for K = −1 (unstable

gravity gradient configuration) larger control magnitude is required. In the steady-

state, the waveforms of the control plate deflections are oscillatory, and δ1 and δ2

are almost 180o out of phase. These oscillations in the solar flap deflections are

required to counter the periodic gravity gradient torqueMg so that the attitude can

be maintained at the target value α∗ = 100o. The SRP-dependent control signal

Csv (appearing in the α-dynamics Eq. (4.2)) has been computed using the signal

v from Eq. (4.3). Fig. (4.1) shows the waveform of the control signal Csv. Its

maximum value remains within 2(rad−1). Because δi, i = 1, 2, exhibit oscillations,

the control signal Csv has also oscillatory pattern.

Case Ib. Robust Attitude Control, effect of K: K = 1,−1,0.5 Cs = 5,

e = 0.2, i = 23.5, φ0 = 45o, α∗ = 0o and disturbance input Md = 0

It is desired to control the pitch angle to a target value 0o. For this purpose, one

sets α∗ = 0o. It is assumed that the actual value of the solar parameter is Cs = 5,

but C∗

s is 7. The disturbance input Md is assumed to be zero. The eccentricity of

the orbit is 0.2. The initial value of the solar aspect angle is assumed to be φ0 = 45o.

The closed-loop responses for the spacecraft model Eq. (4.2) with δ in Eq. (4.1) for

44

Page 57: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

4

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.2

0

0.2

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

(a) (c)(b)

δ2

δ2

δ1δ

2

δ1δ

1

Figure 4.2: C = 5, i = 23.5, e = 0.2, φ0 = 45o;α∗ = 0o(a) response for K = 1, (b)response for K = −1, (c) response for K = 0.5

the values of K = 1, K = −1, K = 0.5 are obtained. It is pointed out that K = 1

refers to a dumbbell satellite aligned with the local vertical while K = −1 refers to a

dumbbell satellite aligned with the local horizontal. Of course, K = −1 refers to an

unstable gravity gradient configuration. Note that because f ∗ = 0, the control law

is not a function of the K-dependent nonlinear function in the pitch dynamics. The

selected responses for K = 1 (in the left column) K = −1 (in the middle column)

and K = 0.5 (in right column) are shown in Fig. 4.2. It is observed that the pitch

angle is smoothly regulated to 0o for all values of K in less than one orbit time.

The maximum values of the control surface deflections are about (27, 28) (deg) for

K = 1, (33, 21) (deg), for K = −1 and (23, 22) (deg) for K = 0.5. It is observed

that for K = −1 (unstable gravity gradient configuration) larger control magnitude

is required. In the steady-state, the waveforms of the control plate deflections are

oscillatory, and δ1 and δ2 are almost 180o out of phase. These oscillations in the

solar flap deflections are required to counter the periodic gravity gradient torque

Mg so that the attitude can be maintained at the target value α∗ = 0o. The SRP-

45

Page 58: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.02

0

0.02

0.04

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

δ2 δ

2

δ1δ

1

(b)(a)

Figure 4.3: K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 1000; (a) response forC = 10, (b) response for C = 4

dependent control signal Csv (appearing in the α-dynamics Eq. (4.2)) has been

computed using the signal v from Eq. (4.3). Fig. (4.2) shows the waveform of

the control signal Csv. Its maximum value remains within 2(rad−1). Because δi,

i = 1, 2, exhibit oscillations, the control signal Csv has also oscillatory pattern.

Case IIa. Robust attitude control, effect of Cs: K = 0.5, Cs = 10 and

4, e = 0.2, i = 23.5o, φ0 = 45o, α∗ = 100o and Md = 0

Now the effect of uncertainty in the control input gain (solar parameter) Cs

on the performance of the controller is examined. For this purpose, simulation is

done for the satellite model with actual values of Cs = 10 and Cs = 4. But the

nominal value C∗

s of Cs for the controller design is assumed to be 7. As such one

has uncertainty of −30% and +75% in Cs, respectively. The value of K is 0.5,

e is 0.2 and φ0 = 45 (deg). The controller parameters of Case Ia are retained.

Selected responses are plotted in Fig. 4.3 for Cs = 10 (left column) and Cs = 4

(right column), respectively. One observes the convergence of the pitch angle to

the target value smoothly. The maximum values of the control surface deflections

46

Page 59: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−10

0

10

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

(a) (b)

δ2

δ1

δ2

δ1

Figure 4.4: K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for C = 10,(b) response for C = 4

are about(12, 11) (deg) for Cs = 10 and (22, 20) flap deflections are periodic in the

steady-state similar to Case Ia. Similar to Fig. 4.1, oscillatory pattern of the control

signal Csv is observed in Fig. 4.3.

Case IIb. Robust attitude control, effect of Cs: K = 0.5, Cs = 10 and

4, e = 0.2, i = 23.5o, φ0 = 45o, α∗ = 0o and Md = 0

Now the effect of uncertainty in the control input gain (solar parameter) Cs

on the performance of the controller is examined. For this purpose, simulation is

done for the satellite model with actual values of Cs = 10 and Cs = 4. But the

nominal value C∗

s of Cs for the controller design is assumed to be 7. As such one

has uncertainty of −30% and +75% in Cs, respectively. The value of K is 0.5,

e is 0.2 and φ0 = 45 (deg). The controller parameters of Case Ib are retained.

Selected responses are plotted in Fig. 4.3 for Cs = 10 (left column) and Cs = 4

(right column), respectively. One observes the convergence of the pitch angle to

the target value smoothly. The maximum values of the control surface deflections

are about(15, 14) (deg) for Cs = 10 and (26, 25) (deg) for Cs = 4. As expected the

47

Page 60: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−100

−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

(a) (b)

δ1

δ2

δ1

δ2

Figure 4.5: C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 100o; (a) response for φ0 = 90o,(b) response for φ0 = 135o

choice of larger Cs gives smaller solar flap deflection because control input matrix

has larger magnitude. The solar flap deflections are periodic in the steady-state

similar to Case Ib. Similar to Fig. 4.2, oscillatory pattern of the control signal Csv

is observed in Fig. 4.4.

Case IIIa. Robust attitude control, effect of φ0: K = 0.5, Cs = 5,

e = 0.2, φ0 = 90o and 135o, α∗ = 100o and Md = 0

Now the performance of the SRP control system for different known values of

the solar aspect angle φ is examined. Simulation is done using the initial value

φ0 = 90 (deg) or φ0 = 135 (deg) of the solar aspect angle. The value of K and Cs

are 0.5 and 5, respectively and the eccentricity e is 0.2. The responses are shown in

Fig. 4.5 for φ0 = 90o (left column) and for φ0 = 135o (right column), respectively.

Despite different initial values of φ0, smooth control of the pitch angle is observed.

The maximum values of the control surface deflections are about (22, 38) (deg) for

φ0 = 90o and (61, 30) (deg) for φ0 = 135o. But the response time is about one orbit

time for both values of φ0. Fig. 4.5 shows the oscillatory waveform of the control

48

Page 61: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ2

δ1δ

1

δ2

(a) (b)

Figure 4.6: C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 0o; (a) response for φ0 = 90o,(b) response for φ0 = 135o

signal Csv.

Case IIIb. Robust attitude control, effect of φ0: K = 0.5, Cs = 5,

e = 0.2, φ0 = 90o and 135o, α∗ = 0o and Md = 0

Now the performance of the SRP control system for different known values of

the solar aspect angle φ is examined. Simulation is done using the initial values

φ0 = 90 (deg) and φ0 = 135 (deg) of the solar aspect angle. The value of K and Cs

are 0.5 and 5, respectively and the eccentricity e is 0.2. The responses are shown in

Fig. 4.6 for φ0 = 90o (left column) and for φ0 = 135o (right column), respectively.

Despite different initial value of φ0, smooth control of the pitch angle is observed.

The maximum values of the control surface deflections are about (28, 29) (deg) for

φ0 = 90o and (14, 14) (deg) for φ0 = 135o. The response time is about one orbit

time for both values of φ0. Fig. 4.6 shows the oscillatory waveform of the control

signal Csv.

Case IVa. Robust attitude control despite sinusoidal, random and

pulse disturbance input Md: K = 0.5, Cs = 5, e = 0.2, i = 23.5o, φ0 = 45o,

49

Page 62: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbitsα−

α r [deg

]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−6

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−1

0

1

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

1

2x 10

−6

Orbits

Dis

turb

ance

Md

δ2 δ

2 δ1

(a) (b) (c)

Figure 4.7: C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 100o; (a) response forsinusoidal disturbance, (b) random disturbance (c) pulse type disturbance

α∗ = 100o

Simulation is done to examine the performance of the adaptive controller in the

presence (i) sinusoidal, (ii) random and (iii) pulse type disturbance inputs, shown in

the left, center, and right column in Fig 4.7, respectively. The random disturbance

is generated by passing a white noise with unit variance through a transfer function

F (s) = 5× 10−10/(s+ 5). The controller of Case Ia is retained. Selected responses

are shown in Fig. 4.7. It is observed that the controller achieves the regulation

of the pitch angle to the target value in the presence of each disturbance input.

In the steady-state, it is observed that flap deflection is a periodic function in the

presence of sinusoidal disturbance (Fig. 4.7, left column). The maximum value

of control surface deflection is about (18, 17) (deg). The control signal Csv also

exhibits periodic oscillations. The maximum values of the control surface deflections

are about (18, 17) (deg) for random (center column) and (18, 17) for the pulse-type

disturbance (right column), respectively. The response time remains of similar order

as in the previous cases.

50

Page 63: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbitsα−

α r [deg

]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−6

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−10

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

1

2x 10

−6

Orbits

Dis

turb

ance

Md

(a) (c)(b)

δ2

δ1

δ1

δ2

δ1

δ2

Figure 4.8: C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response forsinusoidal disturbance, (b) random disturbance (c) pulse type disturbance

Case IVb. Robust attitude control despite sinusoidal, random and

pulse disturbance input Md: K = 0.5, Cs = 5, e = 0.2, i = 23.5o, φ0 = 45o,

α∗ = 0o

Simulation is done to examine the performance of the adaptive controller in the

presence (i) sinusoidal, (ii) random and (iii) pulse type disturbance inputs, shown in

the left, center, and right column in Fig 4.8, respectively. The random disturbance

is generated by passing a white noise with unit variance through a transfer function

F (s) = 5× 10−10/(s+ 5). The controller of Case Ib is retained. Selected responses

are shown in Fig. 4.8. It is observed that the controller achieves the regulation

of the pitch angle to the target value in the presence of each disturbance input.

In the steady-state, it is observed that flap deflection is a periodic function in the

presence of sinusoidal disturbance (Fig. 4.8, left column). The maximum value

of control surface deflection is about (23, 22) (deg). The control signal Csv also

exhibits periodic oscillations. The maximum values of the control surface deflections

are about (23, 22) (deg) for random (center column) and (23, 22) for the pulse-type

51

Page 64: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

(a) (b) (c)

δ1

δ2δ

2

δ1

δ2

δ1

Figure 4.9: C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 100o; (a) response for e = 0.05,(b) response for e = 0.4

disturbance (right column), respectively. The response time remains of similar order

as in the previous cases.

Case Va. Robust attitude control, effect of eccentricity: K = 0.5,

Cs = 5, e = 0.05 and 0.4, i = 23.5o, φ0 = 45o, α∗ = 100o disturbance Md = 0

Now the performance of the controller for different values of the eccentricity of

the orbit is examined. Simulation is done for e = 0.05, e = 0.1 and e = 0.4. Note

that similar to Case (I-IV)a, it is assumed that e and i are known. The controller

parameters of Case Ia are retained. The responses are shown in Fig. 4.9(a) for

e = 0.05, Fig. 4.9(b) for e = 0.1 and Fig. 4.9(c) for e = 0.4. Fig. 4.9 shows

smooth regulation of the pitch angle to 100 (deg). It is observed that compared to

the responses for e = 0.05, larger eccentricity of the orbit causes presence of higher

harmonic terms in the solar flap rotation angle waveforms. The response time is

almost of similar order as in Cases Ia and IIa. The maximum values of the control

surface deflections are about (17, 15) (deg) for e = 0.05, (17, 16) (deg) for e = 0.1

and (22, 20) (deg) for e = 0.4, respectively. It is observed that the peaks in δi and

52

Page 65: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ2

δ1

δ1

δ2

δ1

δ2

(a) (b) (c)

Figure 4.10: C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 0o; (a) response for e = 0.05,(b) response for e = 0.4

the control signal Csv are larger for the control of spacecraft in orbits of higher

eccentricity.

Case Vb. Robust attitude control, effect of eccentricity: K = 0.5,

Cs = 5, e = 0.05 and 0.4, i = 23.5o, φ0 = 45o, α∗ = 100o disturbance Md = 0

Now the performance of the controller for different values of the eccentricity of

the orbit is examined. Simulation is done for e = 0.05, e = 0.1 and e = 0.4. Note

that similar to Case (I-IV)b, it is assumed that e and i are known. The controller

parameters of Case Ia are retained. The responses are shown in Fig. 4.10(a) for

e = 0.05, Fig. 4.10(b) for e = 0.1 and Fig. 4.10(c) for e = 0.4. Fig. 3.10 shows

smooth regulation of the pitch angle to 0 (deg). It is observed that compared to

the responses for e = 0.05, larger eccentricity of the orbit causes presence of higher

harmonic terms in the solar flap rotation angle waveforms. The response time is

almost of similar order as in Cases Ib and IIb. The maximum values of the control

surface deflections are about (26, 23) (deg) for e = 0.05, (25, 22) (deg) for e = 0.1

and (25, 23) (deg) for e = 0.4. It is observed that the peaks in δi and the control

53

Page 66: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.05

0

0.05

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

Orbits

Csv

[rad

−1]

(b)

δ1

δ2δ

2

δ1

(a) (b)

Figure 4.11: C = 5, K = 0.5, φ0 = 45o, α∗ = 100o; (a) response for i = 30o, e = 0.1,(b) response for i = 10o, e = 0.2

signal Csv are larger for the control of spacecraft in orbits of higher eccentricity.

Case VIa. Robust attitude control, effect of e and i : K = 0.5, Cs = 5,

α∗ = 100o, φ0 = 45o, Md = 0

Simulation is done for two sets of values (e, i) = (0.1, 30o) and (0.2, 10o). Of

course, it is assumed that these values of (e, i) are known to the designer. The

closed-loop responses for (e, i) = (0.1, 30o) and (e, i) = (0.2, 10o) are shown in

Fig. 4.11(a) (left column) and Fig. 4.11(b) (right column), respectively. It is seen

that smooth regulation of the pitch angle to 1000 is accomplished for each (e, i).

The maximum values of the control surface deflections are about (18, 16) (deg) for

(e, i) = (0.1, 30o) and (17, 16) (deg) for (e, i) = (0.2, 10o), respectively. It is seen

that the waveform of the control signal Csv is periodic in the steady-state.

Case VIb. Robust attitude control, effect of e and i : K = 0.5, Cs = 5,

α∗ = 0o, φ0 = 45o, Md = 0

Simulation is done for two sets of values (e, i) = (0.1, 30o) and (0.2, 10o). Of

course, it is assumed that these values of (e, i) are known to the designer. The closed-

54

Page 67: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−0.1

0

0.1

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

(a) (b)

δ2

δ1

δ1

δ2

Figure 4.12: C = 5, K = 0.5, φ0 = 45o, α∗ = 0o; (a) response for i = 30o, e = 0.1,(b) response for i = 10o, e = 0.2

loop responses for (e, i) = (0.1, 30o) and (e, i) = (0.2, 10o) are shown in Fig. 4.12(a)

(left column) and Fig. 4.12(b) (right column), respectively. It is seen that smooth

regulation of the pitch angle to 00 is accomplished for each (e, i). The maximum

values of the control surface deflections are about (26, 23) (deg) for (e, i) = (0.1, 30o)

and (22, 21) (deg) for (e, i) = (0.2, 10o), respectively. It is seen that the waveform

of the control signal Csv is periodic in the steady-state.

Extensive simulation has been performed for several values of the solar aspect

angle, the eccentricity e of the orbit, the orbit inclination i, and the model param-

eters K and Cs. These results show that the designed control law accomplishes

robust regulation of the pitch angle trajectory, even in the presence of disturbance

input.

55

Page 68: Robust and Adaptive Attitude Control Of Spacecraft Using

4.7 Conclusions

The design of an adaptive control system for the pitch angle control of spacecraft

with uncertain dynamics in elliptic orbits, using output feedback with high gain

estimator was considered. The parameters of the nonaffine-in-control spacecraft

model were assumed to be unknown, and external disturbance input was assumed

to be acting on the satellite. An adaptive control law was designed for the tracking of

reference pitch angle trajectory. The control system included a high gain estimator

for the estimation of pitch angle derivatives of the satellite. In the closed-loop

system precise pitch attitude control was accomplished, despite uncertainties and

disturbance input.

56

Page 69: Robust and Adaptive Attitude Control Of Spacecraft Using

Chapter 5

Robust Finite Time Satellite

Attitude Control Using Output

Feedback

5.1 Introduction

In the previous chapter, an output feedback controller using a high gain estimator

was designed for adaptive control of pitch angle. In this chapter, the controller

used in Chapter 3 is used but with the pitch angular velocity and acceleration

signals provided by the universal output feedback SISO controller which acts as an

estimator. The estimator is actually a differentiator which guarantees finite time

convergence with the actual derivatives of the signal. This design like the one in

the previous chapter, avoids measurement of pitch angular velocity and acceleration

which would have not been practical.

5.2 Problem Formulation

For the development of the control law, the satellite orbiting in an elliptic orbit

derived in the Chapter 2 is considered. The dynamics of the satellite is described

by

57

Page 70: Robust and Adaptive Attitude Control Of Spacecraft Using

α = f0(α, α, θ) + Csv(α, θ, δ) (5.1)

where the nonlinear functions f0 and v are

f0(α, α, θ) = (1 + e cos θ)−1[2e sin θα− 1.5K sin 2(α− θ) +Mdn]

v(α, θ, δ) = (1 + e cos θ)−1σsMsn (5.2)

Note that the disturbance input Md is included in the nonlinear function f0. (The

argument θ denotes the dependence of f0 onMd.) For the design of the controller, it

is assumed that the nonlinear function f0 as well as the solar parameter Cs are not

known. The Eq. (5.2) describing the rotational motion of the satellite is nonaffine-

in-control variables δ1 and δ2. Therefore, it will be convenient to introduce the

derivatives of δ1 and δ2 as control inputs. Differentiating Eq. (5.2) with respect to

θ yieldsd3α

dθ3= fa(α, α, θ, δ) +B(α, θ, δ)δ (5.3)

where

fa(α, α, θ) =∂f0(α, α, θ)

∂αα +

∂f0(α, α, θ)

∂αα +

∂f0(α, α, θ)

∂θ+ Cs[

∂v

∂αα +

∂v

∂θ]

B(α, θ, δ) = Cs∂v

∂δ

Note that for obtaining the expression for fa, α given in Eq. (5.2) has been sub-

stituted for α. (Often the arguments of functions are suppressed for simplicity in

notation.) It is assumed here that the nonlinear function fa and the matrix B are

not known precisely. Let Ω1 ⊂ R4 be a region defined by

Ω1 = (α, θ, δ) : rankB(α, θ, δ) = 1

In this study, the motion of the satellite evolving in the region Ω1 will be of inter-

est. Note that outside the region Ω1, the input signal δ has no effect on the pitch

dynamics in Eq. (5.4).

58

Page 71: Robust and Adaptive Attitude Control Of Spacecraft Using

Suppose that αr is a given reference pitch angle trajectory. The objective is

to design a robust control law such that the pitch angle α converges to the refer-

ence trajectory αr in a finite time, despite the presence of disturbance input and

parameter uncertainties in fa and B, using only output feedback of pitch angle.

5.3 Robust Control System

The controller used in Chapter 3 is described by

ξ1 = α− αr; ξ2 = α− αr; ξ3 = α− αr

where α − αr is the tracking error. Then controller used in Chapter 3 is given

by

un(ξ) = −k1|ξ1|ν1sgn(ξ1)− k2|ξ2|ν2sgn(ξ2)− k3|ξ3|ν3sgn(ξ3) (5.4)

ur = −G(ξ)sgn(s) (5.5)

s(ξ, ξa) = ξ3 + ξa (5.6)

where the auxillary signal ζa satisfies

ξa = −un (5.7)

The pitch dynamics are assumed to be not available and are replaced with the

signals for pitch angular velocity and acceleration coming from the estimator. There-

fore using previous equations modified to

un(ζ) = −k1|ζ0|ν1sgn(ζ0)− k2|ζ1|ν2sgn(ζ1)− k3|ζ2|ν3sgn(ζ2) (5.8)

ur = −G(ζ)sgn(s) (5.9)

59

Page 72: Robust and Adaptive Attitude Control Of Spacecraft Using

s(ζ, ζa) = ζ2 + ζa (5.10)

where the auxillary signal ζa satisfies

ζa = −un (5.11)

where ζi(i = 0, 1, 2) are the estimated values of (α− αr), (α− αr) and (α− αr)

respectively.

5.4 Universal Output Feedback SISO Controller-Estimator

In this subsection, the structure of the estimator is considered. The design is based

on the results obtained in [45,46]. The estimator used here is a combination of

a sliding mode controller with relative degree r based on an r-sliding-finite-time-

convergent sliding mode and of an exact robust (r − 1)th order differentiator with

finite time convergence. The estimator provides exact tracking after a finite time

transient. The corresponding proof and theorems are provided in [45,46].

The following set of equations constitute the estimator

ζ0 = w0, w0 = −ψ0|ζ0 − fe|(r−1)/rsgn(ζ0 − fe) + ζ1

ζ1 = w1, w1 = −ψ1|ζ0 − ψ0|(r−2)/(r−1)sgn(ζ0 − ψ0) + ζ2

...

ζr−2 = wr−2, wr−2 = −ψr−2|ζ0 − ψr−3|1/2sgn(ζr−2 − ψr−3) + ζr−1

ζr−1 = −ψr−1sgn(ζr−1 − ψr−2)

(5.12)

where parameters fe = α − αr, ψi of the estimator are chosen according to

the condition |f (r)e | ≤ L. L is chosen using ψi = ψ0,iL where psi0,i are chosen by

computer simulation.

60

Page 73: Robust and Adaptive Attitude Control Of Spacecraft Using

5.5 Simulations results

This section presents the results of digital simulation. The complete closed-loop

system including the satellite model Eq. (5.2) with and without external disturbance

moment, and the control law Eqs. (5.8), (5.9), (5.10) and (5.11) is simulated for a

set of values of K, Cs, eccentricity e, orbit inclination i and solar aspect angle φ for

target angles α = 0o, 100o. The solar aspect angle φ is a slowly varying function.

The function φ given by

φ(θ) = φ0 + (∂φ/∂θ)(θ − θ0)

is used here for computation, where φ0 = φ(θ0). The inclination of the orbital plane

of the geosynchronous satellite is i = 23.5o. The semi-major axis is a = 42, 241 km

and Iz is 500 kg.m2. The initial conditions of the spacecraft are chosen as θo = 0,

α(θo) = 100o, 0o and α(θo) = 0. The initial values of the flap deflections are δ1(θo) =

0o and δ2(θo) = 0o. For a limited perturbation in B, it is assumed to be of the form

B = (C∗

s + ∆Cs)∂v∂δ, where the vector function (∂v/∂δ) is known, C∗

s is the known

nominal value, and ∆Cs is the uncertain portion of Cs. The nonlinear nominal

function f ∗(ξ, θ, δ is assumed to be zero for simplicity in implementation; that is,

f(ξ, θ, δ) = ∆f . Apparently, such a choice of ∆f represents a large uncertainty

in the model. The reference pitch angle trajectory is generated by a fifth-order

reference generator given by

d6αr

dθ6= −p5

d5αr

dθ5− p4

d4αr

dθ4− p3

d3αr

dθ3− p2

d2αr

dθ2− p1

dαr

dθ− p0(αr − α∗) (5.13)

where α∗ is the target pitch angle. The parameters in Eq. (5.12) are such that

the poles of the reference generator are [−4,−2.5,−3.5,−3,−3.8,−3.5]. The initial

conditions are αr(0) = 0 for Case Ia, α = 100o for Case Ib and djαr(0)/dθj = 0,

j = 1, 2...6. The feedback gains in the nominal signal un are k1 = 29512, k2 = 2874,

and k3 = 93. Therefore, the roots of Eq. (5.4) are [-28, -31, -34]. The switching gain

in Eq. (3.24) is selected as G = 0.25. Here a constant gain G is used for simplicity in

implementation. The initial values for the estimates of the pitch dynamics tracking

61

Page 74: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−4

−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

(a) (b)

δ2δ

2

δ1

δ1

Figure 5.1: C = 5, i = 23.5, e = 0.2, φ0 = 45o; α∗ = 100o(a) response for K = 1,(b) response for K = 0.5

error are chosen as ζ1 = 10(deg), ζ2 = 1 and ζ3 = 0.66. The Lipschitz constant, L =

1. The parameters ψi are chosen based on previously simulation tested values for

L = 1. These controller parameters have been selected by observing the simulated

responses.

The signum function in the control law is replaced by a saturation function

sat(s). Suppose that 0 < ǫsat ≪ 1. The function sat(s) is then defined as

sat(s) =

1ǫsat

s |s| ≤ ǫsat

1 s > ǫsat

−1 s < −ǫsat

Here ǫsat is taken as 3.5× 10−5

Case Ia. Robust Attitude Control, effect of K: K = 1 and 0.5, Cs = 5,

e = 0.2, i = 23.5, φ0 = 45o, α∗ = 100o and disturbance input Md = 0

It is desired to control the pitch angle to a target value 100o. For this purpose,

one sets α∗ = 100o in Eq. (5.7). It is assumed that the actual value of the solar

62

Page 75: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

αe

αe

αα

α. α.

αe. α

e.

α..

α..

αe..α

e..

(a) (b)

Figure 5.2: Estimates of α, α and α for α∗ = 100o(a) response for K = 1, (b)response for K = 0.5

parameter is Cs = 5, but C∗

s is 7. The disturbance input Md is assumed to be zero.

The eccentricity of the orbit is 0.2. The initial value of the solar aspect angle is

assumed to be φ0 = 45o. The closed-loop responses for the spacecraft model Eq.

(5.2) with δ in Eq. (5.1) for the values of K = 1, K = 0.5 are obtained. It is

pointed out that K = 1 refers to a dumbbell satellite aligned with the local vertical

while K = −1 refers to a dumbbell satellite aligned with the local horizontal. Note

that because f ∗ = 0, the control law is not a function of the K-dependent nonlinear

function in the pitch dynamics. The selected responses for K = 1 (in the left

column) and K = 0.5 (in right column) are shown in Fig. 5.1. It is observed that

the pitch angle is smoothly regulated to 100o for all values of K in less than one

orbit time. The maximum values of the control surface deflections are about (46, 41)

(deg) for K = 1 and (37,31) (deg) for K = 0.5. These oscillations in the solar flap

deflections are required to counter the periodic gravity gradient torque Mg so that

the attitude can be maintained at the target value α∗ = 100o. The SRP-dependent

control signal Csv (appearing in the α-dynamics Eq. 5.2)) has been computed using

63

Page 76: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−1

0

1x 10

−3

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

4

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−4

−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−10

0

10

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

(a) (b)

δ1

δ1δ

2

δ2

Figure 5.3: C = 5, i = 23.5, e = 0.2, φ0 = 45o;α∗ = 0o(a) response for K = 1, (b)response for K = −1, (c) response for K = 0.5

the signal v from Eq. (5.3). Fig. (5.1) shows the waveform of the control signal

Csv. Its maximum value remains within 2(rad−1). Because δi, i = 1, 2, exhibit

oscillations, the control signal Csv has also oscillatory pattern. Fig (5.2) shows the

convergence of estimated (αe) and actual (α) values of pitch angle and its derivatives

for each case.

Case Ib. Robust Attitude Control, effect of K: K = 1,0.5 Cs = 5,

e = 0.2, i = 23.5, φ0 = 45o, α∗ = 0o and disturbance input Md = 0

It is desired to control the pitch angle to a target value 0o. For this purpose,

one sets α∗ = 0o in Eq. (5.7). It is assumed that the actual value of the solar

parameter is Cs = 5, but C∗

s is 7. The disturbance input Md is assumed to be zero.

The eccentricity of the orbit is 0.2. The initial value of the solar aspect angle is

assumed to be φ0 = 45o. The closed-loop responses for the spacecraft model Eq.

(5.2) with δ in Eq. (5.1) for the values of K = 1, K = 0.5 are obtained. It is pointed

out that K = 1 refers to a dumbbell satellite aligned with the local vertical while

K = −1 refers to a dumbbell satellite aligned with the local horizontal. The selected

64

Page 77: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

α..α..

αe α

e

α α

α. α.

αe. α

e.

αe..

αe..

(a) (b)

Figure 5.4: Estimates of α, α and α for α∗ = 0o(a) response for K = 1, (b) responsefor K = 0.5

responses for K = 1 (in the left column) and K = 0.5 (in right column) are shown

in Fig. 5.3. It is observed that the pitch angle is smoothly regulated to 0o for all

values of K in less than one orbit time. The maximum values of the control surface

deflections are about (24, 24) (deg) for K = 1 and (20, 19) (deg) for K = 0.5. In the

steady-state, the waveforms of the control plate deflections are oscillatory, and δ1

and δ2 are almost 180o out of phase. These oscillations in the solar flap deflections

are required to counter the periodic gravity gradient torque Mg so that the attitude

can be maintained at the target value α∗ = 0o. The SRP-dependent control signal

Csv (appearing in the α-dynamics Eq. (5.2)) has been computed using the signal v

from Eq. (5.3). Fig. 5.3 shows the waveform of the control signal Csv. Its maximum

value remains within 2(rad−1). Because δi, i = 1, 2, exhibit oscillations, the control

signal Csv has also oscillatory pattern. Fig. 5.4 shows the convergence of estimated

(αe) and actual (α) values of pitch angle and its derivatives for each case.

Case IIa. Robust attitude control, effect of Cs: K = 0.5, Cs = 10 and

4, e = 0.2, i = 23.5o, φ0 = 45o, α∗ = 100o and Md = 0

65

Page 78: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2

4x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

δ2

δ2 δ

1 δ1

(a) (b)

Figure 5.5: K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 1000; (a) response forC = 10, (b) response for C = 4

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−2

0

2

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−2

0

2

Orbits

α.. αe..

αe.. α

e..

αe

αe

α α

α. α.

αe.

αe.

α.. α..

(a) (b)

Figure 5.6: Estimates of α, α and α for α∗ = 100o(a) response for C = 10, (b)response for C = 4

66

Page 79: Robust and Adaptive Attitude Control Of Spacecraft Using

Now the effect of uncertainty in the control input gain (solar parameter) Cs

on the performance of the controller is examined. For this purpose, simulation is

done for the satellite model with actual values of Cs = 10 and Cs = 4. But the

nominal value C∗

s of Cs for the controller design is assumed to be 7. As such one

has uncertainty of −30% and +75% in Cs, respectively. The value of K is 0.5,

e is 0.2 and φ0 = 45 (deg). The controller parameters of Case Ia are retained.

Selected responses are plotted in Fig. 5.5 for Cs = 10 (left column) and Cs = 4

(right column), respectively. One observes the convergence of the pitch angle to

the target value smoothly. The maximum values of the control surface deflections

are about(25, 19) (deg) for Cs = 10 and (42, 36) (deg) for Cs = 4. As expected the

choice of larger Cs gives smaller solar flap deflection because control input matrix

has larger magnitude. The solar flap deflections are periodic in the steady-state

similar to Case Ia. Similar to Fig. 5.1, oscillatory pattern of the control signal Csv

is observed in Fig. 5.5. Fig. 5.6 shows the convergence of estimated (αe) and actual

(α) values of pitch angle and its derivatives for each case.

Case IIb. Robust attitude control, effect of Cs: K = 0.5, Cs = 10 and

4, e = 0.2, i = 23.5o, φ0 = 45o, α∗ = 0o and Md = 0

Now the effect of uncertainty in the control input gain (solar parameter) Cs

on the performance of the controller is examined. For this purpose, simulation is

done for the satellite model with actual values of Cs = 10 and Cs = 4. But the

nominal value C∗

s of Cs for the controller design is assumed to be 7. As such one

has uncertainty of −30% and +75% in Cs, respectively. The value of K is 0.5,

e is 0.2 and φ0 = 45 (deg). The controller parameters of Case Ib are retained.

Selected responses are plotted in Fig. 5.7 for Cs = 10 (left column) and Cs = 4

(right column), respectively. One observes the convergence of the pitch angle to

the target value smoothly. The maximum values of the control surface deflections

are about(13, 11) (deg) for Cs = 10 and (23, 23) (deg) for Cs = 4. As expected the

choice of larger Cs gives smaller solar flap deflection because control input matrix

has larger magnitude. The solar flap deflections are periodic in the steady-state

67

Page 80: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−10

0

10

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ1 δ

1 δ2

δ2

(a) (b)

Figure 5.7: K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response for C = 10,(b) response for C = 4

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

αe

α

α.

αe.

αe..

α..α..

αe

α

α.

αe.

αe..

(a) (b)

Figure 5.8: Estimates of α, α and α for α∗ = 0o(a) response for C = 10, (b) responsefor C = 4

68

Page 81: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−4

−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−100

−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

(a) (b)

δ1

δ1

δ2

δ2

Figure 5.9: C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 100o; (a) response for φ0 = 90o,(b) response for φ0 = 135o

similar to Case Ib. Similar to Fig. 5.3, oscillatory pattern of the control signal Csv

is observed in Fig. 5.7. Fig. 5.8 shows the convergence of estimated (αe) and actual

(α) values of pitch angle and its derivatives for each case.

Case IIIa. Robust attitude control, effect of φ0: K = 0.5, Cs = 5,

e = 0.2, φ0 = 90o and 135o, α∗ = 100o and Md = 0

Now the performance of the SRP control system for different known values of

the solar aspect angle φ is examined. Simulation is done using the initial value

φ0 = 90 (deg) or φ0 = 135 (deg) of the solar aspect angle. The value of K and Cs

are 0.5 and 5, respectively and the eccentricity e is 0.2. The responses are shown in

Fig. 5.9 for φ0 = 90o (left column) and for φ0 = 135o (right column), respectively.

Despite different initial values of φ0, smooth control of the pitch angle is observed.

The maximum values of the control surface deflections are about (29, 25) (deg) for

φ0 = 90o and (67, 17) (deg) for φ0 = 135o. But the response time is about one orbit

time for both values of φ0. Fig. 5.9 shows the oscillatory waveform of the control

signal Csv. Fig. 5.10 shows the convergence of estimated (αe) and actual (α) values

69

Page 82: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−2

0

2

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−2

0

2

Orbits

α.. αe..

αe

αe

α α

α. α.

αe.

αe.

α.. α..

αe..α

e..

(a) (b)

Figure 5.10: Estimates of α, α and α for α∗ = 100o(a) response for φ0 = 90o, (b)response for φ0 = 135o

of pitch angle and its derivatives for each case.

Case IIIb. Robust attitude control, effect of φ0: K = 0.5, Cs = 5,

e = 0.2, φ0 = 90o and 135o, α∗ = 0o and Md = 0

Now the performance of the SRP control system for different known values of

the solar aspect angle φ is examined. Simulation is done using the initial values

φ0 = 90 (deg) and φ0 = 135 (deg) of the solar aspect angle. The value of K and Cs

are 0.5 and 5, respectively and the eccentricity e is 0.2. The responses are shown in

Fig. 5.11 for φ0 = 90o (left column) and for φ0 = 135o (right column), respectively.

Despite different initial value of φ0, smooth control of the pitch angle is observed.

The maximum values of the control surface deflections are about (29, 29) (deg) for

φ0 = 90o and (13, 13) (deg) for φ0 = 135o. The response time is about one orbit

time for both values of φ0. Fig. 5.11 shows the oscillatory waveform of the control

signal Csv. Fig. 5.12 shows the convergence of estimated (αe) and actual (α) values

of pitch angle and its derivatives for each case.

Case IVa. Robust attitude control despite sinusoidal, random and

70

Page 83: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−4

−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

40

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−4

−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ1

δ1

δ2

δ2

(a) (b)

Figure 5.11: C = 5, K = 0.5, i = 23.5, e = 0.2, α∗ = 0o; (a) response for φ0 = 90o,(b) response for φ0 = 135o

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

αe α

e

α α

α. α.

αe. α

e.

α..α..

αe.. α

e..

(a) (b)

Figure 5.12: Estimates of α, α and α for α∗ = 0o(a) response for φ0 = 90o, (b)response for φ0 = 135o

71

Page 84: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbitsα−

α r [deg

]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−6

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−10

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

1

2x 10

−6

Orbits

Dis

turb

ance

Md

δ2 δ

1

δ2 δ

1

δ2

δ1

(a)(c)(b)

Figure 5.13: C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 100o; (a) responsefor sinusoidal disturbance, (b) random disturbance (c) pulse type disturbance

pulse disturbance input Md: K = 0.5, Cs = 5, e = 0.2, i = 23.5o, φ0 = 45o,

α∗ = 100o

Simulation is done to examine the performance of the adaptive controller in the

presence (i) sinusoidal, (ii) random and (iii) pulse type disturbance inputs, shown in

the left, center, and right column in Fig 3.7, respectively. The random disturbance

is generated by passing a white noise with unit variance through a transfer function

F (s) = 5× 10−10/(s+ 5). The controller of Case Ia is retained. Selected responses

are shown in Fig. 5.13. It is observed that the controller achieves the regulation

of the pitch angle to the target value in the presence of each disturbance input.

In the steady-state, it is observed that flap deflection is a periodic function in the

presence of sinusoidal disturbance (Fig. 5.13, left column). The maximum value

of control surface deflection is about (37, 31) (deg). The control signal Csv also

exhibits periodic oscillations. The maximum values of the control surface deflections

are about (37, 31) (deg) for random (center column) and (37, 31) for the pulse-type

disturbance (right column), respectively. The response time remains of similar order

72

Page 85: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

αe

αeα

eαα α

α. α.α.

αe.α

e.α

e.

α..α..α..

αe.. α

e.. α

e..

(c)(b)(a)

Figure 5.14: Estimates of α, α and α for α∗ = 100o(a) response for sinusoidaldisturbance, (b) random disturbance (c) pulse type disturbance

as in the previous cases. Fig. 5.14 shows the convergence of estimated (αe) and

actual (α) values of pitch angle and its derivatives for each case.

Simulation is done to examine the performance of the adaptive controller in the

presence (i) sinusoidal, (ii) random and (iii) pulse type disturbance inputs, shown in

the left, center, and right column in Fig 3.8, respectively. The random disturbance

is generated by passing a white noise with unit variance through a transfer function

F (s) = 5× 10−10/(s+ 5). The controller of Case Ib is retained. Selected responses

are shown in Fig. 5.15. It is observed that the controller achieves the regulation

of the pitch angle to the target value in the presence of each disturbance input.

In the steady-state, it is observed that flap deflection is a periodic function in the

presence of sinusoidal disturbance (Fig. 5.15, left column). The maximum value

of control surface deflection is about (20, 19) (deg). The control signal Csv also

exhibits periodic oscillations. The maximum values of the control surface deflections

are about (20, 19) (deg) for random (center column) and (20, 19) for the pulse-type

disturbance (right column), respectively. The response time remains of similar order

73

Page 86: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 2−2

0

2x 10

−6

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]0 0.5 1 1.5 2

−2

0

2

OrbitsC

sv [r

ad−1

]

0 0.5 1 1.5 2−2

0

2x 10

−10

Orbits

Dis

turb

ance

Md

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

1

2x 10

−6

Orbits

Dis

turb

ance

Md

(a)(c)(b)

δ2 δ

2 δ1

Figure 5.15: C = 5, K = 0.5, i = 23.5, e = 0.2, φ0 = 45o, α∗ = 0o; (a) response forsinusoidal disturbance, (b) random disturbance (c) pulse type disturbance

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

(a) (b) (c)

αe

αeα

e

α α α

α. α. α.

αe.α

e.α

e.

α..

α..α..

αe..

αe..α

e..

Figure 5.16: Estimates of α, α and α for α∗ = 0o(a) response for sinusoidal distur-bance, (b) random disturbance (c) pulse type disturbance

74

Page 87: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−4

−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−2

0

2

4x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−4

−2

0

2

Orbits

Csv

[rad

−1]

(a) (c)(b)

δ2

δ2 δ

2

δ1

δ1δ

1

Figure 5.17: C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 100o; (a) response fore = 0.05, (b) response for e = 0.4

as in the previous cases. Fig. 5.16 shows the convergence of estimated (αe) and

actual (α) values of pitch angle and its derivatives for each case.

Case Va. Robust attitude control, effect of eccentricity: K = 0.5,

Cs = 5, e = 0.05 and 0.4, i = 23.5o, φ0 = 45o, α∗ = 100o disturbance Md = 0

Now the performance of the controller for different values of the eccentricity of

the orbit is examined. Simulation is done for e = 0.05, e = 0.1 and e = 0.4. Note

that similar to Case (I-IV)a, it is assumed that e and i are known. The controller

parameters of Case Ia are retained. The responses are shown in Fig. 5.17(a) for

e = 0.05, Fig. 5.17(b) for e = 0.1 and Fig. 5.17(c) for e = 0.4. Fig. 5.17 shows

smooth regulation of the pitch angle to 100 (deg). It is observed that compared

to the responses for e = 0.05, larger eccentricity of the orbit causes presence of

higher harmonic terms in the solar flap rotation angle waveforms. The response

time is almost of similar order as in Cases Ia and IIa. The maximum values of the

control surface deflections are about (33, 25) (deg) for e = 0.05, (34, 26) (deg) for

e = 0.1 and (48, 40) (deg) for e = 0.4, respectively. It is observed that the peaks

75

Page 88: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

(a) (c)(b)

αe

αeα

e

α αα

α.

αe. α

e.

α.

αe.

α.

α..α.. α..

αe..

αe..α

e..

Figure 5.18: Estimates of α, α and α for α∗ = 100o (a) response for e = 0.05, (b)response for e = 0.4

in δi and the control signal Csv are larger for the control of spacecraft in orbits of

higher eccentricity. Fig. 5.18 shows the convergence of estimated (αe) and actual

(α) values of pitch angle and its derivatives for each case.

Case Vb. Robust attitude control, effect of eccentricity: K = 0.5,

Cs = 5, e = 0.05 and 0.4, i = 23.5o, φ0 = 45o, α∗ = 100o disturbance Md = 0

Now the performance of the controller for different values of the eccentricity of

the orbit is examined. Simulation is done for e = 0.05, e = 0.1 and e = 0.4. Note

that similar to Case (I-IV)b, it is assumed that e and i are known. The controller

parameters of Case Ia are retained. The responses are shown in Fig. 5.19(a) for

e = 0.05, Fig. 5.19(b) for e = 0.1 and Fig. 5.19(c) for e = 0.4. Fig. 5.19 shows

smooth regulation of the pitch angle to 0 (deg). It is observed that compared to

the responses for e = 0.05, larger eccentricity of the orbit causes presence of higher

harmonic terms in the solar flap rotation angle waveforms. The response time is

almost of similar order as in Cases Ib and IIb. The maximum values of the control

surface deflections are about (9, 13) (deg) for e = 0.05, (12, 15) (deg) for e = 0.1 and

76

Page 89: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

−10

0

10

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−4

−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−50

0

50

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

0

2

Orbits

Csv

[rad

−1]

δ2 δ

1

δ2 δ

1

δ2

δ2

(a)(c)(b)

Figure 5.19: C = 5, K = 0.5, i = 23.5, φ0 = 45o, α∗ = 0o; (a) response for e = 0.05,(b) response for e = 0.4

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4−4

−3

−2

−1

0

1

Orbits

α.. αe..

(a) (b) (c)

αe

αeα

e

α αα

α. α.α.

αe. α

e.α

e.

α..α..α..

αe.. α

e..α

e..

Figure 5.20: Estimates of α, α and α for α∗ = 0o (a) response for e = 0.05, (b)response for e = 0.4

77

Page 90: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−40

−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−2

−1

0

1

Orbits

Csv

[rad

−1]

(a) (b)

δ2

δ2

δ1

δ1

Figure 5.21: C = 5, K = 0.5, φ0 = 45o, α∗ = 100o; (a) response for i = 30o, e = 0.1,(b) response for i = 10o, e = 0.2

(29, 25) (deg) for e = 0.4. It is observed that the peaks in δi and the control signal

Csv are larger for the control of spacecraft in orbits of higher eccentricity. Fig. 5.20

shows the convergence of estimated (αe) and actual (α) values of pitch angle and

its derivatives for each case.

Case VIa. Robust attitude control, effect of e and i : K = 0.5, Cs = 5,

α∗ = 100o, φ0 = 45o, Md = 0

Simulation is done for two sets of values (e, i) = (0.1, 30o) and (0.2, 10o). Of

course, it is assumed that these values of (e, i) are known to the designer. The

closed-loop responses for (e, i) = (0.1, 30o) and (e, i) = (0.2, 10o) are shown in

Fig. 5.21(a) (left column) and Fig. 5.21(b) (right column), respectively. It is seen

that smooth regulation of the pitch angle to 1000 is accomplished for each (e, i).

The maximum values of the control surface deflections are about (35, 27) (deg) for

(e, i) = (0.1, 30o) and (36, 29) (deg) for (e, i) = (0.2, 10o), respectively. It is seen

that the waveform of the control signal Csv is periodic in the steady-state. Fig. 5.22

shows the convergence of estimated (αe) and actual (α) values of pitch angle and

78

Page 91: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−2

0

2

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

1.5

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−2

0

2

Orbits

α.. αe..

(a) (b)

αe

αe

α α

α. α.

αe.

αe.

α.. α..

αe.. α

e..

Figure 5.22: Estimates of α, α and α for α∗ = 100o (a) response for i = 30o, e = 0.1,(b) response for i = 10o, e = 0.2

its derivatives for each case.

Case VIb. Robust attitude control, effect of e and i : K = 0.5, Cs = 5,

α∗ = 0o, φ0 = 45o, Md = 0

Simulation is done for two sets of values (e, i) = (0.1, 30o) and (0.2, 10o). Of

course, it is assumed that these values of (e, i) are known to the designer. The

closed-loop responses for (e, i) = (0.1, 30o) and (e, i) = (0.2, 10o) are shown in

Fig. 5.23(a) (left column) and Fig. 5.23(b) (right column), respectively. It is

seen that smooth regulation of the pitch angle to 00 is accomplished for each (e, i).

The maximum values of the control surface deflections are about (17, 13) (deg) for

(e, i) = (0.1, 30o) and (33, 25) (deg) for (e, i) = (0.2, 10o), respectively. It is seen

that the waveform of the control signal Csv is periodic in the steady-state. Fig. 5.24

shows the convergence of estimated (αe) and actual (α) values of pitch angle and

its derivatives for each case.

Extensive simulation has been performed for several values of the solar aspect

angle, the eccentricity e of the orbit, the orbit inclination i, and the model param-

79

Page 92: Robust and Adaptive Attitude Control Of Spacecraft Using

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−5

0

5x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−20

0

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

0 0.5 1 1.5 20

50

100

Orbits

α [d

eg]

0 0.5 1 1.5 2−4

−2

0

2x 10

−4

Orbits

α−α r [d

eg]

0 0.5 1 1.5 2−10

0

10

20

Orbits

δ i [deg

]

0 0.5 1 1.5 2−1

0

1

2

Orbits

Csv

[rad

−1]

δ2

δ1

δ1

δ2

(a) (b)

Figure 5.23: C = 5, K = 0.5, φ0 = 45o, α∗ = 100o; (a) response for i = 30o, e = 0.1,(b) response for i = 10o, e = 0.2

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

0 0.1 0.2 0.3 0.4 0.5

0

20

40

60

80

100

Orbits

α α e [d

eg]

0 0.1 0.2 0.3 0.4 0.5−1.5

−1

−0.5

0

0.5

1

Orbits

α. αe.

0 0.1 0.2 0.3 0.4 0.5−4

−3

−2

−1

0

1

Orbits

α.. αe..

αe α

e

α α

α. α.

αe. α

e.

α..α..

αe.. α

e..

(a) (b)

Figure 5.24: Estimates of α, α and α for α∗ = 0o (a) response for i = 30o, e = 0.1,(b) response for i = 10o, e = 0.2

80

Page 93: Robust and Adaptive Attitude Control Of Spacecraft Using

eters K and Cs. These results show that the designed control law accomplishes

robust regulation of the pitch angle trajectory, even in the presence of disturbance

input.

5.6 Conclusions

The design of a robust finite-time control system for the pitch angle control of

spacecraft with uncertain dynamics using output feedback in elliptic orbits using

SRP was considered. The parameters of the nonaffine-in-control spacecraft model

were assumed to be unknown, and external disturbance input was assumed to be

acting on the satellite. A robust control law designed in Chapter 3 was used for the

tracking of reference pitch angle trajectory in a finite time along with an estimator

for estimating pitch derivatives. The control system included a nominal nonlinear

control designed for the finite time control of the satellite model without uncertain-

ties. Then a third-order sliding mode scheme was used to design a discontinuous

control signal for robustification. It was shown that the composite control system

including the nominal and discontinuous signals, the pitch angle tracking and its

first two derivatives converged to zero in a finite time. The estimated values also

converged with actual values in finite time. In the closed-loop system precise pitch

attitude control was accomplished, despite uncertainties and disturbance input.

81

Page 94: Robust and Adaptive Attitude Control Of Spacecraft Using

Chapter 6

Conclusion

The objective of this research work was to design adaptive control laws to control

the attitude of spacecraft in orbit. The problems associated with spacecraft attitude

control mainly lies in the nonlinearity and uncertainty of the system. Due to this,

robust adaptive control systems are required to effectively regulate the satellite ori-

entation. This problem is addressed in this work and adaptive controllers have been

designed that circumvent the nonlinearity and uncertainty issues and accurately

control the attitude of the satellite. In Chapter 3, a robust finite time controller

is designed which handles nonlinearity and uncertainty using a nominal nonlinear

finite time continuous tracking control law and a higher order sliding mode control

law. A large uncertainty is assumed in the model of the system and a nominal

uncertainty in its parameters. Simulation was performed for various cases allow-

ing uncertainty in parameters, change in eccentricity of orbit, incidence angle. The

results showed the robustness and effectiveness of the system. However, this de-

sign assumes complete knowledge of the pitch angle derivatives which may not be

possible practically.

In Chapter 4, a nonlinear adaptive controller is designed which uses the inverse

control law and also a high gain estimator. The estimator is used to obtain the

values for the pitch angle derivative. The controller only requires the output pitch

angle of the system and uses the state estimates to produce the required trajectory.

82

Page 95: Robust and Adaptive Attitude Control Of Spacecraft Using

Here too, a large uncertainty in the system and nominal uncertainty in parameters

are assumed. This design is better in the aspect that it does not assume availability

of other states like the pitch angle derivatives and requires only the output of the

system for controlling the pitch angle trajectory. However, the estimator used is a

high gain estimator, which means that any minute disturbance in the signals fed

to the estimator can result in huge variations in the estimates. This makes the

controller extremely susceptible to noise in the system. Simulation was performed

for various cases allowing uncertainty in parameters, change in eccentricity of orbit,

incidence angle. The results showed the robustness and effectiveness of the system.

In Chapter 5, the robust finite time controller designed in Chapter 3 is modified

to replace the measured pitch angle derivatives with estimates from a robust exact

differentiator combined with a higher order sliding mode controller which together

act as an estimator. This controller is now more practical. Simulation was performed

for various cases allowing uncertainty in parameters, change in eccentricity of orbit,

incidence angle. The results showed the robustness and effectiveness of the system.

83

Page 96: Robust and Adaptive Attitude Control Of Spacecraft Using

APPENDIX

Nomenclature

a, e = Semi-major axis and eccentricity

B = Control input matrix

As, l = Control surface area and the moment arm

C′

s, Cs = Solar parameter

f , f0, fa = Nonlinear functions

f ∗, C∗

s , B∗ = Nominal values

G, pi = Controller gains

Ix, Iy, Iz = Moments of inertia of the satellite about x, y, z axes, respectively

K = (Ix − Iy)/Iz

Mg = Gravity gradient torque,

Ms,Md = Solar torque and disturbance torque

X, Y, Z = Inertial frame of reference

i = Inclination of the orbital plane with ecliptic plane

p = Solar radiation pressure, 4.65× 10−6Nm−2

ki, νi = Parameters in nominal control law

s = Sliding variable

un, ur, u = Control input signals

V = Lyapunov function

84

Page 97: Robust and Adaptive Attitude Control Of Spacecraft Using

Xb, Yb, Zb = Principal axes of the satellite

ξ1, (ξ2, ξ3) = Tracking error and its derivatives

ξa = Auxiliary variable

Xo, Yo, Zo = Orbital coordinates with Xo along the local vertical, XoYo defining

the orbital plane

α, αr = Pitch attitude angle, and pitch angle command

δi(i = 1, 2) = Control plate rotations

∆Cs,∆f,∆B = uncertain parameter, function, and matrix

θ = Orbital angle

λ = Pitch attitude angle of the satellite with respect to orbital coordinates

µ, (ν−11 , ν−1

2 , ν−13 ) = Degree of vector field, dilation

ρs = Reflectivity of the control surfaces

φ = Location of the Sun from the line of nodes

Ω = Mean orbital rate

85

Page 98: Robust and Adaptive Attitude Control Of Spacecraft Using

REFERENCES

1. Wright, J. L. (1992). Space sailing. Taylor & Francis US.

2. Radziewsky, V. V., & Chernikov, Y. A. (1965). The perturbations of satellite

orbits by solar radiation pressure. Bulletin of the Astronomical Institutes of

Czechoslovakia, 16, 1.

3. Mignard, F., & Henon, M. (1984). About an unsuspected integrable problem.

Celestial mechanics, 33 (3), 239-250.

4. Scheeres, D. J. (1999). Satellite Dynamics about small bodies: Averaged Solar

Radiation Pressure Effects1. Ann Arbor, 1001, 48109-2140.

5. Frye, W.E. & Stearns, E.V.B. (1959). Stabilization and attitude control of

satellite vehicles, ARS J. 22, 927-931

6. Sohn, R.L., Attitude stabilization by means of solar radiation pressure, ARS

J., 29,371-373

7. Hibbard, R. R. (1969). Attitude stabilization using focused radiation pressure.

ARS J. 31, 844-845

8. Ule, L. A. (1963). Orientation of spinning satellites by radiation pressure.

AIAA Journal, 1 (7), 1575-1578.

9. CROCKER, II, M. C. (1970). Attitude control of a sun-pointing spinning

spacecraft by means of solar radiation pressure. Journal of Spacecraft and

Rockets, 7 (3), 357-359.

86

Page 99: Robust and Adaptive Attitude Control Of Spacecraft Using

10. Falcovitz, J. (1966). Attitude control of a spinning sun-pointing spacecraft by

means of a grated solar sail. Technical Report CSR-TR-66-17, MIT Center

for Space Research, .

11. Modi, V. J., & Pande, K. C. (1973). Solar pressure control of a dual-spin

satellite. Journal of Spacecraft and Rockets, 10 (6), 355-361.

12. Modi, V. J., & Pande, K. C. (1973). Solar pressure induced librations of

spinning axisymmetric satellites. Journal of Spacecraft and Rockets, 10 (9),

615-617.

13. Pande, K. C. (1976). Attitude control of spinning spacecraft by radiation

pressure. Journal of Spacecraft and Rockets, 13 (12), 765-768.

14. Modi, V. J. & Pande, K. C. (1974). A bang-bang solar pressure attitude

control system. Journal of the Astronautical Sciences, 22,1-20.

15. Pande, K. C., Davies, M. S., & Modi, V. J. (1974). Time-optimal pitch control

of satellites using solar radiation pressure. Journal of Spacecraft and Rockets,

11 (8), 601-603.

16. Pande, K. C., & Venkatachalam, R. (1979). Semipassive pitch attitude control

of satellites by solar radiation pressure. Aerospace and Electronic Systems,

IEEE Transactions on, (2), 194-198.

17. Pande, K. C. & Venkatachalam, R. (1981). Solar pressure attitude stabiliza-

tion of earth-pointing spacecraft. Aerospace and Electronic Systems, IEEE

Transactions on, 17 (6), 748-756.

18. Pande, K. C., & Venkatachalam, R. (1980). Inertially fixed pitch stabilization

of satellites by solar radiation pressure. Acta Astronautica, 7 (2), 155-167.

19. Pande, K. C., & Venkatachalam, R. (1982). Optimal solar pressure attitude

control of spacecraftI: Inertially-fixed attitude stabilization. Acta Astronau-

tica, 9 (9), 533-540.

87

Page 100: Robust and Adaptive Attitude Control Of Spacecraft Using

20. Venkatachalam, R. & Pande, K. C. (1982). Optimal solar radiation pres-

sure attitude control of spacecraft - II: large angle attitude maneuvers. Acta

Astronautica, 9 (9),541-545.

21. Venkatachalam, R. (1993). Large angle pitch attitude maneuver of a satel-

lite using solar radiation pressure. Aerospace and Electronic Systems, IEEE

Transactions on, 29 (4), 1164-1169.

22. Singh, S. N. & Yim, W. (1996). Feedback linearization and solar pressure

satellite attitude control. Aerospace and Electronic Systems, IEEE Transac-

tions on, 32 (2), 732-740.

23. Stuck, B. W. (1980). Solar pressure three-axis attitude control. Journal of

Guidance, Control, and Dynamics, 3 (2), 132-139.

24. Modi, V. J., & Kumar, K. (1972). Attitude control of satellites using the solar

radiation pressure. Journal of Spacecraft and Rockets, 9 (9), 711-713.

25. Joshi, V. K., & Kumari, K. (1980). New solar attitude control approach for

satellites in elliptic orbits. Journal of Guidance, Control, and Dynamics, 3 (1),

42-47.

26. Kumar, K. (1988). A solar attitude controller for extending operational life-

span of communications satellites. Acta astronautica, 17 (1), 61-67.

27. Scull, J. R. Mariner IV revisited or the tale of the ancient mariner. (1969).

In Proceedings of the 20th International Astronautical Federation Congress,

Argentina, pp. 747-758.

28. Renner, U. (1979). Attitude control by solar sailing-a promising experiment

with OTS-2. European Space Agency Journal, 3, 35-40.

29. Shrivastava, S. K., & Hablani, H. B. (1979). Analysis of solar radiation pres-

sure induced coupled liberations of a gravity stabilized axisymmetric satellite.

Celestial mechanics, 20 (3), 297-313.

88

Page 101: Robust and Adaptive Attitude Control Of Spacecraft Using

30. Pande, K. C. (1976). Attitude control of spinning spacecraft by radiation

pressure. Journal of Spacecraft and Rockets, 13 (12), 765-768.

31. Modi, V. J., & Pande, K. C. (1973). Solar pressure control of a dual-spin

satellite. Journal of Spacecraft and Rockets, 10 (6), 355-361.

32. Pande, K. C., & Venkatachalam, R. (1980). Inertially fixed pitch stabilization

of satellites by solar radiation pressure. Acta Astronautica, 7 (2), 155-167.

33. Patel, T. R., Kumar, K. D., & Behdinan, K. (2008). Satellite attitude con-

trol using solar radiation pressure based on non-linear sliding mode control.

Proceedings of the Institution of Mechanical Engineers, Part G: Journal of

Aerospace Engineering, 222 (3), 379-392.

34. Patel, T. R., Kumar, K. D., & Behdinan, K. (2009). Variable structure con-

trol for satellite attitude stabilization in elliptic orbits using solar radiation

pressure. Acta Astronautica, 64 (2), 359-373.

35. Varma, S., & Kumar, K. D. (2010). Fault tolerant satellite attitude control

using solar radiation pressure based on nonlinear adaptive sliding mode. Acta

Astronautica, 66 (3), 486-500.

36. Singh, S. N., & Yim, W. (2005). Nonlinear adaptive spacecraft attitude con-

trol using solar radiation pressure. Aerospace and Electronic Systems, IEEE

Transactions on, 41 (3), 770-779.

37. Krstic, M., Kanellakopoulos, I., & Kokotovic, P. V. (1995). Nonlinear and

adaptive control design (Vol. 8). New York: John Wiley & Sons.

38. Lee, K. W., & Singh, S. N. (2009). Non-certainty-equivalent adaptive satellite

attitude control using solar radiation pressure. Proceedings of the Institution

of Mechanical Engineers, Part G: Journal of Aerospace Engineering, 223 (7),

977-988.

89

Page 102: Robust and Adaptive Attitude Control Of Spacecraft Using

39. Lee, K. W., & Singh, S. N. (2011). Attractive manifold-based adaptive solar

attitude control of satellites in elliptic orbits. Acta Astronautica, 68 (1), 185-

196.

40. Lee, K. W., & Singh, S. N. (2013). ”L1 adaptive attitude control of satellites

in elliptic orbits using solar radiation pressure. Proceedings of the Institution

of Mechanical Engineers, Part G: Journal of Aerospace Engineering.

41. Bhat, S. P., & Bernstein, D. S. (2005). Geometric homogeneity with appli-

cations to finite-time stability. Mathematics of Control, Signals and Systems,

17 (2), 101-127.

42. Bhat, S. P., & Bernstein, D. S. (2000). Finite-time stability of continuous

autonomous systems. SIAM Journal on Control and Optimization, 38 (3),

751-766.

43. Defoort, M., Floquet, T., Kokosy, A., & Perruquetti, W. (2009). A novel

higher order sliding mode control scheme. Systems & Control Letters, 58 (2),

102-108.

44. Hughes, P. C. (2012). Spacecraft attitude dynamics. DoverPublications. com.

45. Levant, A. (2001). Universal single-input-single-output (SISO) sliding-mode

controllers with finite-time convergence. Automatic Control, IEEE Transac-

tions on, 46 (9), 1447-1451.

46. Levant, A. (2003). Higher-order sliding modes, differentiation and output-

feedback control. International journal of Control, 76 (9-10), 924-941.

47. Esfandiari, F., & Khalil, H. K. (1992). Output feedback stabilization of fully

linearizable systems. International Journal of Control, 56 (5), 1007-1037.1007-

1037.

90

Page 103: Robust and Adaptive Attitude Control Of Spacecraft Using

48. Esfandiari, F., & Khalil, H. K. (1992). Output feedback stabilization of fully

linearizable systems. International Journal of Control, 56 (5), 1007-1037.

49. Khalil, H. K. (1996). Adaptive output feedback control of nonlinear systems

represented by input-output models. Automatic Control, IEEE Transactions

on, 41 (2), 177-188.

50. Zhang, R., & Singh, S. N. (2001). Adaptive output feedback control of an

aeroelastic system with unstructured uncertainties. Journal of Guidance,

Control, and Dynamics, 24 (3), 502-509.

91

Page 104: Robust and Adaptive Attitude Control Of Spacecraft Using

Graduate College

University of Nevada, Las Vegas

Lakshmi Srinivasan

Local Address:

4248 Spencer St, Apt 126

Las Vegas, Nevada 89119

Home Address:

50, Mandapam Road

Kilpauk, Chennai 600010

Tamil Nadu

India

Degrees:

Bachelors of Technology, Electrical Engineering, 2008

SASTRA University, Thanjavur, India

Thesis Title:

Robust and Adaptive Attitude Control of Spacecraft Using Solar Radiation Pres-

sure

Thesis Examination Committee:

Chairperson, Dr. Sahjendra Singh, Ph.D.

Committee Member, Dr. Ebrahim Saberinia , Ph.D.

92

Page 105: Robust and Adaptive Attitude Control Of Spacecraft Using

Committee Member Dr. Venkatesan Muthukumar, Ph.D.

Graduate Faculty Representative, Dr. Woosoon Yim, Ph.D.

93