140
AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT-TOLERANT CONTROL IN PRESENCE OF UNCERTAINTIES AND INPUT CONSTRAINTS A Dissertation Submitted to the Faculty of Purdue University by Shreekant Gayaka In Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy December 2008 Purdue University West Lafayette, Indiana

AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT-TOLERANT

CONTROL IN PRESENCE OF UNCERTAINTIES AND INPUT CONSTRAINTS

A Dissertation

Submitted to the Faculty

of

Purdue University

by

Shreekant Gayaka

In Partial Fulfillment of the

Requirements for the Degree

of

Doctor of Philosophy

December 2008

Purdue University

West Lafayette, Indiana

Page 2: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

ii

TABLE OF CONTENTS

Page

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Need for actuator fault-tolerant control . . . . . . . . . . . . . . . . 11.2 Existing techniques: Scope and Limitations . . . . . . . . . . . . . . 2

1.2.1 Literature review . . . . . . . . . . . . . . . . . . . . . . . . 21.2.2 Desired features of a New Fault-Tolerant Controller . . . . . 4

1.3 An Adaptive Robust Approach to Actuator Fault-Tolerant Control 41.4 Assumption on Input Saturation . . . . . . . . . . . . . . . . . . . . 5

1.4.1 Input saturation and its effect of actuator fault-tolerant control 51.4.2 Robust global stabilization of an integrator chain . . . . . . 61.4.3 Saturated Adaptive Robust Actuator Fault-Tolerant Control 6

1.5 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71.6 Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2 Accommodation of Unknown Actuator Faults Using Output Feedback basedAdaptive Robust Control . . . . . . . . . . . . . . . . . . . . . . . . . . . 102.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102.2 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . 122.3 Output Feedback based ARFTC . . . . . . . . . . . . . . . . . . . . 14

2.3.1 Observer Canonical Form . . . . . . . . . . . . . . . . . . . 142.3.2 State Estimation . . . . . . . . . . . . . . . . . . . . . . . . 202.3.3 Parameter Projection . . . . . . . . . . . . . . . . . . . . . . 222.3.4 Controller Design . . . . . . . . . . . . . . . . . . . . . . . . 23

2.4 Simulation Example: Linearized Boeing 747 Model . . . . . . . . . 332.4.1 Detailed ARFTC Controller design . . . . . . . . . . . . . . 36

2.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

3 Output Feedback based Adaptive Robust Fault-Tolerant Control for a Classof Uncertain Nonlinear Systems . . . . . . . . . . . . . . . . . . . . . . . 433.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 433.2 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . 443.3 Output Feedback based ARFTC . . . . . . . . . . . . . . . . . . . . 46

3.3.1 State estimation . . . . . . . . . . . . . . . . . . . . . . . . . 473.3.2 Parameter Projection . . . . . . . . . . . . . . . . . . . . . . 503.3.3 Controller Design . . . . . . . . . . . . . . . . . . . . . . . . 50

Page 3: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

iii

Page3.4 Simulation Example: A Nonlinear Hypersonic Aircraft Model . . . . 593.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4 A Backstepping based Approach to Robust Global Stabilization of a Chainof Integrators with Input Saturation . . . . . . . . . . . . . . . . . . . . . 684.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 684.2 Motivation and Problem Formulation . . . . . . . . . . . . . . . . . 714.3 Main Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

4.3.1 Convergence to Unsaturated Region . . . . . . . . . . . . . . 754.3.2 Controller Parameter Selection . . . . . . . . . . . . . . . . 80

4.4 Simulation Example: Third Order Integrator Chain . . . . . . . . . 874.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

5 Saturated Adaptive Robust Actuator Fault-Tolerant Control for FeedbackLinearizable Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 955.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 955.2 Problem Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . 975.3 Adaptive Robust Actuator Fault-Tolerant Control . . . . . . . . . . 99

5.3.1 Parameter Estimation . . . . . . . . . . . . . . . . . . . . . 1035.3.2 Controller design . . . . . . . . . . . . . . . . . . . . . . . . 1045.3.3 Controller Parameter Selection . . . . . . . . . . . . . . . . 1115.3.4 Asymptotic Stability . . . . . . . . . . . . . . . . . . . . . . 118

5.4 Simulation Example: Nonlinear Hypersonic Aircraft Model . . . . . 1205.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125

6 Conclusions and Future Work . . . . . . . . . . . . . . . . . . . . . . . . 1266.0.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 1266.0.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

LIST OF REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

Page 4: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

iv

LIST OF FIGURES

Figure Page

2.1 Reference tracking, control signals and tracking error for MRAC versusARC based fault-tolerant schemes in absence of disturbances . . . . . . 40

2.2 Reference tracking, control signals and tracking error for MRAC versusARC based fault-tolerant schemes in presence of disturbances . . . . . 41

3.1 Structure of ARC and RAC based fault-tolerant controllers . . . . . . . 53

3.2 Reference tracking, control signals and tracking error for RAC versus ARCbased fault-tolerant schemes in absence of disturbances . . . . . . . . . 65

3.3 Reference tracking, control signals and tracking error for RAC versus ARCbased fault-tolerant schemes in presence of disturbances . . . . . . . . . 66

4.1 Saturation function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.2 Comparative results for stabilization in absence of disturbances . . . . 91

4.3 Comparative results for stabilization in presence of disturbance . . . . 92

4.4 Tracking in presence of large disturbance . . . . . . . . . . . . . . . . . 93

5.1 Saturation function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

5.2 Comparative results for stabilization in absence of disturbances . . . . 123

5.3 Comparative results for stabilization in presence of disturbance . . . . 124

Page 5: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

v

ABSTRACT

Gayaka, Shreekant Ph.D., Purdue University, December 2008. An Adaptive RobustApproach to Actuator Fault-Tolerant Control in Presence of Uncertainties and InputConstraints . Major Professor: Bin Yao, School of Mechanical Engineering.

In this work, we develop adaptive robust schemes for actuator fault-tolerant con-

trol in presence of uncertainties and input saturation. The type of faults considered

in the present work encompass hardover-failure, loss in efficiency and stuck actua-

tors. The two chief ways in which the system performance can degrade following an

actuator-fault are undesirable transients and unacceptably large steady-state tracking

errors. Adaptive control based schemes are ideal for handling the jump in parameter

values following an actuator fault, and can guarantee good final tracking accuracy.

However, such schemes may not be able to suppress the transients due to sudden

change in system parameters. Furthermore, the performance of adaptive control

based schemes deteriorate significantly in presence of unknown modeling errors and

disturbances. Robust control based schemes, on the other hand, can guarantee desired

transient response due to the sudden jump in system parameters and attenuate the

effect of modeling uncertainties on the tracking error. But, in face of large paramet-

ric uncertainties due to actuator faults, the final tracking accuracy of robust control

based schemes may degrade as they cannot reduce the extent of parametric uncer-

tainties. In the present work, we claim that an adaptive robust fault-tolerant control

scheme can solve both the problems, as it seamlessly integrates adaptive and robust

control design techniques. Comparative simulation studies are performed using linear

and nonlinear aircraft models to illustrate the superior performance of the proposed

scheme over robust MRAC and robust backstepping based adaptive control designs

respectively.

Page 6: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

vi

One of the standard assumptions made in the design of adaptive fault-tolerant

control is that the healthy actuators have sufficient control authority despite faults

to recover desired closed-loop performance. In reality, however, the controller could

generate large control commands to suppress the undesired transients, leading to ac-

tuator saturation. Furthermore, in direct adaptive schemes, the estimator may fail

to generate reliable parameter estimates due to saturation. This could further de-

grade the performance of a actuator fault-tolerant control. As a first step towards

developing an approach which can deal with input constraints, we propose a concep-

tually different approach for global stabilization of a chain of integrators. A novel and

elegant approach to solve this problem was proposed by Teel [1] using saturation func-

tions and coordinate transformation. With Teel’s work as foundation, many results

have been proposed to improve the performance of tracking/stabilizing controllers for

chain of integrators. Naturally, all such approaches also inherited the limitations of

Teel’s approach. Most importantly, in presence of uncertainties, such a transforma-

tion would considerably shrink the region where the controller is unsaturated, and in

some cases, may even render the task of designing a stabilizing controller impossible.

We combine the backstepping based design with saturation functions to develop a

simple controller which does not rely on coordinate transformation and meets all the

desired objectives. Furthermore, necessary and sufficient conditions for the existence

of the proposed control law, as well as a systematic way of choosing the controller

parameters is also presented. Comparative simulation studies are performed on a

third order integrator chain which shows the effectiveness of the proposed scheme.

Finally, an actuator fault-tolerant controller is designed which combines the pro-

posed backstepping based saturation functions approach with a least-square estima-

tor. The indirect scheme ensures that the adaptation mechanism is not affected

adversely due to actuator saturation. Simulation studies performed on a hypersonic

aircraft model demonstrate the effectiveness of the proposed scheme in addressing

actuator faults in presence of input constraints.

Page 7: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

1

1. INTRODUCTION

1.1 Need for actuator fault-tolerant control

On 4 October 1992, El Al Flight 1862, a Boeing 747 cargo plane of the Israeli airline

El Al, crashed into the Groeneveen and Klein-Kruitberg flats in the Bijlmermeer

neighbourhood of Amsterdam, Netherlands. This unfortunate incident resulted in

43 casualties, consisting of the plane’s crew of three and a non-revenue passenger in

a jump seat, plus 39 persons on the ground. Among many factors that caused this

disaster, one of the chief reasons was the damaged control surfaces on the right wing.

On 22 September 1981, in a similar incident, Easter Airlines L-1011 flying from

Newark, New Jersey to San Juan, Puerto Rico, suffered a massive failure of its number

two (tail) engine. A shrapnel from the engine damaged 3 out of 4 hydraulic systems

in the tail structure. But, the fluid which remained pressurized in that 4th system

enabled the captain to land the plane safely at John F. Kennedy International Airport,

with remaining healthy components of the control system. There were no injuries.

Thus, the additional 4th hydraulic control system saved the plane and all on board.

Unfortunately, there are more examples like the El Al Flight where final outcome

was a disaster, and few success stories like that of Eastern Airlines. This naturally

raises the question - after a failure, is it at all possible to use the remaining healthy

components of the control system in a fashion to avoid complete breakdown? Many

researchers have tried to simulate the conditions which lead to such incidents, and

tried to come up with schemes which can avoid such disasters. In [2], it was concluded

that had a fault-tolerant control scheme been in place, the ill fate of the El Al Flight

could have been avoided. One of the reasons Easter Airlines L-1011 could be saved is

Page 8: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

2

because the captain and the crew had sufficient reaction time, (i.e., time required to

detect, isolate and identify the fault) to reconfigure the remaining actuators to land

the aircraft safely. From the two incidents cited above, it can be concluded that the

key to avoid a catastrophic outcome is to use the working actuators effectively. Ac-

tuator fault-tolerant controller precisely accomplishes this by exploiting the available

redundancy in modern machines e.g., multiple control surfaces on an aircraft.

1.2 Existing techniques: Scope and Limitations

1.2.1 Literature review

From the preceding section, it should be evident that as complexity of modern

day machines increase, it may not be safe to rely on human interference to detect,

isolate and compensate faults. Ensuring safety and reliability in machines with large

number of subsystems and components call for advanced algorithms. Thus, in the

past few decades, many researchers have focused their attention to this important

problem, leading to a steady increase in literature devoted to actuator fault-tolerant

control. The effect of such actuator faults on the system dynamics can be captured

as unknown, sudden change in system parameters, and it can degrade the system

system performance in two chief ways: (a) it can cause large transients, which may

eventually cause instability and, (b) it may result in unacceptably large steady-state

tracking errors.

Most of the available literature on this topic can broadly be categorized in two

groups - adaptive control and robust control based designs. Adaptive schemes are a

promising approach to deal with such failures as it can learn the change in system

parameters by virtue of their on-line learning capability. Not surprisingly, many

adaptive schemes have been developed to solve this problem. Tao et al. proposed

a model reference adaptive control (MRAC) based scheme for unknown actuator

failure compensation for linear systems in [3]. They further extended their direct

fault-compensation scheme to various classes of nonlinear systems in [4], [5] using

Page 9: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

3

backstepping based adaptive control. Another popular adaptive approach to solve

this problem is multiple model adaptive control (MMAC), switching and tuning [6].

But, none of these papers considered unstructured or non-parametric uncertainties

e.g., unknown nonlinearities and disturbances, which can be a limiting-factor of the

achievable system performance. Robust adaptive control (RAC) based schemes, that

can guarantee boundedness of closed loop signals in presence of unknown modeling

uncertainties and disturbances, were investigated in [5,7,8]. But, RAC is a variant of

adaptive control and lacks two desirable properties inherent to robust control based

techniques. First, there is no convenient and transparent way to attenuate the effect

of non-parametric uncertainties, like external disturbances, on system response and

steady-state tracking error. Second, such techniques are not well suited to suppress

the undesirable transients following a sudden change in system parameters due to

unknown actuator faults. As poor transients in adaptive control based schemes can

be attributed to the learning phase of the controller, it may appear that increasing

the adaptation gain can improve the transient response as it speeds up the learning

process. In fact, this result has been claimed in many articles (see [9], [10]). However,

as projection type of robustness modifications are present in RAC based techniques

to avoid parameter drift, the use of high adaptation gains may cause the estimated

parameters to bounce back and forth between the present upper and lower limits.

This could introduce a high frequency component in the control signal, which may

ultimately excite the high-frequency ignored dynamics. Thus, even though all signals

can be shown to be bounded, obtaining guaranteed transient response in a RAC

framework still remains a challenging problem. Robust control based schemes [11–

14], on the other hand, have guaranteed transient response in presence of various

uncertainties. Furthermore, the effect of such disturbances can be attenuated to any

desired extent on the steady-state tracking error. However, in face of large parametric

uncertainties due to actuator faults, such schemes will either lead to undesirable

control input chatter or poor steady-state tracking error due to smoothing techniques.

Page 10: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

4

1.2.2 Desired features of a New Fault-Tolerant Controller

A critical review of the existing literature reveals that adaptive and robust control

based fault-tolerant schemes can each address a part of the problem, but not all the

issues associated with actuator fault-tolerant control when used individually. Thus,

a fault-tolerant control should possess the desirable properties of both the schemes

in order to satisfactorily solve the problem. More specifically, the control objectives

a new fault-tolerant control scheme are

1. ability to handle large parametric uncertainties with desired transients

2. ensure small steady-state tracking error, despite actuator faults and non-parametric

uncertainties like external disturbances and unknown nonlinearities

3. stable controller with desired closed-loop properties, in case the uncertainties

disrupt the functioning of estimation module

1.3 An Adaptive Robust Approach to Actuator Fault-Tolerant Control

Given the need for stability in safety critical missions, the large parametric un-

certainties introduced due to unknown actuator failures and the inherent limitations

of adaptive control, the idea of safe adaptive control is coming to forefront. Safe

adaptive control ensures certain stability properties even without adaptation [15,16].

ARC based schemes have already resolved this issue [17,18] and may be classified as

the so-called safe adaptive control. In the proposed ARC based fault-tolerant scheme,

adaptive and robust control designs are integrated seamless fashion. This allows all

control objectives, enlisted in the preceding section, to be achieved. In fact, switching

the adaptation off at any instant converts the adaptive robust controller into a de-

terministic robust controller with guaranteed transient performance. Moreover, the

design procedure allows us to calculate explicit upper bound for tracking errors over

the entire time history in terms of certain controller parameters and achieve prespec-

Page 11: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

5

ified final tracking accuracy. Thus, ARC based fault-tolerant schemes are natural

choices for safety sensitive systems over conventional adaptive and robust schemes.

1.4 Assumption on Input Saturation

An important assumption in the design of fault-tolerant control in the literature

cited above is that the working actuators have sufficient authority to accommodate the

fault and fulfill the desired control objective. In reality, however, as actuators fail, it

may not be possible to achieve all performance criteria, as there is only limited control

authority. Especially, in safety critical missions like flight control systems, saturation

can not only deteriorate the closed loop performance significantly [19], but also lead

to catastrophic outcomes in the worst case scenario. A good example of harmful

effects of saturation is the crash of JAS 39 Gripen in 1993. Post analysis revealed

that there were no component malfunction which could have caused the accident. It

was concluded after further analysis that Pilot Induced Oscillation (PIO) caused the

aircraft to crash. The PIO was a direct consequence of overlooking the saturation

limits in the design of the control systems. The situation becomes much more grave

when there are actuator faults.

1.4.1 Input saturation and its effect of actuator fault-tolerant control

In order to understand how actuator faults and actuator saturation negatively

reinforce their concomitant destructive effects, we must first understand the inter-

play of the actuator fault-tolerant control and the actuator faults. Following actuator

faults, the tracking errors tend to stray away from zero due to undesired transients.

If the controller is designed without any regard to saturation, it may generate large

control signals to attenuate the effect of undesired transients, thereby saturating the

actuators. As the actuators saturate, the error further increases, which in turn in-

creases the commanded control input, so on and so forth. This may ultimately lead

the closed-loop system to become unstable. Second, in direct adaptive schemes, the

Page 12: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

6

state estimation error – the difference in actual state and the reconstructed state

which uses estimated parameters, correlates directly to the mismatch in estimated

parameters. This allows the actuator-fault tolerant controller to learn the change in

system parameters due to faults, and adjust the model compensation accordingly,

thereby resulting in small steady-state tracking error. However, in presence of satu-

ration, in addition to the parameter mismatch, the difference between the actual and

commanded control input is also responsible for the state-estimation error. Hence, the

adaptation can no longer guarantee reliable estimates for the unknown parameters.

As a first step towards designing a saturated adaptive robust actuator fault-tolerant

control, we investigate the design of saturated controller for an integrator chain with

input disturbances.

1.4.2 Robust global stabilization of an integrator chain

Controller design in presence of input saturation is a theoretically challenging

problem with deep practical implications. The significance of this problem cannot be

overemphasized, as proper functioning of any control system depends on actuators,

all of which have physical limitations e.g., limited valve opening, available voltage

for servo-motors etc. Especially, in safety critical missions like flight control systems,

input saturation can not only deteriorate the closed loop performance significantly

[19], but also lead to catastrophic outcomes in the worst case scenario [20]. Not

surprisingly, many design techniques which account for this input nonlinearity have

been proposed (see [21], [22], [23] and the references there in).

1.4.3 Saturated Adaptive Robust Actuator Fault-Tolerant Control

From the preceding discussion, it should be clear that if a fault-tolerant scheme

can be designed which (a) relaxes the performance criteria when the states deviate

far away from zero following an actuator fault, and (b) ensures reliable adaptation

in spite of actuator faults, then the actuator fault-tolerant scheme will become more

Page 13: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

7

practical. In the proposed approach, bounded feedback control laws are designed

such that performance is compromised when the error-variables are far away from

zero and the actuator is saturated, and more emphasis is put on bringing the error-

variables in a region where the controller is unsaturated. This implies unrealistically

high control inputs are not demanded of the saturating actuators after a fault. Once

within the unsaturated region, the desired performance can be recovered. To tackle

the second problem, we use an indirect adaptive scheme with controller and estimator

modularity. Many schemes have been proposed in the literature to avoid the harmful

effects of saturation on adaptation. For example in [24,25], a modified tracking error

was used to avoid the harmful effect of saturation on adaptation. Pseudo Control

Hedging (PCH) based technique was used in [26], which involves altering the com-

manded reference to a level that allows controller operation without saturation. Such

techniques unnecessarily increase the complexity, without significantly improving the

performance. An indirect scheme, in contrast, does not require such modifications,

as the model structure of the system does not change despite saturation. This facil-

itates an uninterrupted and reliable estimation of parameters. Surprisingly, indirect

adaptive schemes, which are the simplest and surest way of avoiding the effect of

saturation on adaptation, have largely been overlooked.

1.5 Contributions

In this dissertation, standard assumptions and design approaches for actuator

fault-tolerant control and saturated control problem were carefully examined to ex-

plore their drawbacks in terms of achievable performance. A quantitative analysis of

the shortcomings naturally lead us to consider these problems from a fundamentally

different viewpoint, which resulted in novel control design techniques. Following prob-

lems were recognized as fundamental issues which have not been effectively addressed

in the existing literature

Page 14: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

8

1. accommodation of unknown actuator faults and the limitations of existing tech-

niques in effectively addressing these issues,

2. degradation in performance of an actuator fault-tolerant control in presence of

saturation,

3. conservativeness of the conventional coordinate transformation based approach

for controller design in presence of saturation and input disturbance.

Following schemes have been developed in this dissertation to solve the aforemen-

tioned problems,

1. proposed a performance oriented adaptive robust control (ARC) based fault-

tolerant design which can accommodate actuator faults with guaranteed tran-

sient response and small steady-state tracking error,

2. developed a conceptually different approach to design a globally stable controller

for integrator chain with input disturbance,

3. designed an indirect adaptive robust actuator fault-tolerant control scheme

which explicitly considers the effect of actuator saturation.

1.6 Organization

This dissertation is divided in six chapters. The contents of chapters are summa-

rized as follows:

1. Chapter 1, Introduction: motivates the study of actuator fault-tolerant control.

This chapter gives an overview of the existing techniques, which leads to the

research objectives and problem formulation. Research approach and some

features of the techniques proposed in this dissertation are also discussed.

2. Chapter 2, Accommodation of Unknown Actuator Faults Using Output Feedback

based Adaptive Robust Control: presents an output feedback based adaptive ro-

Page 15: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

9

bust fault-tolerant scheme to accommodate unknown actuator faults. Compar-

ative simulation studies are also presented to show the superior performance of

the proposed design.

3. Chapter 3, Output Feedback based Adaptive Robust Fault-Tolerant Control for

a Class of Uncertain Nonlinear Systems: addresses the problem of unknown

actuator fault-accommodation for a class of nonlinear systems. Comparative

simulation studies demonstrate that the shortcomings of a robust adaptive back-

stepping based design can be overcome by using the proposed scheme.

4. Chapter 4, A Backstepping based Approach to Robust Global Stabilization of a

Chain of Integrators with Input Saturation: proposes a novel way of stabilizing

an integrator chain in presence of input disturbances. It is shown through sim-

ulations that improved disturbance rejection properties and faster convergence

rate can be achieved by using the proposed controller.

5. Chapter 5, Saturated Adaptive Robust Actuator Fault-Tolerant Control for Feed-

back Linearizable Systems: addresses the problem of actuator fault-tolerant

control in presence of uncertainties and input saturation. The results obtained

indicate the effectiveness of the proposed scheme in accommodating actuator

faults, despite input saturation.

6. Chapter 6, Conclusions and Future Work: discusses the results and summarizes

the thesis work. Some ideas for future research are also presented.

Page 16: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

10

2. ACCOMMODATION OF UNKNOWN ACTUATOR FAULTS USING

OUTPUT FEEDBACK BASED ADAPTIVE ROBUST CONTROL

2.1 Introduction

In this chapter, we solve the problem of fault accommodation for unknown actua-

tor failures in an uncertain linear system with unknown parameters and subjected to

bounded disturbances. This problem is of prime importance as conventional feedback

control design for complex systems may result in unacceptable degradation of perfor-

mance or even instability in the event of actuator failures. The faults are modeled as

actuators experiencing bounded disturbance about an unknown constant value, loss

in actuator efficiency or a combination of the two. We do not assume the knowledge

of failed actuators or the instant of failure. Fortunately, adaptive schemes, by virtue

of their on-line learning capability can bypass this problem. Consequently, many

adaptive schemes have been developed to solve this problem.

A novel approach for solving the problem of unknown actuator failure compen-

sation was proposed in [3] for linear systems. But, the schemes use conventional

MRAC [3] which suffer from poor transients during the learning phase and have diffi-

culty in ensuring tracking error bounds in absolute sense in the presence of exogenous

disturbances. Robust schemes for actuator fault accommodation, which can handle

such disturbances and unstructured uncertainties with guaranteed transient perfor-

mance, include LMI based techniques [27] and sliding mode control (SMC) based

approaches [13]. But, robust control based direct fault-tolerant schemes lack the abil-

ity to learn the large extent of parametric uncertainty introduced by failed actuators.

Thus, such schemes rely on high gain to attenuate the effect of uncertainties on the

steady-state tracking error. But, as all actuators have limited bandwidth, this may

lead to degraded tracking performance. One approach which potentially alleviates

Page 17: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

11

these problems is multiple model adaptive control (MMAC), switching and tuning.

In [28], the authors propose a MMAC based actuator fault accommodation scheme.

The proposed scheme, however, did not take into account modeling uncertainties and

used full state-feedback.

In the present work, we develop an output feedback based Adaptive Robust Fault-

Tolerant Control (ARFTC) scheme for accommodation of unknown actuator faults.

Actuator faults manifest themselves as jump in parameter values in the system model

and pose a twofold problem. First, it can introduce undesirable transients due to the

jump in parameter values and second, it can lead to unacceptable steady-state track-

ing error. Robust schemes can address the first issue and adaptive control can deal

with the second one, but none of them can resolve both issues simultaneously. In

the present work, we claim that ARFTC can tackle both the problems at the same

time. This is to be expected as an Adaptive Robust Control (ARC) based fault-

tolerant scheme combines robust and adaptive control design philosophies seamlessly

(see [17], [18]). In fact, parameter projection is used not only to guarantee the bound-

edness of the estimated parameters but, also in the design of the robust component

of the control law. Thus, the known extent of the jumps in parameter values is

incorporated in the controller design to guarantee desired transient response. Fur-

thermore, as ARFTC design puts more emphasis on the underlying robust control law

design, switching the adaptation off at any instant converts the ARFTC controller

into a robust controller which guarantees desired transient performance and certain

steady-state accuracy. This feature is extremely desirable for safety critical systems.

The adaptation mechanism of the proposed controller, on the other hand, tries to

learn the change in system parameters due to actuator faults and adjusts the model-

compensation of the control law accordingly. By doing so, an improved steady-state

tracking performance – asymptotic tracking in the presence of parametric uncertain-

ties only – is achieved. Also, it is seldom the case that a parametric model can

completely capture the effect of fault. In the proposed approach, an unknown time

varying disturbance is added to the parametric fault model, and the underlying ro-

Page 18: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

12

bust controller is designed in a way to attenuate its effect. This considerably enlarges

the class of faults that can be addressed using the proposed scheme. Furthermore, in

addition to robust feedback for dealing with unstructured uncertainties, we also use

a priori information about the exogenous disturbances to achieve better disturbance

rejection properties.

In this chapter, a linearized Boeing 747 model is used to show the effectiveness of

the proposed scheme in dealing with unknown actuator faults. Additionally, compar-

ative studies performed with respect to a robust MRAC based fault-tolerant design

clearly demonstrate the superior performance in terms of guaranteed transient re-

sponse and steady-state tracking error.

2.2 Problem Statement

In the present work, we consider systems which can be represented in the input-

output form as follows,

y(t) =

q∑

j=1

Bj(s)

A(s)uj(t) +

D(s)

A(s)∆(y, t) + dy(t) (2.1)

where, A(s) = sn + an−1sn−1 + . . . + a1s + a0, Bj(s) = bjms

m + . . . + bj1s + bj0 and

D(s) = dlsl + . . . + d1s + d0, l ≥ m. Note that l ≥ m implies we do not restrict

ourselves to matched uncertainties only. The plant parameters ai and bi are unknown

constants. The coefficients di corresponding to the disturbance distribution are as-

sumed to be known but, the results can be readily extended to the case where they are

unknown constants. uj(t) represents the actual output from the jth actuator. dy(t)

represents the output disturbance, and ∆(y, t) represents any disturbance coming

from the intermediate channels of the plant. An implicit assumption in the system

representation (2.1) is,

A1: The relative degree ρ = n−m is known and same for any input uj.

Page 19: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

13

In this work, we will consider actuator failures which can be modeled as,

uj(t) =

uj + uj(t), ∀t ≥ Tf ,

if jth actuator gets stuck at Tf

ηjju∗j(t), ∀t ≥ Tf ,

if jth actuator loses efficiency at Tf

(2.2)

where u∗j(t) represents the control command to the jth actuator, uj is an unknown

constant value at which the actuator gets stuck, uj(t) is a bounded unknown signal

about uj, Tf is the unknown instant of failure and ηjj represents actuator loss in

efficiency with ηjj ∈ [(ηjj)min, 1], (ηjj)min ≥ 0. Note that conventional MRAC based

schemes are not well suited to handle time-varying disturbances like uj(t). Hence,

the class of actuator faults addressed here is larger and more practical than what is

typically considered. It is noted that the fault model presented here encompass both

lock-in-place and hard-over failure actuator faults. Without actuator redundancy,

actuator faults cannot be accommodated. This is formally stated in the following

assumption,

A2: System (2.1) is such that the desired control objective can be fulfilled with

up to q − 1 stuck actuators and any number of actuators with loss in efficiency.

Now, the problem we attempt to solve in this work can be stated precisely as

follows. For the system described by (2.1), subjected to unknown actuator failures

(2.2) and bounded disturbances, the goal is to design an output feedback control

law such that the steady-state output tracking error converges exponentially to an

acceptable bound and has a guaranteed transient performance.

In addition to actuator fault compensation, it is also desirable that the closed-

loop system posses good disturbance rejection properties. In the present approach,

such properties are achieved by using robust feedback as well as explicitly tak-

ing into account ∆(y, t) (see [29]). We use prior information about the nature of

disturbance to construct a nominal disturbance model ∆n(y, t) = q(y, t)T c, where

q(y, t) = [qp(y, t), . . . , q1(y, t)]T ∈ R

p represents the vector of known basis shape func-

tions and c = [cp, . . . , c1]T represents the vector of unknown magnitudes. Thus, the

Page 20: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

14

disturbance can be represented as, ∆ = ∆n + ∆, where ∆ is the modeling error.

Adaptation will be used to compensate for the effect of ∆n on the output tracking

performance and ∆ will be dealt with via certain robust feedback for robust perfor-

mance.

2.3 Output Feedback based ARFTC

2.3.1 Observer Canonical Form

In the present work, we will assume that control commands to all actuators are

the same i.e., u∗1 = . . . = u∗q = u∗. With this choice of control inputs, and the fault

model described by (2.2), the healthy and faulty actuators can be parameterized in

the following way,

uj(t) = ηjj(1 − σjj)u∗(t) + σjj(uj + uj(t))

⇒ uj(t) = κju∗(t) + σjj uj + σjjuj(t)) (2.3)

where

σjj =

0 before jth actuator gets stuck

1 after jth actuator gets stuck

ηjj =

1 before jth actuator loses efficiency

[(ηjj)min, 1] after jth actuator loses efficiency

κj = ηjj(1 − σjj)

Page 21: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

15

Now, system (2.1) can be rewritten as,

y(t) =

q∑

j=1

κj

Bj(s)

A(s)u∗j(t) +

q∑

j=1

σjj

Bj(s)

A(s)uj

+

q∑

j=1

σjj

Bj(s)

A(s)uj(t) +

D(s)

A(s)∆(y, t) + dy(t) (2.4)

=bfms

m + . . .+ bf1s+ bf0sn + an−1sn−1 + . . .+ a1s+ a0

u∗(t)

+bfms

m + . . .+ bf1s+ bf0sn + an−1sn−1 + . . .+ a1s+ a0

1

+

q∑

j=1

σjj

bjmsm + . . .+ bj1s+ bj0

sn + an−1sn−1 + . . .+ a1s+ a0uj(t)

+dls

l + . . .+ d1s+ d0

sn + an−1sn−1 + . . .+ a1s+ a0∆ + dy(t) (2.5)

where,

bfi = κ1b1i + · · ·+ κqbqi, i = 0, · · · , m

bfj = σ11u1b1j + · · · + σqquqbqj j = 0, · · · , m (2.6)

Note that as a consequence of assumption A1 and A2, bfm 6= 0 for any fault.

Before we proceed to present the observer canonical form of the system (2.5), we

recall the standard result that for a transfer function

G(s) =βms

m + ... + β0

sn + αn−1sn−1 + ... + α0

(2.7)

the observer canonical form is given by

x =

−αn−1 1 0 . . . 0

−αn−2 0 1 . . . 0...

. . ....

.... . .

...

−α0 0 . . . 0 1

x+

0...

βm

...

β0

u (2.8)

Page 22: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

16

This is a standard result and can be found in any book on linear systems like [30].

Now we present an observer canonical realization of the above input-output model

which is more suitable for the controller design technique presented here,

y(t) =

q∑

j=1

κjBj(s)

A(s)u∗(t) +

q∑

j=1

σjjBj(s)

A(s)uj +

D(s)

A(s)∆(y, t)

+

q∑

j=1

σjjBj(s)

A(s)uj(t) + dy(t)

Now, expanding each term in the transfer function, we have

q∑

j=1

κjBj(s) =

q∑

j=1

κj(bjmsm + . . .+ bj1s+ bj0)

=

(

q∑

j=1

κjbjm

)

sm + . . .+

(

q∑

j=1

κjbj1

)

s+

(

q∑

j=1

κjbj0

)

= bfmsm + . . .+ bf1s+ bf0 , using the definition of bfj

q∑

j=1

σjjBj(s)uj =

q∑

j=1

σjj(bjmsm + . . .+ bj1s+ bj0)uj

=

(

q∑

j=1

σjjbjmuj

)

sm + . . .+

(

q∑

j=1

σjjbj1uj

)

s+

(

q∑

j=1

σjjbj0uj

)

= bfmsm + . . .+ bf1s+ bf0 , using the definition of bfj

D(s) = dlsl + . . .+ d1s+ d0

q∑

j=1

σjjBj(s) =

q∑

j=1

σjj(bjmsm + . . .+ bj1s+ bj0)

A(s) = sn + an−1sn−1 + . . .+ a1s+ a0

which leads to

y(t) =bfms

m + . . .+ bf1s+ bf0sn + an−1sn−1 + . . .+ a1s+ a0

u∗(t) +bfms

m + . . .+ bf1s+ bf0sn + an−1sn−1 + . . .+ a1s+ a0

1

+dls

l + . . .+ d1s+ d0

sn + an−1sn−1 + . . .+ a1s+ a0

∆(y, t)

+

q∑

j=1

σjj

bjmsm + . . .+ bj1s+ bj0

sn + an−1sn−1 + . . .+ a1s+ a0uj(t) + dy(t)

= Y1 + Y2 + Y3 +

q∑

j=1

Y j4 + dy(t)

Page 23: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

17

The corresponding observer canonical form for TF =∑q

j=1κjBj(s)

A(s)is given by

Y11

Y12

...

Y1n

=

−an−1 1 0 · · · 0

−an−2 0 1 · · · 0...

.... . .

. . ....

−a0 0 0 0 0

Y11

Y12

...

Y1n

+

0...

bfm...

bf0

u∗

Y1 = Y11

For TF =∑q

j=1σjjBj(s)uj

A(s)

Y21

Y22

...

Y2n

=

−an−1 1 0 · · · 0

−an−2 0 1 · · · 0...

.... . .

. . ....

−a0 0 0 0 0

Y21

Y22

...

Y2n

+

0...

bfm...

bf0

u0

Y2 = Y21

where u0 = 1.

For TF = D(s)A(s)

,

Y31

Y32

...

Y3n

=

−an−1 1 0 · · · 0

−an−2 0 1 · · · 0...

.... . .

. . ....

−a0 0 0 0 0

Y31

Y32

...

Y3n

+

0...

dl

...

d0

Y3 = Y31

Page 24: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

18

And similarly for TF =σjjBj(s)

A(s), for j = 1, 2, ..., q

Y j41

Y j42

...

Y j4n

=

−an−1 1 0 · · · 0

−an−2 0 1 · · · 0...

.... . .

. . ....

−a0 0 0 0 0

Y j41

Y j42

...

Y j4n

+

0...

bjm...

bj0

σjjuj(t)

Y j4 = Y j

41

Next, we define a coordinate transformation and rewrite the dynamics in terms

of the transformed coordinates as follows

xi = Y1i + Y2i + Y3i +

q∑

j=1

Y j4i

Let us pick i < min(l,m) for clarity of presentation. Then, taking the derivative of

xi, we obtain

xi = Y1i + Y2i + Y3i +

q∑

j=1

Y j4i

= Y1(i+1) − an−iY11 + Y2(i+1) − an−iY21 + Y3(i+1) − an−iY31 +

q∑

j=1

(Y j

4(i+1) − an−iYj41)

=

(

Y1(i+1) + Y2(i+1) + Y3(i+1) +

q∑

j=1

Y j

4(i+1)

)

− an−i

(

Y11 + Y21 + Y31 +

q∑

j=1

Y j41

)

= xi+1 − an−ix1

Page 25: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

19

Similar expressions can be derived for all i = 1, . . . , n by substitution, which gives us

the following state-space realization

x1 = x2 − an−1x1

...

xn−l−1 = xn−l − al+1x1

xn−l = xn−l+1 − alx1 + dlqT (y, t)c+ dl∆

...

xρ−1 = xρ − am+1x1 + dm+1qT (y, t)c+ dm+1∆

xρ = xρ+1 − amx1 + bfmu∗(t) + bfm

+

q∑

j=1

σjjbjmuj(t) + dmqT (y, t)c+ dm∆

...

xn = −a0x1 + bf0u∗(t) + bf0 +

q∑

j=1

σjjbj0uj(t)

+d0qT (y, t)c+ d0∆

y = Y11 + Y21 + Y31 +

q∑

j=1

Y j41 + dy(t) = x1 + dy(t) (2.9)

In addition to the assumptions made previously, we will make the following real-

istic assumptions. The first assumption is standard in adaptive control designs.

A3: The polynomial bfmsm + · · ·+bf0 is Hurwitz and the sign of the high frequency

gain (sign(bfm)) is known, irrespective of the failed actuators.

A4: The extent of parametric uncertainties, modeling error ∆(y, t), output dis-

turbance dy(t) as well as derivative dy(t) satisfy,

θ ∈ Ωθ , θ : θmin < θ < θmax

∆ ∈ Ω∆ , ∆ : |∆(y, t)| ≤ δ(t)

dy ∈ Ωd , dy : |dy(t)| ≤ δd(t)

dy ∈ Ωf , dy : |dy(t)| ≤ δf (t) (2.10)

Page 26: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

20

where

θ = [−an−1, ...,−a0, bfm, .., b

f0 , cp, .., c1, b

fm, . . . , b

f0 ]

T ∈ R2m+n+p+2, θmin and θmax

are assumed to be known, and δ(t), δd(t), and δf(t) are unknown but bounded func-

tions. Note that ∆(y, t) is assumed to be globally bounded w.r.t y. Thus, in the

present work we consider only bounded uncertainties.

2.3.2 State Estimation

In this section, we will describe the design of K-filters [31] for state estimation.

The state-space equations (3.3) can be rewritten as,

x = A0x+ (k − a)x1 + dqT (y, t)c+ bfu∗ + bf + ∆

y = x1 + dy(t) (2.11)

where,

A0 =

−k1

... In−1

−kn 0 . . . 0

k =

k1

...

kn

d =

0(n−l−1)×1

dl

...

d0

bf =

0(ρ−1)×1

bfm...

bf0

bf =

0(ρ−1)×1

bfm...

bf0

a =

an−1

...

a0

(2.12)

and

∆ =

0(l−1)×1

dl∆...

dm∆ +∑q

j=1 σjjbjmuj(t)...

d0∆ +∑q

j=1 σjjbj0uj(t)

(2.13)

Page 27: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

21

represents the bounded disturbances.

The observer matrix A0 can be made stable by a suitable choice of k. Thus, there

exists a symmetric positive definite matrix P such that,

PA0 + AT0 P = −I, P = P T > 0 (2.14)

For the purpose of state-estimation, the following set of K-filters is defined,

ξn = A0ξn + ky

ξi = A0ξi + en−iy, 0 ≤ i ≤ n− 1

υi = A0υi + en−iu∗, 0 ≤ i ≤ m

ψi = A0ψi + dqi(y, t), 1 ≤ i ≤ p

ζi = A0ζi + en−i, 0 ≤ i ≤ m (2.15)

where ei denotes the ith standard basis vector in Rn and ξn, ξi, υi, ψi and ζi ∈ R

n×1.

Note that due to the special structure of A0, the order of the K-filters described above

can be reduced by using the following two filters and certain algebraic expressions,

η = A0η + eny

λ = A0λ+ enu (2.16)

Now, the ξi and υi filter states can be obtained using the following expression,

ξn = −An0η

ξi = Ai0η 0 ≤ i ≤ n− 1

υi = Ai0λ 0 ≤ i ≤ m (2.17)

Using the above filters, the state estimate x ∈ Rn×1 is given by,

x = ξn −n−1∑

i=0

aiξi +m∑

i=0

bfi υi +m∑

i=0

bfi ζi +

p∑

i=1

ciψi (2.18)

Let εx = x − x be the estimation error. Then, using (4.46), (3.10) and the filters

described above, the estimation error dynamics is given by,

εx = A0εx + (a− k)dy + ∆ (2.19)

Page 28: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

22

Using assumption A4 and taking derivative of the positive semi-definite (psd)

function Vεx = εTxPεx,

Vεx = εTx (AT

0 P + PA0)εx + 2εTx ((a− k)dy + ∆)

≤ −|εx|2 + 2|εx|δ(t) = −|εx|22

− 1

2(|εx| − 2δ)2 + 2δ2

≤ −|εx|22

+ 2δ2 ≤ − Vε

2pmax

+ 2δ2 (2.20)

where δ = ||(a − k)|δd + |∆||, pmin = mineig(P ), pmax = maxeig(P ) and using

pmin|εx(t)|2 ≤ Vεx(t) ≤ pmax|εx(t)|2. Using comparison lemma, from (2.20) we obtain

Vεx(t) ≤ exp

(

− t

2pmax

)

Vεx(0) + 4pmax‖δ‖2∞

(

1 − exp

(

− t

2pmax

))

⇒ |εx(t)|2 ≤ pmax

pminexp

(

− t

2pmax

)

|εx(0)|2 + 4pmax

pmin‖δ‖2

(

1 − exp

(

− t

2pmax

))

⇒ |εx(t)| ≤√

pmax

pmin

exp

(

− t

4pmax

)

|εx(0)| + 2

pmax

pmin

‖δ‖∞

(

1 − exp

(

− t

2pmax

))

(2.21)

In (3.15), the first term is exponentially vanishing and the second term is a bounded

function of time. Thus, we have

εx ∈ Ωε , εx : |εx| ≤ δε(t) (2.22)

i.e., the estimation error remains bounded within a ball of unknown radius δε.

2.3.3 Parameter Projection

Let θ denote the estimate of θ and θ = θ−θ denote the estimation error.Parameter

projection is used to ensure that the parameter estimates remain within a known

bounded region. The update law and the projection mapping used here have the

following form,

˙θ = Projθ(Γτ) (2.23)

Projθi(•i) =

0 if θi = θi,max and •i > 0

0 if θi = θi,min and •i < 0

•i otherwise

(2.24)

Page 29: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

23

where Γ > 0 is a diagonal matrix, and τ is any adaptation function. The projection

mapping guarantees that the following two properties are always satisfied,

P1 θ ∈ Ωθ = θ : θmin ≤ θ ≤ θmax (2.25)

P2 θT (Γ−1Projθ(Γτ) − τ) ≤ 0, ∀τ (2.26)

2.3.4 Controller Design

In this section, we present the design of output feedback based Adaptive Robust

Fault-Tolerant Control (ARFTC) scheme. The main idea is to synthesize a virtual

control law at each step which will replace the estimated state and, the estimation

error will be dealt with at each step using robust feedback.

Step 1: The derivative of the output tracking error z1 = y − yr is given by,

z1 = x2 − an−1y + an−1dy + dy − yr (2.27)

But, x2 is not measured and is replaced by its estimate using (3.10),

x2 = ξn,2 −n−1∑

i=0

aiξi,2 +m∑

i=0

bfi υi,2

+m∑

i=0

bfi ζi,2 +

p∑

i=1

ciψi,2 + εx2 (2.28)

where εx2 is the estimation error of x2, and denote

ξ(2) = [ξn−1,2, . . . , ξ0,2], υ(2) = [υm,2, . . . , υ0,2],

ζ(2) = [ζm,2, . . . , ζ0,2], ψ(2) = [ψp,2, . . . , ψ1,2]. (2.29)

Substituting (3.21) back in (3.20), we obtain

z1 = bfmυm,2 + ω0 + θT ω − yr + ∆1 (2.30)

where ωT = [ξ(2), υ(2), ψ(2), ζ(2)] + e∗T1 y, ω = ω − e∗n+1υm,2, ω0 = ξn,2, ∆1 = an−1dy +

dy + εx2 and e∗Ti is the ith standard basis vector in Rn+2m+p+2. (3.23) suggests a

Page 30: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

24

natural choice for the virtual input is υm,2, which will be used for synthesizing the

virtual control law α1,

α1(y, η, λm+1, ψ, θ, t) = α1a + α1s,

α1a = − 1

bfmω0 + θT ω − yr (2.31)

where λi = [λ1, . . . , λi]T is obtained from (2.16). In (3.24), α1a is the model compen-

sation component of the control law used to achieve an improved model compensation

through on-line parameter adaptation. Thus, the fault is partly accommodated using

model compensation, as [θn+m+p+2, ..., θn+2m+p+2]T = [bfm, ...,

ˆbf0 ]T . Let z2 = υm,2 − α1

denote the input discrepancy. Substituting (3.24) into (3.23), we get

z1 = bfm(z2 + α1s) − θTφ1 + ∆1, φ1 , ω + e∗n+1α1a (2.32)

We now present the design of the robust component of the control law α1s, which

suppresses the potential destabilizing effect of parameter estimation transients.

α1s = α1s1 + α1s2 + α1s3, α1s1 = − 1

(bfm)min

k1sz1 (2.33)

where (bfm)min = minj(ηjjbjm) = (θn+1)min and k1s is a nonlinear gain, such that

k1s ≥ g1+ ‖ Cφ1Γφ1 ‖2, g1 ≥ 0 (2.34)

in which Cφ1is a positive definite constant diagonal matrix to be specified later

and g1 is a constant. Consider the positive semi-definite (p.s.d) function V1 = 12z21 .

Substituting (3.26) and (3.27) in (3.25), V1 is given by

V1 ≤ bfmz1z2 − k1sz21 + z1(b

fmα1s2 − θTφ1)

+z1(bfmα1s3 + ∆1) (2.35)

As a result of assumption A4, we have

‖ θTφ1 ‖≤‖ θM ‖‖ φ1 ‖ (2.36)

Page 31: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

25

where θM = θmax − θmin. Thus, ‖ θTφ1 ‖ is bounded by a known function, which

ensures that there exists a robust control function satisfying the following conditions

[17]:

(a) z1bfmα1s2 − θTφ1 ≤ ǫ11

(b) z1α1s2 ≤ 0 (2.37)

where ǫ11 is a positive design parameter.

Remark 1: Condition (a) of (3.31) shows that α1s2 is synthesized to attenuate the

effect of parametric uncertainties θ with the level of control accuracy being measured

by ǫ11. Condition (b) is to make sure that α1s2 is dissipative in nature so that it does

not interfere with the functionality of adaptive control law α1a.

Similarly, from assumption A4 and (3.14), we obtain

|∆1| ≤ δ1(t) , |an−1|δd(t) + δf (t) + δε2(t) (2.38)

Note that δ1 is a bounded function, and the same strategy as in (3.31) can be used

to design a robust control law. However, since the bound of ∆1 is not known, it is

impossible to prespecify the level of control accuracy exactly. So, a more relaxed

requirement compared to the condition (a) of (3.31) is sought to be satisfied,

z1bfmα1s3 + ∆1 ≤ ǫ12δ21 (2.39)

where ǫ12 is a controller parameter which can be freely tuned. This implies that by

choosing a sufficiently small ǫ12, right hand side of equation (3.37) can be made smaller

than any desired value, even though it cannot be prespecified exactly. Examples of

smooth α1s2 and α1s3 satisfying (3.31) and (3.37) respectively are given by [32],

α1s2 = − h1

4(bfm)minǫ11z1, α1s3 = − 1

4(bfm)minǫ12z1 (2.40)

where h1 ≥‖ θM ‖2‖ φ1 ‖2. Also, it should be noted that the effect of unstructured

uncertainties like uj(t), dy(t), and ∆(y, t), which manifests itself in δε2, is compensated

by α1s3 i.e., robust feedback .

Page 32: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

26

Step 2: From (3.24), (2.16), (2.17) and rearrangement of (3.20-3.23), the derivative

of α1 can be written as,

α1 = α1c + α1u

α1c =∂α1

∂y(ω0 + θTω) +

∂α1

∂ηη +

m+1∑

j=1

∂α1

∂λj

λj

+

p∑

j=1

∂α1

∂ψj

ψj +∂α1

∂t

α1u =∂α1

∂y(−θTω + ∆1) +

∂α1

∂θθ (2.41)

Using (2.15) and (2.16), α1c is calculable and can be used in the design of control

functions. However, α1u is not calculable due to various uncertainties and hence, will

be dealt with via robust feedback in this step. From (2.15) and (3.39), the derivative

of the z2 = υm,2 − α1 is

z2 = υm,3 − k2υm,1 − α1c − α1u (2.42)

Now, consider the augmented p.s.d function V2 = V1 + 12z22 . From (3.29) and (3.40),

the derivative of V2 is given by

V2 ≤ V1|α1+ z2bfmz1 + υm,3 − k2υm,1 − α1c − α1u (2.43)

where V1|α1= −k1sz

21 + z1(b

fmα1s2 − θ1φ1) + z1(b

fmα1s3 + ∆1). As in (3.24), the ARC

control function α2 for the virtual control input υm,3 in (3.40) consists of

α2(y, η, λm+2, ψ, θ, t) = α2a + α2s

α2a = −bfmz1 + k2υm,1 + α1c

α2s = α2s1 + α2s2 + α2s3, α2s1 = −k2sz2

k2s ≥ g2 +

∂α1

∂θCθ2

+ ‖Cφ2Γφ2‖2 (2.44)

where g2 > 0 is a constant and Cθ2 and Cφ2 are positive definite constant diagonal

matrices, α2s2 and α2s3 are robust control functions to be chosen later. As mentioned

previously, due to use of discontinuous projection, we cannot use tuning functions

Page 33: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

27

which anticipates and compensates for the effect of˙θ. The last two terms in the

inequality for k2s compensates for this loss of information. Substituting (3.42) and

(3.39) in (3.41), and using similar techniques as in (3.25), we have

V2 ≤ V1|α1+ z2z3 − k2sz

22 + z2(α2s2 − θTφ2)

+z2(α2s3 + ∆2) − z2∂α1

∂θ

˙θ (2.45)

where z3 = υm,3 − α2 represents the input discrepancy and

φ2 = e∗n+1z1 −∂α1

∂yω , ∆2 = −∂α1

∂y∆1 (2.46)

From (3.36), it follows that ∆2 ≤ |∂α1/∂y|δ1. Similar to (3.31) and (3.37), the robust

control functions α2s2 and α2s3 are chosen to satisfy

(a) z2(α2s2 − θTφ2) ≤ ǫ21

(b) z2(α2s3 + ∆2) ≤ ǫ22δ21

(c) z2α2s2 ≤ 0 , z2α2s3 ≤ 0 (2.47)

where ǫ21 and ǫ22 are positive design parameters. As in step 1, α2s2 and α2s3 can be

chosen as,

α2s2 = − h2

4ǫ21z2 , α2s3 = − 1

4ǫ21

(

∂α1

∂y

)2

z2 (2.48)

where h2 is any smooth function satisfying h2 ≥‖ θM ‖2‖ φ2 ‖2. From (3.29) and h2

defined above, the derivative of V2 satisfies

V2 ≤ z2z3 −2∑

j=1

kjsz2j + z1(b

fmα1s2 − θ1φ1)

+z1(bfmα1s3 + ∆1) + z2(α2s2 − θTφ2)

+z2(α2s3 + ∆2) −∂α1

∂θ

˙θz2 (2.49)

Step i (3 ≤ i < ρ): For any j ∈ [3, i− 1], let zj = υm,j − αj−1 and recursively define

φj = −∂αj−1

∂yω , ∆j = −∂αj−1

∂y∆1 (2.50)

Page 34: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

28

At step i, choose the desired ARC control function αi as

αi(y, η, λm+i, ψ, θ, t) = αia + αis

αia = −zi−1 + kiυm,1 + α(i−1)c

αis = αis1 + αis2 + αis3, αis1 = −kiszi

kis ≥ gi +

∂αi−1

∂θCθi

+ ‖CφiΓφi‖2 (2.51)

where gi > 0 is a constant, and Cθi and Cφi are positive definite constant diagonal

matrices, αis2 and αis3 are robust control functions satisfying,

(a) zi(αis2 − θTφi) ≤ ǫi1

(b) zi(αis3 + ∆1) ≤ ǫi2δ21

(c) ziαis2 ≤ 0 , ziαis3 ≤ 0 (2.52)

and

α(i−1)c =∂αi−1

∂y(ω0 + θTω) +

m+i+1∑

j=1

∂αi−1

∂λj

λj

+

p∑

j=1

∂αi−1

∂ψj

ψj +∂αi−1

∂ηη +

∂αi−1

∂t(2.53)

Then, the ith error subsystem is

zi = zi+1 − zi−1 − kiszi + (αis2 − θTφi)

+(αis3 + ∆i) −∂αi−1

∂θ

˙θ (2.54)

and the derivative of the augmented p.s.d function Vi = Vi−1 + 1/2z2i satisfies,

Vi ≤ zizi+1 −i∑

j=1

kjsz2j + z1(b

fmα1s2 − θTφ1)

+i∑

j=2

zj(αjs2 − θTφj) + z1(bfmα1s3 + ∆1)

+

i∑

j=2

zj(αjs3 + ∆j) −i∑

j=2

∂αj−1

∂θ

˙θzj (2.55)

Page 35: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

29

Step ρ: In this final step, the actual control law u∗ will be synthesized such

that υm,ρ tracks the desired ARC control function αρ−1. The derivative of zρ can be

obtained as

zρ = υm,ρ+1 + u∗ − kρυm,1 − α(ρ−1)c

−∂αρ−1

∂y(−θTω + ∆1) −

∂αρ−1

∂θ

˙θ (2.56)

If υm,ρ+1 + u∗ were the virtual input, (2.56) would have the same form as the inter-

mediate step i. Therefore, the general form, (3.48-2.56) applies to step ρ. Since u is

the actual control input, it can be chosen as,

u∗ = αρ − υm,ρ+1 (2.57)

where αρ is given by (3.49). Then, zρ+1 = u∗ + υm,ρ+1 − αρ = 0.

Theorem 1. Let the parameter estimates be updated using adaptation law (5.20)

in which τ is chosen as

τ =

ρ∑

j=1

φjzj (2.58)

If diagonal controller gain matrices Cθj and Cφk are chosen such that c2φkr ≥ ρ

4

∑ρ

j=1 1/c2θjr,

where cθjr and cφkr are the rth diagonal element of Cθj and Cφk respectively. Then,

the control law (3.51) guarantees that,

1. In general the control input and all internal signals are bounded. Furthermore,

Vρ is bounded above by,

Vρ(t) ≤ exp(−λρt)Vρ(0)

+ǫρ1 + ǫρ2 ‖ δ1 ‖2

λρ

[1 − exp(−λρt)] (2.59)

where λρ = 2ming1, . . . , gρ, ǫρ1 =∑ρ

j=1 ǫj1, ǫρ2 =∑ρ

j=1 ǫj2 and ‖ δ1 ‖2∞

stands for the infinity norm of δ1.

2. If after a finite time t0, ∆ = 0, uj = 0 and dy = 0 (i.e., in the presence of

parametric uncertainties and modeled disturbances only) then, in addition to

results in (1), asymptotic output tracking control is also achieved.

Page 36: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

30

Proof. Using (3.51), we known zρ+1 = 0. From (3.49), we have

αρ(y, η, λn, ψ, θ, t) = αρa + αρs

αρa = −zρ−1 + kρυm,1 + α(ρ−1)c

αρs = αρs1 + αρs2 + αρs3, αρs1 = −kρszρ

kρs ≥ gρ+ ‖ ∂αρ−1

∂θCθρ ‖ + ‖ CφρΓφρ ‖2 (2.60)

Substituting in (2.56), we obtain

zρ = −zρ−1 − kρszρ + (αρs2 − θTφρ)

+(αρs3 + ∆ρ) −∂αρ−1

∂θ

˙θ (2.61)

Also, substituting i = ρ− 1 in (2.55) gives

Vρ−1 ≤ zρ−1zρ −ρ−1∑

j=1

kjsz2j + z1(b

fmα1s2 − θTφ1)

+

ρ−1∑

j=2

zj(αjs2 − θTφj) + z1(bfmα1s3 + ∆1)

+

ρ−1∑

j=2

zj(αjs3 + ∆j) −ρ−1∑

j=2

∂αj−1

∂θ

˙θzj (2.62)

Taking derivative of the Lyapunov function Vρ = Vρ−1|αρ−1+ 1

2z2

ρ and using (2.61)

and (2.62) we obtain

Vρ = Vρ−1|αρ−1+ zρzρ

≤ −ρ∑

j=1

kjsz2j + z1(b

fmα1s2 − θTφ1)

+

ρ∑

j=2

zj(αjs2 − θTφj) + z1(bfmα1s3 + ∆1)

+

ρ∑

j=2

zj(αjs3 + ∆j) −ρ∑

j=2

zj

∂αj−1

∂θ

˙θ (2.63)

Page 37: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

31

Now, using the fact that robust component of the virtual control law is designed to

satisfy inequalities (3.31),(3.45) and (3.50), we have

Vρ ≤ −ρ∑

j=1

(gj +

∂αj−1

∂θCθj

+ ‖ CφjΓφj ‖2)z2j

+

ρ∑

j=1

(ǫj1 + ǫj2δ21) −

ρ∑

j=2

zj

∂αj−1

∂θ

˙θ (2.64)

By AM-GM inequality y1+...+yp

p≥ p

√y1.y2...yp, the last term of the above inequality

satisfies

−ρ∑

j=2

zj

∂αj−1

∂θ

˙θ ≤

ρ∑

j=2

|zj |∣

∂αj−1

∂θCθjC

−1θj

˙θ

≤ρ∑

j=2

(

∂αj−1

∂θCθj

2

z2j +

1

4‖C−1

θj

˙θ‖2

)

(2.65)

Once again, using AM-GM inequality we have

‖C−1θj

˙θ‖2 = ‖C−1

θj Proj(Γτ)‖2 ≤ ‖C−1θj Γτ‖2

≤ (

ρ∑

k=1

‖C−1θj Γφkzk‖)2 ≤ ρ(

ρ∑

k=1

‖C−1θj Γφk‖)2z2

k (2.66)

Putting (2.65) and (2.66) together, we get

−ρ∑

j=2

zj

∂αj−1

∂θ

˙θ ≤

ρ∑

j=2

∂αj−1

∂θCθj

2

z2j

4

ρ∑

k=1

ρ∑

j=2

‖C−1θj Γφk‖2z2

k

≤ρ∑

j=2

∂αj−1

∂θCθj

2

z2j

+

ρ∑

k=1

‖CφkΓφk‖2z2k (2.67)

Finally combining (2.64) and (2.67), we have

Vρ ≤ −ρ∑

j=1

gjz2j +

ρ∑

j=1

(ǫj1 + ǫj2δ21)

≤ −λρVρ + ǫρ1 + ǫρ2δ21 (2.68)

Page 38: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

32

which proves first part of the theorem.

To prove the second part, we note that in presence of parametric uncertainties

only i.e., ∆ = 0, dy = 0 and uj = 0, ∆1 = ǫ2. From (3.37) and condition b of (3.50),

we have |z1(bfmα1s3 +∆1) ≤ ǫ12ε22| and |zj(αjs3 +∆j) ≤ ǫj2ε

22|, j = 2, ..., ρ. Thus, from

(3.54), (2.55), condition b of (3.31), and (3.45), Vρ satisfies

Vρ ≤ −ρ∑

j=1

(θTφ1zj + gjz2j − ǫj2ε

22)

= −ρ∑

j=1

gjz2j − θT τ + ǫρ2ε

22 (2.69)

Now, we use the augmented positive definite function

Vθ = Vρ +1

2θT Γ−1θ + γεTPε (2.70)

where γ ≥ ǫρ2. Now using (5.24) and (2.14), the derivative of Vθ is given by

Vθ ≤ −ρ∑

j=1

gjz2j − θT τ + ǫρ2ε

22 + θT Γ−1θ − γ‖ε‖2

≤ −ρ∑

j=1

gjz2j (2.71)

It may appear that we have neglected the ρ+1 to n states in the present analysis.

But, due to the assumption of stable zero dynamics and bounded disturbances as

well as actuator faults, it can be easily proved using standard adaptive control argu-

ments that all internal signals remain bounded and do not interfere with the tracking

performance.

Remark 2: In context of actuator fault compensation, first part of theorem 1

guarantees that the jump in parameter values due to failed actuator does not interfere

with the desired transient performance. By using trajectory initialization techniques

as described in [17], we can set Vρ(0) = 0 and then, from (3.52) we have

|z1(t)| ≤√

2(ǫρ1 + ǫρ2 ‖ δ1 ‖2∞)

λρ

[

1 − exp(−λρt)

2

]

(2.72)

This equation provides an upperbound for the output tracking error z1 and charac-

terizes the transient response. ǫρ1, ǫρ2 and λρ are controller parameters which we

Page 39: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

33

can choose. Note that from remark 1, we know by properly tuning ǫρ2 we can make

ǫρ2 ‖ δ1 ‖2∞ as small as desired. In this sense, we have guaranteed transient response.

This result on transient response of the system is a direct consequence of underlying

robust filter structure of the ARC controller.

2.4 Simulation Example: Linearized Boeing 747 Model

For simulation purposes, we will use the linearized model of Boeing 747. It should

be noted that the same example was used in [3, 33], and thus, provides a platform

to compare the MRAC in [3] and its robust adaptive control version in [33] with the

backstepping based ARC algorithm proposed here for actuator fault accommodation.

Plant Model: The linearized model for the lateral motion of Boeing 747 without

disturbances can be represented as,

x(t) = Ax(t) +Bu(t), y(t) = x2(t) = yr(t)

x(t) = [β, yr, p, φ]T , B = [b1, b2, b3] (2.73)

where, β is the side-slip angle, yr is the yaw-rate, p is the roll rate, φ is the roll

angle, y is the output which needs to follow the reference trajectory r(t) and u is

the control input vector consisting of three control signals representing three rudder

servos δr1, δr2, δr3. Note that the B matrix has been augmented by b2 and b3 for

studying actuator failure compensation properties of the proposed algorithm. From

Page 40: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

34

the data provided in [34] for Boeing 747 in horizontal flight at 40,000 ft and nominal

forward speed 774 ft/s, the perturbation dynamics matrices are

A =

−0.0558 −0.9968 0.0802 0.0415

0.598 −0.115 −0.0318 0

−3.05 0.388 −0.465 0

0 0.0805 1 0

(2.74)

b1 =

0.00729

−0.475

0.153

0

b2 =

0.01

−0.5

0.2

0

b3 =

0.005

−0.3

0.1

0

(2.75)

It can be easily verified that this plant satisfies the assumptions A1-A4.

Simulation results: Simulations are done using r(t) = 0.02sin(0.2t) as the refer-

ence signals for MRAC and ARC based fault compensation techniques. Three faults

are introduced in the system during the simulations: the third actuator gets stuck

at u3 = 0.5 rads at 40 seconds and the second actuator loses 90% efficiency at 60

seconds. These faults can be represented as follows,

u2(t) =

u∗(t) for t ≤ 60 sec

0.1u∗(t) for t > 60 sec

u3(t) =

u∗(t) for t ≤ 40 sec

0.5 rads for t > 40 sec(2.76)

and u1 is assumed to be healthy for all times i.e., u1(t) = u∗(t), ∀t. For the state

estimators, the gain was chosen to be k = [4, 6, 4, 1]T , which places all the poles of

A0 at -1. The initial conditions for the plant was x = [0, 0.02, 00]′

. The controller

parameters for the MRAC based scheme can be obtained from [3]. The details of

ARC controller is given at the end of this section. Two scenarios are studied through

simulations: with and without disturbance. In the first case, as all disturbances are

assumed to be zero, we have ∆(y, t) = dy(t) = 0. As can be seen in Fig.1, both the

systems perform well initially, and have similar commanded control input profiles.

Page 41: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

35

Note that the magnitude of the commanded control input increases with each fault.

This is to be expected as initially three actuators were performing the task, whereas

with each failure, the same task needs to be done with fewer remaining healthy

actuators. As the faults chosen are fairly severe, they cause large jump in the plant

parameters. Consequently, the transient tracking error deviates significantly for the

MRAC based scheme, but stays close to zero for the proposed adaptive robust fault-

tolerant control (ARFTC). This can be explained as follows. The design of robust

component of the ARFTC law has already incorporated the extent of such jumps

in parameter values, and hence, is better suited to handle the transients introduced

by actuator failures. This shows the guaranteed transient performance of ARFTC

which cannot be achieved using any adaptive control based scheme. On the other

hand, it should also be noted that due to the learning capability of the ARFTC,

the steady-state error tends to zero. This cannot be achieved using a fault-tolerant

scheme which relies on robust feedback only.

In the next set of simulations, intermediate disturbance ∆(y, t) = 0.004sin(2.1t)

with d = [1, 1.50, 0.5625, 0.0625]′

and completely unknown output disturbance dy(t) =

0.002sin(4t) are added to the system. Note that in context of aircraft control, many

disturbances can be modeled using harmonic functions, e.g., wind-shear [35] which

corresponds to the ∆y term. Also, rate-gyros which are used for measuring yaw-rates

suffer from harmonic disturbances (see [36]) as well. Hence, sinusoidal disturbances

were used for simulation studies. Fig.2 shows result with the same set of faults as

described earlier for r(t) = 0.02sin(0.2t). To make the simulation studies more mean-

ingful, robustness modifications and disturbance estimation was added to the MRAC

based design, as outlined in [33], with bounds given by [θi,min, θi,max] , [0.80θ∗i , 1.20θ∗i ].

These simulations demonstrate the strength of ARFTC in attenuating the effect of

disturbance and modeling error in addition to large parametric uncertainties. In fact,

even though the magnitude of the control signals remain comparable for MRAC and

ARFTC based designs, there is an order of magnitude difference in the tracking error.

As seen in Fig. 2, even with robust adaptation mechanisms and disturbance estima-

Page 42: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

36

tion, MRAC based schemes can neither attenuate the effect of unknown modeling

errors like dy(t) on the steady-state error nor guarantee desired transient response.

On the other hand, due to the underlying robust controller structure of ARFTC,

it can attenuate the effect of such uncertainties on the steady-state error as well as

guarantee desired transient response.

2.4.1 Detailed ARFTC Controller design

The plant model in the transfer function (TF) form is given by

[y(t)] =−0.4750s3 − 0.2479s2 − 0.1187s− 0.0563

s4 + 0.6358s3 + 0.9389s2 + 0.5116s+ 0.0037[u1(t)]

+−0.5s3 − 0.2608s2 − 0.1223s− 0.0583

s4 + 0.6358s3 + 0.9389s2 + 0.5116s+ 0.0037[u2(t)]

+−0.3s3 − 0.1564s2 − 0.0747s− 0.0355

s4 + 0.6358s3 + 0.9389s2 + 0.5116s+ 0.0037[u3(t)]

+s3 + 1.50s2 + 0.5625s+ 0.0625

s4 + 0.6358s3 + 0.9389s2 + 0.5116s+ 0.0037[∆(y, t)] + dy(t)(2.77)

where ∆(y, t) = 0.004 sin(2.1t), dy(t) = 0.002 sin(4t). The corresponding observer

canonical form is given by

x1 = x2 − 0.6358x1 − 0.4750u1 − 0.5u2 − 0.3u3 + 0.004sin(2.1t)

x2 = x3 − 0.9389x1 − 0.2479u1 − 0.2608u2 − 0.1564u3 + 0.0.006sin(2.1t)

x3 = x4 − 0.5116x1 − 0.1187u1 − 0.1223u2 − 0.0747u3 + 0.0023sin(2.1t)

x4 = −0.0037x1 − 0.0563u1 − 0.0583u2 − 0.0355u3 + 0.0003sin(2.1t)

y = x1 + dy(t)

For the filters, we use k = [4, 6, 4, 1], which places all the poles of A0 at −1. The

filters are given by

η1

η2

η3

η4

=

−4 1 0 0

−6 0 1 0

−4 0 0 1

−1 0 0 0

η1

η2

η3

η4

+

0

0

0

1

y

Page 43: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

37

from which we can obtain the required ξi, i = 0, 1, 2, 3, 4 filter-states as follows

ξ0 = η, ξ1 = A0η, ξ2 = A20η, ξ3 = A3

0η, ξ4 = −A40η

The next set of filter is given by

λ1

λ2

λ3

λ4

=

−4 1 0 0

−6 0 1 0

−4 0 0 1

−1 0 0 0

λ1

λ2

λ3

λ4

+

0

0

0

1

u

from which we can obtain υi, i = 0, 1, 2, 3

υ0 = λ, υ1 = A0λ, υ2 = A20λ, υ3 = A3

ψ1

ψ2

ψ3

ψ4

=

−4 1 0 0

−6 0 1 0

−4 0 0 1

−1 0 0 0

ψ1

ψ2

ψ3

ψ4

+

1

1.5

0.5625

0.0625

sin(2.1t)

Note that in this case it assumed that D(s) = s3 +1.50s2 +0.5625s+0.0625 is known.

But, the controller can be designed when D(s) has unknown coefficients, with suitable

modifications, the details of which were presented in the previous response.

The last set of filters is given by

ζi = A0ζi + e4−i, i = 1, 2, 3, 4 (2.78)

where e4−i is the standard basis in R4

Now, we are ready to present the controller design. Using the controller parametriza-

tion proposed (see equation (3)), the derivative of the error is given be

z1 = y − yr = (x1 + dy) − yr

z1 = x2 − 0.6358x1 − 0.4750u1 − 0.5u2 − 0.3u3 + 0.004sin(2.1t) + dy − yr

= x2 − 0.6358x1 + [0.4750η11(1 − σ11) + 0.5η22(1 − σ22) + 0.3η33(1 − σ33)]u∗(t)

+0.4750σ11(u11 + u11(t)) + 0.5σ22(u22 + u22(t)) + 0.3σ33(u33 + u33(t))

+0.004sin(2.1t) + 0.6358dy + dy − yr

Page 44: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

38

The unmeasured state x2 is replaced by the estimated state x2 = ξ4(2)−ξ(2)a+υ(2)bf +

ζ(2)bf +ψ2c+ εx2, where the unknown parameters a, bf , bf and c are as defined in the

manuscript. With this, the dynamics can be rewritten as

z1 = ξ4(2) − ξ(2)a+ υ(2)bf + ζ(2)b

f + ψ2c+ εx2 − 0.6358y + 0.004 sin(2.1t) + 0.6358dy + dy − yr

= b3υ3,2 + ω0 + θT ω − yr − ∆1

where ωT = [ξ(2), υ(2), ψ(2), ζ(2)] + e∗T1 y ∈ R13, ω = ω− e∗5υ3,2, (e∗j is jth standard basis

in R13×1) ω0 = ξ4,2, ∆1 = 0.004 sin(2.1t) + 0.6358dy + dy + εx2.

The control law is given by

α1 = α1a + α1s

α1a =1

bf3ω0 + θT ω − yr

The robust component of the control law is designed as

α1s1 = − 1(

bf3

)

min

k1sz1

α1s2 = − h1

4(bfm)minǫ11z1

α1s3 = − 1

4(bfm)minǫ12z1

(2.79)

where k1s ≥ g1 + ‖Cφ1Γφ1‖, g1 ≥ 0. However, as the relative degree of the system is

1, there is no restriction on the choice of g1 and Cφ1, as long as both are greater than

0. Thus, they can always be selected such that g1 + ‖Cφ1Γφ1‖ ≤ 60 = k1s. The other

parameter h1 is chosen as

h1 = ‖θM‖‖φ1‖

where φ1 = ω + e∗5α1a and θM = ‖θmax − θmin‖2. The other parameters are given by

ǫ11 = ǫ22 = 0.2.

˙θ = ProjθΓτ1, where τ1 = φ1z1, Γ = 50I13×13

Page 45: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

39

θ = [−a3,−a2,−a1,−a0,−bf3 ,−bf2 ,−bf1 ,−bf0 , c, bf3 , bf2 , bf1 , bf0 ]

θmin = [0.5086, 0.7511, 0.4093, 0.0030, 0.0240, 0.0125, 0.0060, 0.0028, 0.0032,

−1.506,−0.7856,−0.3728,−0.1773]

θmax = [0.7630, 1.1267, 0.6139, 0.0044, 1.5300, 0.7981, 0.3788, 0.1801, 0.0048,

1.506, 0.7856, 0.3728, 0.1773]

This completes the design of the ARC based fault-tolerant control for the linearized

Boeing 747 model.

Page 46: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

40

0 20 40 60 80−0.01

0

0.01

0.02

0.03

0.04

time (sec)y

and

ym (r

ad/s

ec)

Plant output and reference signal

MRACARCRef

0 20 40 60 80−0.01

0

0.01

0.02

0.03

time (sec)

e(t)

(rad

/sec

)Tracking error

MRACARC

0 20 40 60 80−0.6

−0.4

−0.2

0

0.2

0.4

time (sec)

u(t)

(rad

)

Control signals: MRAC

u1(t)

u2(t)

u3(t)

0 20 40 60 80−0.6

−0.4

−0.2

0

0.2

0.4

time (sec)

u(t)

(rad

)

Control signals: ARC

u1(t)

u2(t)

u3(t)

Figure 2.1. Reference tracking, control signals and tracking error forMRAC versus ARC based fault-tolerant schemes in absence of distur-bances

Page 47: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

41

0 20 40 60 80−0.01

0

0.01

0.02

0.03

0.04

time (sec)y

and

ym (r

ad/s

ec)

Plant output and reference signal

Robust MRACARCRef

0 20 40 60 80−0.01

0

0.01

0.02

0.03

time (sec)

e(t)

(rad

/sec

)Tracking error

Robust MRAC

ARC

0 20 40 60 80−0.6

−0.4

−0.2

0

0.2

0.4

time (sec)

u(t)

(rad

)

Control signals: Robust MRAC

u1(t)

u2(t)

u3(t)

0 20 40 60 80−0.6

−0.4

−0.2

0

0.2

0.4

time (sec)

u(t)

(rad

)

Control signals: ARC

u1(t)

u2(t)

u3(t)

Figure 2.2. Reference tracking, control signals and tracking error forMRAC versus ARC based fault-tolerant schemes in presence of distur-bances

2.5 Conclusion

In this work, an output feedback based Adaptive Robust Fault-Tolerant Control

(ARFTC) scheme was presented for the fault accommodation of uncertain linear sys-

Page 48: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

42

tems with a larger class of unknown actuator faults. Adaptation and robust feedback

are used simultaneously to maintain tracking performance in face of large parametric

uncertainties introduced due to failing actuators, exogenous disturbances and other

modeling uncertainties. Comparative simulation studies were done using a linearized

model for lateral motion of Boeing 747 which confirmed the superior performance of

the proposed ARFTC strategy over conventional robust MRAC based schemes. In

summary, some of the salient features of the ARFTC scheme presented in the chapter

are, (1) capability to handle large parametric uncertainties due to unknown actuator

failures with guaranteed transient performance with desired steady-state tracking er-

ror, (2) better disturbance rejection properties and (3) guaranteed robust performance

when adaptation is switched off.

Page 49: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

43

3. OUTPUT FEEDBACK BASED ADAPTIVE ROBUST FAULT-TOLERANT

CONTROL FOR A CLASS OF UNCERTAIN NONLINEAR SYSTEMS

3.1 Introduction

In this chapter, we extend the output feedback based adaptive robust fault-tolerant

control (ARFTC) developed in the previous chapter to a class uncertain nonlinear

systems. The faults considered here comprise of stuck actuators, loss in actuator

efficiency or a combination of the two faults. Tao et al. proposed direct fault-

compensation schemes to deal with such faults in [4], [5] using backstepping based

adaptive control. Another popular adaptive approach to solve this problem is mul-

tiple model adaptive control (MMAC), switching and tuning [6]. Robust adaptive

control (RAC) based schemes, that can guarantee boundedness of closed loop signals

in presence of unknown modeling uncertainties and disturbances, were investigated

in [5,7,8]. As, backstepping based RAC schemes accommodate the effect of parame-

ter estimation transients ( ˙θ) through tuning functions, they have better performance

than MRAC based techniques, which rely on certainty-equivalence principle. How-

ever, they still lack the ability to attenuate the effect of non-parametric uncertainties

like external disturbances and unknown nonlinearities on the tracking error. Fur-

thermore, as discussed in the first chapter, using high adaptation gain in RAC based

design in presence of disturbances to improve transient performance can results in

high frequency oscillations in the control input, which is extremely undesirable.

A nonlinear Hypersonic aircraft model is used to demonstrate the effectiveness

of the proposed scheme. Note that in this chapter, the comparative studies are car-

ried out with respect to a robust adaptive backstepping based design, whereas in

our previous chapter, the comparative studies were conducted with a model reference

adaptive control (MRAC) based scheme. As MRAC design is fundamentally different

Page 50: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

44

from backstepping based designs, it was not possible to compare and contrast the

two designs at each step. In this chapter, however, the comparison is done between

two backstepping based approaches, which makes it possible to emphasize the under-

lying subtle but important differences in the design of the two fault-tolerant control

strategies. This makes the comparative studies much more compelling.

3.2 Problem Statement

In the present work, we will consider systems which can be represented in the

following output feedback form subjected to output dependent uncertainties

x1 = x2 + ϕ0,1(y) +

p∑

j=1

ajϕ1,j(y) + ∆1(y, t)

...

xρ−1 = xρ + ϕ0,ρ−1(y) +

p∑

j=1

ajϕρ−1,j(y) + ∆ρ−1(y, t)

xρ = xρ+1 + ϕ0,ρ(y) +

p∑

j=1

ajϕρ,j(y) +

q∑

j=1

bm,jβj(y)uj(t) + ∆ρ(y, t)

...

xn = ϕ0,n(y) +

p∑

j=1

ajϕn,j(y) +

q∑

j=1

b0,jβj(y)uj(t) + ∆n(y, t) (3.1)

where ρ = n − m is the relative degree, uj is the control input, y = x1 is the

measured output, ϕ0,i(y) and βj(y) are known smooth functions of y. Furthermore,

it will be assumed that |βj(y)| > 0 and is uniformly bounded above by a constant

i.e., |βj(y)|≤βj for any y. ∆i(y, t) represents uncertainties e.g., modeling error and

disturbances. ai, bi,j are unknown constants such that sign of the high frequency gain

(sgn(bm,j)) is known.

Page 51: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

45

System (3.1) we will be subjected to actuator faults which can be represented as

uj(t) =

uj + uj(t), ∀t ≥ Tf ,

if jth actuator gets stuck at Tf

ηjju∗j(t), ∀t ≥ Tf ,

if jth actuator loses efficiency at Tf

(3.2)

where u∗j represents the control command to the jth actuator, Tf is the unknown

instant of failure, uj is an unknown constant value at which the actuator gets stuck,

uj(t) represents an unknown but bounded disturbance about uj, and ηjj ∈ [(ηjj)min, 1]

represents actuator loss in efficiency. Note that the class of faults considered here is

more practical and significantly larger than what has been considered in the literature

due to the presence of uj(t).

Without actuator redundancy and sufficient control authority, actuator faults can-

not be accommodated and the same is stated formally in the following assumption

A1. System (3.1) is such that the desired control objective can be fulfilled with

up to q − 1 stuck actuators and any number of actuators with loss in efficiency.

The problem we attempt to solve in this work can now be stated as follows. For

the uncertain nonlinear system (3.1), subjected to faults (3.5) the goal is to design an

output feedback based control strategy such that the output tracking error converges

exponentially to a prespecified bound and has a guaranteed transient performance.

Remark 1. In the present work, a simple fault model (3.2) was chosen in order

to clearly illustrate the advantages of using ARFTC over robust adaptive control

(RAC) based scheme, without unnecessarily complicating the presentation. Any fault

which can be linearly parameterized in terms on known basis functions (i.e., uFj (t) =

∑rk=1 bkuj,k(t)+ uj(t), where uF

j (t) is the signal from faulty actuator, bk are unknown

constants, uj,k(t) are known basis functions) can be addressed using the ARFTC

framework.

Page 52: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

46

3.3 Output Feedback based ARFTC

Control signals uj(t) are designed such that βj(y)uj(t) = u∗(t). With fault model

(3.2) and the chosen actuation scheme, we can rewrite the control inputs as follows,

uj(t) =ηjj

βj(y)(1 − σjj)u

∗(t) + σjj(uj + uj(t)) (3.3)

for j = 1, . . . , q where σjj = 1 corresponds to stuck actuators, σjj = 0, (ηjj)min ≤ηjj ≤ 1 represents actuator loss of efficiency and σjj = 0, ηjj = 1 corresponds to

healthy actuators.

With this, we can rewrite the system as follows

xi = x2 + ϕ0,i(y) +

p∑

j=1

ajϕi,j(y) + ∆i(y, t)

for i = 1, 2, .., ρ− 1, and

xρ = xρ+1 + ϕ0,ρ(y) +

p∑

j=1

ajϕρ,j(y) +

q∑

j=1

µm,jβj(y)

+κmu∗(t) +

q∑

j=1

bm,jσjjβj(y)uj(t) + ∆ρ(y, t)

...

xn = ϕ0,n(y) +

p∑

j=1

ajϕn,j(y) +

q∑

j=1

µ0,jβj(y)

+κ0u∗(t) +

q∑

j=1

b0,jσjjβj(y)uj(t) + ∆n(y, t) (3.4)

where

κi =

q∑

j=1

ηjj(1 − σjj)bi,j,

µi,j = σjj ujbi,j, i = 0, 1, .., m j = 1, 2, .., q

Note that κi is the unknown measure of actuator effectiveness after faults and µi,j is

the unknown measure of the fault magnitude which needs to be compensated. Thus,

the system experiences jump in parameter values and bounded disturbances with the

Page 53: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

47

occurrence of each new fault. The proposed scheme accommodates such faults by

estimating µi,j and κi, and relies on robust feedback scheme to deal with the bounded

disturbances uj(t) and the jump in parameter values.

We will make the following realistic assumptions regarding the uncertainties present

in the system

A2. The extent of parametric uncertainties and uncertain nonlinearities satisfy

ai ∈ Ωa , ai : (ai)min < ai < (ai)max

κi ∈ Ωκ , κi : (κi)min < κi < (κi)max

µi,j ∈ Ωµ , µi,j : (µi,j)min < µi,j < (µi,j)max

∆i ∈ Ω∆ , ∆i : |∆i(y, t)| ≤ δi(t) (3.5)

where (ai)min, (ai)max, (κi)min, (κi)max, (µi,j)min, (µi,j)max are known and δi(t) is an

unknown but bounded function.

We will make another standard assumption which guarantees stability of the zero

dynamics

A3. The polynomial κmsm + κm−1s

m−1 + . . . + κ0 is a stable polynomial and

sign(κm) is known, irrespective of the failed actuators.

3.3.1 State estimation

We need to construct state-estimator for

x = A0x+ ky + ϕ0(y) + Φ(y)a+m∑

i=0

en−iκiu∗ +

q∑

j=1

m∑

i=0

en−iµi,jβj(y) + ∆(y, t)

y = Cx, C = [1 0 0 . . . 0] (3.6)

where

A0 =

−k1

... In−1

−kn 0 . . . 0

k =

k1

...

kn

ϕ0(y) =

ϕ0,1

...

ϕ0,n

(3.7)

Page 54: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

48

and

Φ(y) =

ϕ1,1 · · · ϕ1,p

.... . .

...

ϕn,1 · · · ϕn,p

∆(y, t) =

∆1

...

∆ρ−1

∆ρ +∑q

j=1 bm,jσjjβj(y)uj(t)...

∆n +∑q

j=1 b0,jσjjβj(y)uj(t)

(3.8)

Note that A = A0 +kCT and the observer matrix A0 can be made stable by a suitable

choice of observer gain k such that there exists a symmetric positive definite matrix

P satisfying

PA0 + AT0 P = −I, P = P T > 0 (3.9)

We will define the following set of filters for the purpose of state-estimation,

ξ0 = A0ξ0 + ky + ϕ0(y), ξ0 ∈ Rn×1

ξ = A0ξ + Φ(y), ξ ∈ Rn×p

ϑi = A0ϑi + en−iu∗, ϑi ∈ R

n×1

ψi,j = A0ψi,j + en−iβj(y), ψi,j ∈ Rn×1 (3.10)

where i = 0, 1, .., m and j = 1, 2, .., q.

Due to the special structure of A0, the order of K-filters can be reduced by using

the following two filters

λ = A0λ + enu∗

ζj = A0ζj + enβj , j = 1, 2, .., q (3.11)

and the following algebraic equations

ϑi = Ai0λ

ψi,j = Ai0ζj, i = 0, 1, .., m (3.12)

Page 55: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

49

The estimated state can be written as

x = ξ0 + ξa+

m∑

i=0

κiϑi +

q∑

j=1

m∑

i=0

µi,jψi,j (3.13)

Let ε = x− x be the estimation error. Then, the state-estimation error dynamics

is given by

ε = A0ε+ ∆(y, t) (3.14)

From (4.46) and using βj(y) ≤ βj, we know there exists an unknown but bounded

function which upperbounds ∆(y, t) i.e.,

|∆(y, t)| ≤ δ(t) (3.15)

where | • | denotes the L2 norm.

Vεx = εTx (AT

0 P + PA0)εx + 2εTx ∆(y, t)

≤ −|εx|2 + 2|εx|δ(t)

= −|εx|22

− 1

2(|εx| − 2δ(t))2 + 2δ2(t)

≤ −|εx|22

+ 2δ2(t) (3.16)

using (3.15), pmin = mineig(P ), pmax = maxeig(P ) and pmin|εx(t)|2 ≤ Vεx(t) ≤pmax|εx(t)|2. Using comparison lemma, from (3.16) we obtain

Vεx(t) ≤ exp

(

− t

2pmax

)

Vεx(0) + 4

∫ t

0

exp

(

− t− τ

2pmax

)

δ(τ)2dτ (3.17)

⇒ |εx(t)|2 ≤ pmax

pmin

exp

(

− t

2pmax

)

|εx(0)|2 +2

pmin

∫ t

0

exp

(

− t− τ

2pmax

)

δ(τ)2dτ

⇒ |εx(t)| ≤√

pmax

pminexp

(

− t

4pmax

)

|εx(0)| +√

2

pmin

[∫ t

0

exp

(

− t− τ

2pmax

)

δ(τ)2dτ

]

(3.18)

In (3.18), the first term is exponentially vanishing, and the second term is unknown

but bounded. Hence, the state-estimation error remains bounded and converges ex-

ponentially to a residual-ball whose size depends on the extent of unknown modeling

uncertainties i.e.,

|εi| = |xi − xi| ≤ δεi(t) for i = 1, ..., n (3.19)

Page 56: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

50

3.3.2 Parameter Projection

Discontinuous projection, as defined in the previous chapter, will be used in this

work.

3.3.3 Controller Design

The output feedback based controller design presented here combines the output

feedback based adaptive backstepping [31] and discontinuous projection based ARC

[18], which uses state-feedback. The main idea is to synthesize a virtual control law

which will drive the error to a small residual ball. But, as in this case only a single

state is available for measurement, the synthesized virtual control law will replace

the reconstructed state at each step, and the state estimation error will be dealt

with via robust feedback. Also, it should be noted that the use of discontinuous

projection implies a tuning function based backstepping cannot be used, and hence

a stronger robust control law is needed to negate the effects of parameter estimation

transients. For advantages of discontinuous projection based technique over smooth

modifications of adaptive law like smooth projection and, the full-state feedback based

ARC controller design, the reader is referred to [18].

Step 1: The derivative of the output tracking error z1 = y − yd is given by,

z1 = y − yd

= x2 + ϕ0,1(y) +

p∑

j=1

ajϕ1,j(y) − yd + ∆1(y, t)

= x2 + ϕ0,1(y) +

p∑

j=1

ajϕ1,j − yd + ∆1(y, t)

= ω0 + ωTθ + ∆1(y, t)

= κmϑm,2 + ω0 + ωTθ + ∆1(y, t) (3.20)

Page 57: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

51

where

ω0 = [ξ0,2 + ϕ0,1],

ω = [ξ(2) + Φ(1), ϑm,2, ϑm−1,2, .., ϑ0,2,

ψm,1(2), .., ψm,q(2), .., ψ0,1(2), .., ψ0,q(2)]T

ω = ω − e∗p+1ϑm,2

θ = [a1, a2, .., ap, κm, .., κ0,

µm,1, .., µm,q, .., µ0,1, .., µ0,q]T

∆1(y, t) = ∆1(y, t) + ε2 (3.21)

and e∗k is the kth basis vector in Rp+m+qm+1.

If ϑm,2 were the input, we would synthesize a virtual control law α1 to make z1 as

small as possible

α1(y, ξ0, ξ, λm+1, ψi,j,2, θ, t) = α1a + α1s

α1a = − 1

κm

ω0 + ωTθ − yd (3.22)

In (3.22), α1a is the model compensation component of the control law used to achieve

an improved model compensation through on-line parameter adaptation. Thus, the

fault is partly accommodated using model compensation, as

[θp+m+2, ..., θp+m+qm+1]T = [µm,1, .., µm,q, .., µ0,1, .., µ0,q]

T

Now, as we assumed the sign of κm is known, without loss of generality, one can as-

sume κm > 0 and it is lower bounded by a non-zero positive constant i.e., (κm)min =

(θp+1)min > 0 where (θp+1)min is independent of the failure pattern. Also, that

(θp+1)min is known from assumption A4. Then, the projection mapping (5.22) guaran-

tees that κm ≥ (κm)min > 0, which implies that the control law (3.22) is well defined.

Substituting (3.22) into (3.20), we get

z1 = κm(z1 + α1s) − θφ1 + ∆1 (3.23)

Page 58: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

52

The robust component is designed to compensate for the potential destabilizing effect

of the uncertainties on the right hand side of (3.23) as follows

α1s = α1s1 + α1s2 + α1s3, α1s1 = − 1

κm,min

k1sz1 (3.24)

where k1s is a nonlinear gain, such that

k1s = g1 + ||Cφ1Γφ1||2, g1 ≥ 0 (3.25)

in which Cφ1is a positive definite constant diagonal matrix to be specified later. As

discontinuous projection is used, tuning functions cannot be used to compensate for

parameter estimation-error transients.

Substituting (3.25) in (3.23), we obtain

z1 = κmz2 −κm

κm,min

k1sz1 + κm(α1s2 + α1s3) − θTφ1 + ∆1 (3.26)

Define a positive semi-definite (p.s.d) function V1 = 12z21 . Its derivative is given by

V1 ≤ κmz1z2 − k1sz21 + z1(κmα1s2 − θTφ1) + z1(κmα1s3 + ∆1) (3.27)

From assumption A1,

‖ θTφ1 ‖≤‖ θM ‖‖ φ1 ‖, θM = θmax − θmin (3.28)

As ‖ θTφ1 ‖ is bounded by a known function, there exists a robust control function

satisfying the following conditions

(a) z1κmα1s2 − θTφ1 ≤ ǫ11

(b) z1α1s2 ≤ 0 (3.29)

Similarly, from assumption A2 and (3.19), we have

|∆1| ≤ |ε2| + |∆1| = δε2(t) + δ1(t) , δ1(t) (3.30)

Now, we can follow the same strategy as in (3.29) to design a robust control law. But,

as δ1(y, t) is unknown, we cannot prespecify the level of control accuracy. Hence, we

seek to achieve the following relaxed conditions

(a) z1κmα1s3 + ∆1(y, t) ≤ ǫ12δ21

(b) z1α1s2 ≤ 0 (3.31)

Page 59: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

53

Remark 2. Condition (a) of (3.29) shows that α1s2 is synthesized to attenuate the

effect of parametric uncertainties θ with the level of control accuracy being measured

by ǫ11. Condition (b) ensures that α1s2 is dissipative in nature so that it does not

interfere with the functionality of adaptive control law α1a. One smooth example of

α1s2 satisfying (3.29) is

α1s2 = − h1

4κm,minǫ11z1, h1 ≥ ‖θM‖2‖φ1‖2 (3.32)

Similarly, an example of α1s3 satisfying (3.31), which is synthesized to attenuate the

effect of unstructured uncertainties ∆1(y, t), is given by given by

α1s3 = − 1

4κm,minǫ12z1 (3.33)

Remark 3. There are subtle but important design differences between an ARC

Figure 3.1. Structure of ARC and RAC based fault-tolerant controllers

and RAC based fault-tolerant control scheme. Figure (3.1) shows the underlying

structure of ARC and RAC based designs. Note that the ARC based fault-tolerant

controller, the emphasis is on the inner loop robust controller, and the adaptation

mechanism in the outer loop is used to reduce the extent of modeling uncertainties.

As unknown actuator faults introduce severe estimation error (θ) and estimation

error transients (˙θ), it is necessary to suppress their undesirable effect on the system

dynamics. The coordination mechanism ensures that the potential destabilizing effect

Page 60: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

54

of θ and ˙θ are effectively suppressed by the robust controller. Furthermore, the

bounded uncertainties are also attenuated to desired extent by the robust controller.

Thus, desired transient response is guaranteed. On the other hand, the RAC based

designs use an adaptive controller in conjunction with robustness modifications to the

adaptation scheme to guarantee the boundedness of all the signals. They lack the

extra design freedom present in ARC due to the underlying robust controller. Thus,

they cannot guarantee desired transient response.

Step 2: From (3.11-3.13) and (3.20-3.22), we can obtain the derivative of α1 as

follows

α1 = α1c + α1u

α1c =∂α1

∂yω0 + ωT θ +

∂α1

∂ξ0A0ξ0 + ky + ϕ0(y)

+∂α1

∂ξA0ξ + Φ(y) +

m+1∑

i=1

∂α1

∂λi

λi +

q∑

j=1

m+1∑

i=1

∂α1

∂ζi,jζi,j +

∂α1

∂t(3.34)

α1u =∂α1

∂y(−θTω + ∆1) +

∂α1

∂θ

˙θ (3.35)

α1c is calculable and will be used for control function design. α1u, however, is not

calculable and will be dealt with via certain robust terms. From (3.10), the derivative

of z2 = ϑm,2 − α1 is given by

z2 = ϑm,3 − k2ϑm,1 − α1c − α1u (3.36)

Define a p.s.d function V2 = V1 + 12z22 . Then, derivative of V2 using (3.27) and (3.36)

is given by

V2 ≤ V1|α1+ z2κmz1 + ϑm,3 − k2ϑm,1 − α1c − α1u (3.37)

Page 61: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

55

where V1|α1= −k1sz

21 + z1(κmα1s2 − θTφ1) + z1(κmα1s3 + ∆1). Similar to (3.22), we

can now define α2 for ϑm,3 as follows

α2(y, ξ0, ξ, λm+2, ψi,j,3, θ, t) = α2a + α2s

α2a = −κmz1 + k2ϑm,1 + α1c

α2s = α2s1 + α2s2 + α2s3, α2s1 = −k2sz2

k2s ≥ g2 + ‖∂α1

∂θCθ2‖ + ‖Cφ2Γφ2 ‖2 (3.38)

where g2 ≥ 0 is a constant, Cθ2 and Cφ2 are positive definite constant diagonal

matrices, α2s2 and α2s3 are robust control functions to be synthesized later. Due

to use of discontinuous projection, we cannot use tuning functions which anticipates

and compensates for the effect of parameter estimation transients. α2s1 is the robust

control term which compensates for this loss of information. The reason for choosing

this form for α2s1 will become apparent in the proof of theorem 1. Substituting (3.38)

in (3.37), we obtain

V2 ≤ V1|α1+ z2z3 − k2sz

22 + z2(α2s2 − θTφ2) + z2(α2s3 + ∆2) − z2

∂α1

∂θ

˙θ (3.39)

where z3 = ϑm,3 − α2 represents the input discrepancy and

φ2 = e∗n+1z1 −∂α1

∂yω, ∆2 = −∂α1

∂y∆1 (3.40)

From (3.30), it follows that ∆2 ≤ |∂α1/∂y|δ1. Similar to (3.31) and (3.37), the robust

control functions α2s2 and α2s3 are chosen to satisfy

(a) z2(α2s2 − θTφ2) ≤ ǫ21

(b) z2(α2s3 + ∆2) ≤ ǫ22δ21

(c) z2α2s2 ≤ 0 , z2α2s3 ≤ 0 (3.41)

where ǫ21 and ǫ22 are positive design parameters. As in (3.32) and (3.33), α2s2 and

α2s3 can be chosen as,

α2s2 = − h2

4ǫ21z2 , α2s3 = − 1

4ǫ21

(

∂α1

∂y

)2

z2 (3.42)

Page 62: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

56

where h2 is any smooth function satisfying h2 ≥‖ θM ‖2‖ φ2 ‖2. From (3.27) and h2

defined above, the derivative of V2 satisfies

V2 ≤ z2z3 −2∑

j=1

kjsz2j + z1(κmα1s2 − θ1φ1)

+ z1(κmα1s3 + ∆1) + z2(α2s2 − θTφ2) + z2(α2s3 + ∆2) −∂α1

∂θ

˙θz2 (3.43)

Step i (3 ≤ i < ρ): Mathematical induction will be used to prove the general result

for all the intermediate steps. At each step i, the ARC control function αi will be

constructed for virtual control input ϑm,i+1. For any j ∈ [3, i−1], let zj = ϑm,j −αj−1

and recursively design

φj = −∂αj−1

∂yω , ∆j = −∂αj−1

∂y∆1 (3.44)

Lemma 1: At step i, choose the desired ARC control function αi as

αi(y, ξ0, ξ, λm+i, ψk,j,i+1, θ, t) = αia + αis

αia = −zi + kiϑm,i + α(i−1)c

αis = αis1 + αis2 + αis3 αis1 = −kiszi

kis ≥ gi+ ‖ ∂αi−1

∂θCθi ‖ + ‖ CφiΓφi ‖2 (3.45)

where gi > 0 is a constant, and Cθi and Cφi are positive definite constant diagonal

matrices, αis2 and αis3 are robust control functions satisfying,

(a) zi(αis2 − θTφi) ≤ ǫi1

(b) zi(αis3 + ∆i) ≤ ǫi2δ21

(c) ziαis2 ≤ 0 , ziαis3 ≤ 0 (3.46)

and

α(i−1)c =∂α1

∂yω0 + ωT θ +

∂α1

∂ξ0A0ξ0 + ky + ϕ0(y)

+∂α1

∂ξA0ξ + Φ(y) +

m+1∑

i=1

∂α1

∂λi

λi +

q∑

j=1

m+1∑

i=1

∂α1

∂ζi,jζi,j +

∂α1

∂t(3.47)

Page 63: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

57

Then, the ith error subsystem is

zi = zi+1 − zi−1 − kiszi + (αis2 − θTφi) + (αis3 + ∆i) −∂αi−1

∂θ

˙θ (3.48)

and the derivative of the augmented p.s.d function Vi = Vi−1 + 1/2z2i satisfies,

Vi ≤ zizi+1 −i∑

j=1

kjsz2j + z1(κmα1s2 − θTφ1) +

i∑

j=2

zj(κmαjs2

−θTφj) + z1(κmα1s3 + ∆1) +

i∑

j=2

zj(αjs3 + ∆j) −i∑

j=2

∂αj−1

∂θ

˙θzj (3.49)

The lemma can be easily verified by recursively writing the various expressions and

substituting the expressions obtained in step 1 and 2.

Step ρ: In this final step, the actual control law u∗ will be synthesized such that

ϑm,ρ tracks the desired ARC control function αρ−1. The derivative of zρ can be

obtained as

zρ = ϑm,ρ+1 + u∗ − kρϑm,1 − α(ρ−1)c −∂αρ−1

∂y(−θTω + ∆1) −

∂αρ−1

∂θ

˙θ (3.50)

If ϑm,ρ+1 + u∗ were the virtual input, (3.50) would have the same form as the inter-

mediate step i. Therefore, the general form, (3.44-3.50) applies to step ρ. Since u∗ is

the actual control input, it can be chosen as,

u∗ = αρ − ϑm,ρ+1 (3.51)

where αρ is given by (3.49). Then, zρ+1 = u∗ + ϑm,ρ+1 − αρ = 0.

Theorem 1. Let the parameter estimates be updated using adaptation law (5.20)

in which τ is chosen as

τ =

ρ∑

j=1

φjzj (3.52)

If diagonal controller gain matrices Cθj and Cφk are chosen such that c2φkr ≥ ρ

4

∑ρ

j=1 1/c2θjr,

where cθjr and cφkr are the rth diagonal element of Cθj and Cφk respectively. Then,

the control law (3.51) guarantees that,

Page 64: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

58

1. In general the control input and all internal signals are bounded. Furthermore,

Vρ is bounded above by,

Vρ(t) ≤ ǫρ1 + ǫρ2 ‖ δ1 ‖2∞

λρ

[1 − exp(−λρt)] (3.53)

where λρ = 2ming1, . . . , gρ, ǫρ1 =∑ρ

j=1 ǫj1, ǫρ2 =∑ρ

j=1 ǫj2 and ‖ δ1 ‖2∞

stands for the infinity norm of δ1.

2. If after a finite time t0, ∆(y, t) = 0 (i.e., in the presence of parametric uncer-

tainties only) then, in addition to results in (3.53), asymptotic output tracking

control is also achieved.

Proof. The proof is similar to that of theorem 1 of chapter 1.

Remark 4. In context of actuator fault compensation, first part of theorem 1

guarantees that the jump in parameter values due to failed actuator does not interfere

with the desired transient performance. By using trajectory initialization techniques

[17], we can set Vρ(0) = 0 and then, from (3.52) we have

|z1(t)| ≤√

2(ǫρ1 + ǫρ2 ‖ δ1 ‖2∞)

λρ

[

1 − exp(−λρt)

2

]

(3.54)

This equation provides an upperbound for the output tracking error z1 and charac-

terizes the transient response. ǫρ1, ǫρ2 and λρ are controller parameters which we

can choose. Note that from remark 1, we know by properly tuning ǫρ2 we can make

ǫρ2 ‖ δ1 ‖2∞ as small as desired. Thus, we can tune the parameter ǫρ2 to make |z1(t)|

smaller than any predetermined bound. In this sense, we have guaranteed transient

response. This result on transient response of the system is a direct consequence

of underlying robust filter structure of the ARC controller. The second part of the

theorem guarantees asymptotic output tracking in presence of actuators failures and

parametric uncertainties only i.e., ∆i(y, t) = 0. As desirable properties of robust and

adaptive control designs are preserved, we get desired transient response with small

steady-state tracking error in spite of actuator faults.

Page 65: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

59

3.4 Simulation Example: A Nonlinear Hypersonic Aircraft Model

The proposed scheme is implemented on a nonlinear longitudinal model of hy-

personic aircraft cruising at a velocity of 15 mach, at an altitude of 110, 000 feet.

The control objective is to track a desired trajectory in presence of various modeling

uncertainties and elevator segment failures. Nominal model of the system is

α = q − γ

q =Myy

Iyy

γ =L+ T sinα

mV− (µ− V 2r) cos(γ)

V r2(3.55)

where

α = angle of attack, rad

γ = flight-path angle, rad

V = velocity, ft/sec

q = pitch rate, rad/sec

T = thrust, lbf

L = lift, lbf

Myy = pitching moment, lbf.ft

Iyy = moment of inertia, slug.ft2

Details of the full order model can be found in [37]. Note that the same reduced order

model was used by Tang et al. [5] and thus will provide a platform to compare the

robust adaptive control (RAC) based fault-tolerant control (FTC) and the adaptive

robust fault-tolerant control (ARFTC) schemes. The nominal model of the system

does not take into account any unstructured modeling uncertainties and external

disturbances. As modeling uncertainties are inherent to any realistic system model,

Page 66: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

60

unmatched uncertainties will be introduced in order to make the simulation studies

more meaningful. The state-space representation is given by

x1 = x2 + a1y + a2 sin(y) + a3y2 sin(y) + a4 cos(x3) + ∆1(y, t)

x2 = a5y2 + a6y + (a7 + a8y + a9y

2)x2 + b1u1 + b2u2 + ∆2(y, t)

x3 = a10 cos(x3) − a1y − a2 sin(y)

y = x1 (3.56)

where [x1, x2, x3] = [α, q, γ] and ∆i represents uncertain nonlinearities and distur-

bances. The numerical values of the nominal plant parameters are

a1 = −0.0427, a2 = −3.4496×10−4, a3 = 5×10−5, a4 = 0.0014, a5 = −4.2006,

a6 = 1.0821, a7 = −3.6896, a8 = 0.1637, a9 = −0.1242, a10 = 0.0014, b1 = 0.8, b2 = 0.8

The initial conditions are set to x(0) = [0, 0.01, 0]T . Due to the presence of the

term a4 cos(x3) in x1, (5.64) is not in the output-feedback form. However, since

|a4 cos(x3)| ≤ a4, it can be considered a bounded uncertainty and can be dealt with

using robust feedback.

The reference command chosen, yd(t) = 0.01 sin(0.1t) is in accordance with [5].

Details of the ARFTC controller is given at the end of this section. Details of RAC

based FTC can be obtained from [5]. Additionally, the controller parameters for

ARFTC scheme were chosen such that the control input profiles would be comparable

for both schemes.

Using the parametrization (3.3), before actuator faults the actual control signal is

given by

u1(t) = 1b1u∗(t) rads

u2(t) = 1b2u∗(t) rads

for t < 50 secs

Page 67: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

61

Two faults are introduced: at t = 50 seconds the first actuator loses 40% efficiency

and at t = 75 seconds, the second actuator gets stuck at u2 = 0.1 radians. These can

be represented in terms commanded control signals as follows

u1(t) = 0.6b1u∗(t) rads

u2(t) = 1b2u∗(t) rads

for t ≥ 50 secs

and

u1(t) = 0.6b1u∗(t) rads

u2(t) = 0.1 rads

for t ≥ 75 secs

Two cases are considered in order to illustrate the effectiveness of the proposed

scheme. In the first set of simulations, all unstructured modeling errors and distur-

bances are assumed to be zero i.e., ∆i = 0 for i = 1, 2. This provides a level ground to

compare the transient performance of the two schemes after an actuator fails. After

the first fault, the tracking error remains close to zero for both schemes. However,

after the second fault, which is more severe as compared to the first one, tracking

error remains close to zero for ARFTC, but deviates significantly for RAC based FTC

and can be explained as follows. The second fault causes a large jump in parameter

value which can lead to significant transient tracking error. In the ARFTC scheme,

however, the robust component of the control law is designed to suppress the effect

of such parameter jumps (αis2 in equations (3.29), (3.31), (3.41) and (3.46)). An

explicit upper bound for the transient tracking error can also be obtained in terms

of controller parameters when an upper bound for the unstructured uncertainties is

known a priori. RAC based FTC, on the other hand, can only guarantee the bound-

edness of all signals at best. Also, as a4 cos(x3) is present even when when other

uncertainties are set to zero, asymptotic tracking cannot be achieved using either

of the schemes. But, in ARFTC, its effect on the final tracking accuracy can be

attenuated by adjusting the gain of the robust component of the control law in a

transparent manner. Another interesting observation is that the steady-state track-

ing error is small in ARFTC without control chattering. Any scheme which relies on

Page 68: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

62

robust control only, would either lead to control input chattering in presence of such

large parametric uncertainties, or result in poor steady-state tracking accuracy due

to smoothing techniques. In contrast, the learning mechanism of ARFTC leads to

improved model compensation and guarantees acceptable steady-state tracking error,

as seen from Fig.1. Thus, desired transient response is guaranteed due to robust

filter structure of the ARFTC. Smaller steady-state tracking error is also achieved

due to the combined effect of reduction in modeling uncertainties through parameter

adaptation, as well as the use of gain to attenuate the effect of bounded uncertainties

on steady-state error.

A second set of simulations were performed where disturbances were introduced

to test the performance of the two schemes in presence of unstructured bounded

uncertainties. We set ∆1(y, t) = 0.01 sin(2t) and ∆2(y, t) = 0.01 sin(3t). Although,

these disturbances are chosen for illustrative purposes only, it is worth mentioning that

many disturbances which affect the aircraft dynamics can indeed be modeled using

harmonic basis functions e.g., wind-shear [35]. Faults and controller parameters are

same as used in the previous case. As evident from Fig.2, the performance of the

RAC based scheme deteriorates significantly in presence of unstructured modeling

uncertainties, which are inherent to any realistic system model. From the tracking

error plots, we see that not only the transient error, but the steady-state error is also

unacceptably large. The limitation of adaptive schemes in presence of disturbances,

even with robustness modifications, becomes obvious by comparing Fig.1 and Fig.2.

It can be explained as follows. Although, parameter projection is used in the RAC

based scheme, it can neither improve the transient performance, nor can it attenuate

the effect of disturbances on the steady-state tracking error. On the other hand, in

ARFTC, the parameter bounds are used not only for projection, but also to design

the robust component of the control law. The parametric uncertainty (which also

accounts for the jump in parameter values) and unstructured uncertainty bounds

are incorporated in the design of the baseline robust controller in ARFTC (αis3 in

equations (3.29), (3.31), (3.41) and (3.46)), guaranteeing desired transient response

Page 69: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

63

and acceptable steady-state tracking error. Therefore, the achievable performance

using the proposed scheme is superior to that of RAC based schemes.

ARC based fault-tolerant controller

The unknown parameter vector θ is

θ = [a1, a2, a3, a5, a6, κ0, µ0]T

and the initial values and bounds for the parameter estimates are

θ(0) = [−0.05,−4 × 10−4, 0,−4.0, 0.9, 1.5, 0]

θmin = [−0.06,−5 × 10−4, 4 × 10−4,−5, 0.8, 0.2,−1]

θmax = [−0.03,−2 × 10−4, 7 × 10−4,−3.5, 1.2, 2, 1]

The gain matrix for parameter estimation is given by Γ = diag1, 1, 1, 1, 1, 5, 0.1.The observer gain matrix is chosen to be k = [2, 1]T and the filters used for state-

reconstruction are

ξ0 = A0ξ0 + ky + ϕ2(y) − yϕ2(y)e1

ξ = A0ξ + Φ(y) + ϕ2(y)ξ

ϑ0 = A0ϑ0 + e2u∗ + ϕ2(y)ϑ0

ψ = A0ψ + ϕ2(y)ψ + e2 (3.57)

where

Φ(y) =

y sin(y) y2 sin(y) 0 0

0 0 0 y2 y

ϕ2(y) = a7y2 + a8y + a9

Page 70: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

64

Initial condition for all filter states are set to 0. The control law is given by

α1 = α1a + α1s, z1 = y − yd

α1a = − 1

κ0(ω0 + ωT θ − yd), α1s = −k1sz1

α1c =∂α1

∂y(ω0 + ωT θ) +

∂α1

∂ξ0,2

˙ξ0,2 +∂α1

∂ξ(2)ξ(2)

+∂α1

∂ϑ0,2

ϑ0,2 +∂α1

∂ψ2

ψ2 +∂α1

∂t

u∗ = u∗a + u∗s, z2 = ϑ0,2 − α1

u∗a = k2ϑ0,1 + α1c − κ0z1, u∗s = −k2sz2 (3.58)

where k1s = −50, k2 = 1 and k2s = −90. The parameter update law is given by

(5.20) where

τ = z1φ1 + z2φ2

φ1 = ωT + eT6 α1a, φ2 = eT

6 z1 −∂α1

∂yω (3.59)

Page 71: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

65

0 20 40 60 80 100 120 140−0.02

−0.015

−0.01

−0.005

0

0.005

0.01

0.015

0.02

time (sec)

y an

d y

m (rad

/sec

)

Plant output and reference signal

RefRACARC

0 20 40 60 80 100 120 140−0.01

−0.005

0

0.005

0.01

time (sec)

e(t)

(rad/

sec)

Tracking error

RACARC

0 20 40 60 80 100 120 140−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

time (sec)

u(t)

(rad)

Control signals: RAC

u1(t)

u2(t)

0 20 40 60 80 100 120 140−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

time (sec)

u(t)

(rad)

Control signals: ARC

u1(t)

u2(t)

Figure 3.2. Reference tracking, control signals and tracking error for RACversus ARC based fault-tolerant schemes in absence of disturbances

Page 72: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

66

0 20 40 60 80 100 120 140−0.03

−0.02

−0.01

0

0.01

0.02

0.03

time (sec)

y an

d y

m (rad

/sec

)

Plant output and reference signal

RefRACARC

0 20 40 60 80 100 120 140−0.02

−0.01

0

0.01

0.02

0.03

time (sec)

e(t)

(rad/

sec)

Tracking error

RACARC

0 20 40 60 80 100 120 140−0.4

−0.3

−0.2

−0.1

0

0.1

0.2

0.3

time (sec)

u(t)

(rad)

Control signals: RAC

u1(t)

u2(t)

0 20 40 60 80 100 120 140−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

time (sec)

u(t)

(rad)

Control signals: ARC

u1(t)

u2(t)

Figure 3.3. Reference tracking, control signals and tracking error for RACversus ARC based fault-tolerant schemes in presence of disturbances

Page 73: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

67

3.5 Conclusion

In this work, an adaptive robust output feedback based scheme is presented for

unknown actuator fault accommodation for a class of uncertain nonlinear system.

Adaptation and robust feedback are used simultaneously to maintain tracking perfor-

mance in face of large parametric uncertainties introduced due to failing actuators,

exogenous disturbances and other modeling uncertainties.

A nonlinear model of hypersonic aircraft is used for simulation studies, which

clearly demonstrates the effectiveness of the proposed scheme in accommodating ac-

tuator faults. In summary, some of the salient features of the fault accommodation

scheme presented in this chapter are,

1. capability to handle large parametric uncertainties due to unknown actuator

failures like stuck actuators and actuator loss in efficiency with guaranteed tran-

sient performance

2. guaranteed robust performance when adaptation is switched off

3. calculable upper bound for tracking error based on controller parameters and

ability to achieve prespecified final tracking accuracy

Page 74: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

68

4. A BACKSTEPPING BASED APPROACH TO ROBUST GLOBAL

STABILIZATION OF A CHAIN OF INTEGRATORS WITH INPUT

SATURATION

4.1 Introduction

In the previous two chapters, an adaptive robust fault-tolerant scheme was devel-

oped to accommodate unknown actuator faults. An important underlying assumption

was that the healthy actuators had sufficient actuator authority to compensate for

the unknown actuator faults. In reality, however this may not be the case. In fact,

achieving small transient tracking error in presence of actuator faults may require

large control signals, which can easily saturate the healthy actuators. Thus, it is ex-

tremely important that the actuator limits be taken into consideration while designing

an actuator fault-tolerant controller.

As a first step towards developing a saturated fault-tolerant controller, we develop

a backstepping based controller for a celebrated problem - stabilization of a chain of

integrators in presence of input saturation. There are two important reasons for

the popularity of this problem within the input saturation community. First, an

integrator chain is not a stable system. For a long time, most researchers focused

on global stabilization of stable systems only. In fact, it was shown by Sontag in [],

that a third or higher order integrator chain cannot be stabilized by linear control

laws. Thus, integrator chain is complex enough that simple linear controllers cannot

work, but at the same time, owing to its simple form, the analysis remains clear and

tractable. Second, many schemes which were proposed for stabilizing the integrator

chain have been successfully generalized to classes of nonlinear systems.

Among various approaches for dealing with input saturation, anti-windup schemes,

nested saturation functions and model predictive control are most popular. In anti-

Page 75: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

69

windup based schemes ( [38], [39], [40], [41]), first a controller is designed without

any regard to the actuator limits, and then a modification is introduced to minimize

the adverse effects of saturation. But, such designs tend to be more conservative in

terms of achievable performance. Model predictive control (MPC), which involves

solving an open-loop optimization problem at each step, is adept at dealing with

hard constraints and is fast becoming a useful tool in dealing with saturation prob-

lems [42], [43]. The main drawback and challenge of MPC based schemes is to incor-

porate the modeling uncertainties, which are inherent to any realistic system model.

Lyapunov function based techniques for stabilizing a system in feedforward ( [44])

and feedback form ( [45], [46]) with bounded control have also been proposed. Note

that backstepping based approaches suffer from the so called “explosion” of terms,

and showing boundedness in such a setting is very challenging. This problem was

circumvented by imposing some restrictions on the growth rate of various nonlinear-

ities and assuming that bounded control laws and control Lyapunov functions are

known a priori for the reduced order system. However, they do not provide very

clear and practical guidelines for designing a controller in presence of uncertainties.

Nested saturation functions and small-gain theorem are amongst the most widely

used tools for dealing with input saturation. This methodology was first proposed by

Teel in [1] for a chain of integrators. The first step of the design involved a coordinate

transformation, which transformed the system to a feedforward form. In the second

step, saturation functions were used to construct a nested control law in terms of the

transformed coordinates. Subsequently, this approach has been extended to various

classes of nonlinear systems in feedforward form under various assumptions [47], [48].

For chain of integrators, many modifications to Teel’s original design have been

proposed to improve the transient response and robustness of the controller [49], [50],

[51]. But, as all of these designs were based on [1], they also inherited the limitations

of that approach. Particularly, in presence of bounded input disturbance, the region

where the controller is unsaturated shrinks drastically, and in the worst case, renders

the controller design impossible even when magnitude of the disturbance is less than

Page 76: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

70

the available control input. This is to be expected as all the transformed coordinates

depend on the last state of the integrator chain, the dynamics of which includes the

input disturbance. Thus, the input disturbance affects the dynamics of every state

in the transformed coordinates. This implies that the parameters for all saturation

functions need to be chosen conservatively to accommodate the effect of disturbance.

Furthermore, the coordinate transformation makes it more difficult to tune the con-

troller parameters to achieve a desired objective. For example, say the objective is for

the output x1 to track yd with desired transient response in the unsaturated region.

In Teel’s approach, as x1 depends on all the transformed coordinates yi, i = 1, ..., n,

the tracking problem for x1 translates to designing a controller with desired transient

response for all the states - a more challenging problem which can further add to

the conservativeness of the overall design procedure. Moreover, it is expected that

in presence of input saturation, there will be trade-off between high controller gains

and the region where it operates unsaturated. The coordinate transformation makes

it difficult to understand these trade-offs and thus, reduces the design flexibility.

From the preceding discussion, it should be clear that the design conservativeness

in presence of disturbances can be reduced and it can be made more transparent if

a controller can be designed without the coordinate transformation. In the present

work, we take a conceptually different approach and solve the problem without using

coordinate transformation. We combine a backstepping based approach with satura-

tion functions to design a globally stabilizing controller for a chain of integrators. As

no transformation is required, the input disturbance appears only in the last channel

and thus, affects the choice of only the last saturation function parameters. Also,

in the proposed technique, bounded virtual control laws are designed at each step

to drive the intermediate error variable to zero. This makes the design more trans-

parent, as we need not deal with a combination of states but only one state at each

step of the design to achieve the desired transient response. In fact, the proposed

scheme not only preserves the clarity present in backstepping based designs, but also

Page 77: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

71

allows design of stabilizing controllers in presence of bounded disturbances when a

coordinate transformation based approach fails.

In addition to being less conservative, the proposed design can also guarantee

desired closed-loop performance within the unsaturated region of controller operation.

However, this implies that the controller gains be selected in a fashion that allows the

closed-loop poles to be placed at desired locations. Thus, the procedure for choosing

the controller parameters become a non-trivial and fairly challenging problem, as they

need to satisfy two sets of constraints - first, a set of inequalities which guarantee that

all the states can be steered to the unsaturated region of controller operation, and

second, a set of constraints which ensure that desired closed-loop performance can be

achieved. To this end, we provide necessary and sufficient conditions for the existence

of the proposed control law, and propose a systematic way of choosing the controller

parameters based on rigorous analysis.

Comparative simulation studies are presented in this chapter show the effective-

ness of the proposed scheme. Note that for comparative studies, we use the technique

proposed in [50], which is based on coordinate transformation. The results demon-

strate that better disturbance rejection capabilities and faster convergence rate can

be achieved using the proposed design.

4.2 Motivation and Problem Formulation

Consider a chain of integrator with input disturbance

x1 = x2

x2 = x3

...

xn = u+ d(t)

y = x1 (4.1)

Page 78: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

72

where |u| ≤ uM , |d(t)| ≤ dM and dM is a known constant. The objective of the

present work is to design a controller such that

1. x1 tracks yd with steady-state error |ess| ≤ δ

2. ess(t) should have desired transient performance in the unsaturated region.

We will make the following assumption regarding the extent of input disturbance

and the reference trajectory

A1: The extent of disturbances and the desired trajectory is such that

dnyd

dtn

≤ λM

uM > dM + λM (4.2)

Let us first investigate the effect of this disturbance on Teel’s approach. In all such

approaches, the coordinate transformation used takes the following generic form [49]

yi =

n∑

j=i

tij xj , j = 1, . . . , n

(4.3)

where xj = xj − y(j−1)d and T = [tij] is the transformation matrix given by

tij =

1, 1 ≤ i ≤ n, j = n

0, i = n, 1 ≤ j ≤ n− 1∑n

m=i+1 kmtm,j+1, i ≥ 1, j ≤ n− 1

With this, the dynamics of the transformed states can be expressed as

y1

...

yn−1

yn

=

0 k1 · · · 1...

. . .. . .

...

0 · · · 0 1

0 · · · · · · 0

y1

...

yn−1

yn

+

1...

1

1

u(t) +

1...

1

1

d(t) (4.4)

The proposed control law takes the form [1]

u = −σn(knyn + σn−1(kn−1yn−1 + ...+ σ1(k1x1))...) (4.5)

Page 79: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

73

where the saturation functions σi are given by

(1) sσi(s) ≥ 0, ∀s 6= 0

(2) σi(s) = s, ∀|s| < Li

(3) |σi(s)| ≤ Mi, ∀s ∈ R, Li < Mi (4.6)

To account for the disturbances, the inequalities proposed in [1] can be modified as

follows to ensure that there is available control input for steering the tracking error

to a neighborhood of zero

|y(n)d | ≤ Ln+1 −Mn

Li+1 ≥ 2Mi + dM , i = 0, ..., n− 1 (4.7)

with M0 = 0. From (4.4), we see that due to the coordinate transformation, the

disturbance affects all the states, although it was present only in the input channel in

the original coordinates. Consequently, the linear unsaturated region (i.e., the region

where the linear controller is unsaturated) of all the transformed states shrink by

dM , as seen from (4.7). This makes the design procedure very conservative. In fact,

even when the extent of uncertainties is less than the available control input, the

coordinate transformation may render the design of a stable controller impossible, as

shown below

uM > Mn+1 > Ln+1

> Mn > Ln

> dM + 2Mn−1 > dM + 2Ln−1

> dM + 2(dM + 2Mn−2) > dM + 2(d+ 2Ln−2)...

...

> dM + 2dM + · · ·+ 2n−1dM = (2n − 1)dM

(4.8)

From the above inequalities, it should be clear that if d(t) is such that dM ≥ uM/(2n − 1),

then this approach would not work. For example, consider the stabilization problem

for a third order chain of integrator with uM = 10, and dM = 4. Then, dM = 4 > 10/7,

Page 80: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

74

which implies (4.8) cannot be satisfied. This shows the conservativeness of any ap-

proach which relies on coordinate transformation. In this work, we will show that such

conservativeness can be removed by using a backstepping based technique, which does

not require coordinate transformation. In contrast to the condition uM > (2n − 1)dM

given by (4.8), we need a much less restrictive condition uM > dM for the stabilization

problem, as given by assumption A1.

4.3 Main Result

In this section, we present the backstepping based controller design for robust

global tracking for an integrator chain and present the main results. In the first

step, a set of inequalities is proposed such that when satisfied, it would guarantee

that for any given set of initial conditions, the error can be driven to an invariant

set where the proposed controller is not saturated. Once within the unsaturated

region, the desired properties of a linear controller, e.g. exponential convergence,

desired transient response and arbitrary disturbance attenuation, can be guaranteed.

However, this implies that the chosen controller gains should not only satisfy the

inequalities proposed in the first step, but should also be able to place the closed-loop

poles at the desired location. In the second step, we provide necessary and sufficient

conditions for the existence of such a control law, and also present a systematic way

of choosing the controller gains to achieve the desired closed-loop performance.

Page 81: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

75

4.3.1 Convergence to Unsaturated Region

We will use the coordinate transformation xi = xi−y(i−1)d to simplify the analysis.

The designed virtual control law αi, error variables zi and the xi-dynamics in terms

of αi and zi are given by

zi = xi − αi−1(zi−1)

αi = −σi(zi)

˙xi = zi+1 + αi (4.9)

where the σi(zi), i = 1, ..., n (see figure (5.1)) are saturation functions.

Remark 1. Note that the form of the virtual control laws designed is signifi-

cantly different, and at the same time much simpler than the typical “cancellation”

backstepping design. In cancellation backstepping design, the detrimental effect of

the virtual control law of step i− 1 i.e., αi−1 is completely canceled at step i, by in-

corporating the appropriate terms in the design of αi. This results in a simple linear

dynamics for the error variables, such that the closed-loop poles can be placed at any

desired location. But, this simplicity comes at the expense of a complicated control

law, with exponentially growing number of terms as the order of the system increases.

In the context of bounded control, it is extremely difficult to show the boundedness

of such a control law. For chain of integrators, as we shall see, the error dynamics can

be stabilized with a much simple control law σi(zi), and without canceling the terms

resulting from the derivative of αi−1. Although linear, the resulting error-dynamics

without cancellation leads to a set of constraints that the controller parameters must

satisfy for global stabilization. This makes the parameter selection process slightly

more complicated than the cancellation backstepping design.

The saturation functions used in this work play a significant role in guaranteeing

the overall system stability. In fact, it is only through a careful and rigorous analysis

of the effects of saturation function parameters on system dynamics that a set of

Page 82: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

76

conditions will be derived, which will ensure all desired objectives can be achieved.

The saturation functions are defined as

(a) ziσi(zi) ≥ 0, ∀zi (4.10)

(b) σi(zi) = kizi, ∀|zi|≤li,

σi(zi) = Mi, ∀|zi|≥Li (4.11)

(c) |σi(zi)| ≤Mi, ∀zi (4.12)

(d)∂σi

∂zi

≤ ki, ∀zi (4.13)

Also, li = βiLi with βi ≤ 1, and Mi = kili(1 + γi) with γi > 0. The interval for

zi is divided into three different regions - Ωi1 = zi : |zi| ≤ li, Ωi2 = zi : |zi| ≤Li and Ωi3 = zi : |zi| > Li. Note that the nonlinear transition region of the

saturation function (Ωi2\Ωi1) needs to be at least second order differentiable, as the

backstepping design involves taking derivatives of σi. σn, however, need not have a

smooth transition region, as this appears in the last step. Hence, we choose γn = 0,

βn = 1 for σn, which implies Ωn1 = Ωn2. Substituting (4.9) in (4.1), the error dynamics

Figure 4.1. Saturation function

Page 83: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

77

can be written as

z1 = z2 − σ1(z1)

· · ·

zi = zi+1 − σi(zi) +i−1∑

j=1

[

j∏

r=1

∂σi−r

∂zi−r

]

(zi−j+1 − σi−j(zi−j))

· · ·

zn = u+ d(t) − y(n)d +

n−1∑

j=1

[

j∏

r=1

∂σn−r

∂zn−r

]

(zn−j+1 − σn−j(zn−j))

(4.14)

Now, we are ready to state the main result of the present work.

Theorem 1. Consider system (4.14). Let the control input be

u = y(n)d − σn(zn) (4.15)

If the controller parameters can be chosen such that

kili > li+1 + ki−1Ni, i = 1, 2, ..., n− 1 (4.16)

knln > kn−1Nn + dM (4.17)

knln ≤ uM − λM , u′M (4.18)

where

Ni , Li +Mi−1 +i−2∑

j=1

[

(

j∏

r=1

ki−1−r)(Li−j +Mi−1−j)

]

and k0 , 0, ln+1 , 0. Also, Mn used in the design of σn(zn) is chosen such that

λM + Mn = uM . Then, for any set of initial conditions, all states reach the linear

unsaturated region i.e., Ωi1 in a finite time.

Proof. Consider the following claims.

Claim1. For any initial condition zn(0)

(a) if |zn(0)| > Ln, then it reaches Ωn1 in a finite time

(b) if zn(0) ∈ Ωn1, then zn(t) ∈ Ωn1, ∀ t > 0

Claim 2. Assume |zi+1| ≤ li+1, then

(a) if |zi(t0)| > Li, then it reaches Ωi2 in a finite time

(b) if zi(t0) ∈ Ωi2\Ωi1, then it reaches Ωi1 in a finite time

(c) if zi(t0) ∈ Ωi1, then zi(t) ∈ Ωi1 ∀ time t > t0

Page 84: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

78

Assume both claims are true. Then, from claim 1, we have |zn| < Ln = ln after

a finite time. Theorem 1 then follows from a recursive application of claim 2 to the

intermediate states zj , j = n − 1, · · · , 1. Thus, we will be done if we can show that

both the claims are true.

Proof of Claim 1, Part (a). Without loss of generality (w.l.o.g), assume

zn(0) ≥ Ln. Then, u = −Mn + y(n)d and from (4.9) we get

˙xn = −Mn + d(t)

⇒ xn(t) ≤ xn(0) − (Mn − dM)t

⇒ zn(t) + αn−1(t) ≤ zn(0) + αn−1(0) − (Mn − dM)t

⇒ (Mn − dM)t ≤ zn(0) − zn(t) + αn−1(0) − αn−1(t)

Using |αn−1| ≤Mn−1, and substituting zn(t) = Ln an upperbound for the time taken

to reach Ωn1 is given by

tn32 ≤zn(0) − Ln + 2Mn−1

Mn − dM

(4.19)

From assumption A1, it is easy to see that as Mn = uM − λM > dM , zn(t) reaches

Ωn1 in a finite time.

Remark 2. Before we proceed further, we note the following facts about zn−dynamics

1. zi, for i < n, affects the zn−dynamics only through bounded terms like σi,∂σi

∂zi

and multiplication of ∂σi

∂ziwith other bounded terms. Note that ∂σi

∂zizi is also

bounded, as it vanishes for |zi| > Li.

2. zn is also upper bounded by Ln in Ωn1.

As zn-dynamics is affected by only bounded terms once |zn| ≤ Ln, it is possible to

use a bounded feedback which can make the tangent vector point inwards at Ωn1

boundary, which is essential in proving the next part of the claim.

Part (b). From (4.17), we know there exists δn > 0 such that

kn(ln − δn) =n−1∑

j=1

[

(

j∏

r=1

kn−r)(Ln−j+1 +Mn−j)

]

+ dM (4.20)

Page 85: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

79

In order to show that any trajectory starting in Ωn1 will never leave this set, we need

to show that the tangent vector points inward at the boundaries i.e., at zn = ±Ln.

Assume w.l.o.g zn > 0 Then, at zn = Ln, znzn is given by

znzn = zn

(

−σn(zn) +n−1∑

j=1

[

j∏

r=1

∂σn−r

∂zn−r

]

(zn−j+1 − σn−j(zn−j))

+ d(t)

)

≤ |zn|(

−knln +n−1∑

j=1

[

(

j∏

r=1

kn−r)(Ln−j+1 +Mn−j)

]

+ dM

)

(4.21)

where we have used ziσi(zi) > 0, ∀ zi. Thus, combining (4.20) and (4.21) we get

znzn ≤ −|zn|knδn < 0 (4.22)

This shows that the tangent-vector points inward at the boundaries and completes

the proof of the claim.

Proof of Claim 2, Part (a). When zi ∈ Ωi3 i.e., |zi| > Li, we have αi = −Mi

(w.l.o.g assume zi > 0) from the definition of σi. Using the assumption of claim 2

i.e., |zi+1|≤li+1 and substituting αi = −Mi in (4.9), we obtain

˙xi = zi+1 + αi

≤ li+1 −Mi = −(Mi − li+1)

⇒ xi(t) ≤ xi(t0) − (Mi − li+1)(t− t0)

⇒ (Mi − li+1)(t− t0) ≤ xi(t0) − xi(t) + αi−1(t0) − αi−1(t)

⇒ ti32 ≤ xi(t0) − Li + 2Mi−1

Mi − li+1(4.23)

where we substituted zi(t) = Li and ti32 , t − t0, the time taken to reach Ωi2 from

Ωi3. It follows from (4.16) that Mi > kili > li+1. This proves the first part of the

claim.

Part (b). To prove the next two parts, we proceed as done in the proof of first

claim. From (4.16), there exists a δi such that

ki(li − δi) = li+1 + ki−1Ni (4.24)

Page 86: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

80

Without loss of generality, assume zi(t0) > 0. Then, from the zi-dynamics (4.14), we

have

zi = zi+1 − σi(zi) +i−1∑

j=1

[

j∏

r=1

∂σi−r

∂zi−r

]

(zi−j+1 − σi−j(zi−j))

≤ li+1 − kili +i−1∑

j=1

[

(

j∏

r=1

ki−r)(Li−j+1 +Mi−j)

]

≤ li+1 − kili + ki−1Ni (4.25)

= −kiδi

⇒ ti21 ≤ Li − likiδi

(4.26)

where we have used the fact that Li − li≥|zi(t) − li|, ∀ zi(t) ∈ Ωi2\Ωi1. Thus, there

exists a finite time-interval ti21, within which any trajectory starting in Ωi2\Ωi1 reaches

Ωi1. This completes the proof of second part of the claim.

Part (c). In order to show that any trajectory starting in Ωi1(t0) stays there ∀t > t0, we need to show that zizi < 0 at zi = ±li. For zi = li, we know σi(zi) = kili

and using the assumption |zi+1|≤li+1, from (4.14) the zi-dynamics can be written as

zizi = zi

(

zi+1 − σi(zi) +i−1∑

j=1

[

j∏

r=1

∂σi−r

∂zi−r

]

(zi−j+1 − σi−j(zi−j))

)

≤ |zi|(

li+1 − kili +i−1∑

j=1

[

(

j∏

r=1

ki−r)(Li−j+1 +Mi−j)

])

≤ −ki|zi|δi (4.27)

This shows that Ωi1 is positively invariant. This completes the proof.

4.3.2 Controller Parameter Selection

There are two important questions which need to be answered next. First, when

can the existence of a solution to inequalities (4.16)-(4.18) be guaranteed. Note that

even when such a solution exists, it only guarantees the convergence of all intermediate

error variables to the linear unsaturated region (an invariant set in the neighborhood

of zero). The system response within the linear unsaturated region is governed by the

Page 87: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

81

gains ki, and leads to the second important question - how to select the gains such

that the desired closed-loop performance can be achieved, without violating (4.16)-

(4.18). It is natural to first select a set of gains ki such that desired performance

can be achieved, and then seek a set of lis such that inequalities (4.16)-(4.18) can

be satisfied. However, such an approach may not work as after fixing the gains ki,

solving the inequalities (4.16)-(4.18) may result in negative lis, which will make the

control law unrealizable. Hence, in the following section, we first state and prove

the main result regarding the necessary and sufficient condition for the existence of a

solution to the inequalities. Then, a systematic way of choosing the controller gains in

accordance with this theorem is proposed, such that desired closed-loop performance

can be achieved.

Necessary and Sufficient Conditions for the Existence of Controller Pa-

rameters

After a series of derivations, (4.16)-(4.17) can be rewritten in a matrix form

AL < D, (4.28)

where L = [l1, l2, · · · , ln−1, ln]T , D = [0, 0, · · · , 0,−dM ]T . And A is a function of ki

given by

A =

−k1 1 0 · · · 0

k21(1 + γ1)

(

k1

β2− k2

)

1 · · · 0

· · · · · · · · · · · · · · ·an1 an2 an3 · · ·

(

kn−1

βn− kn

)

, (4.29)

where

aij = ki−1kj

i−j−1∏

r=1

ki−r−1(1 + γj) + ki−1

i−j∏

r=1

ki−r−11βj, ∀i > j.

aij = −ki + ki−1

βi, ∀i = j

aij = 1, ∀j = i+ 1

aij = 0, ∀j > i+ 1

(4.30)

Page 88: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

82

For any γi > 0 and 0 < βi ≤ 1, if we fix a set of positive kis, then the control law

is feasible if and only if the there exist l1, l2,..., ln > 0 such that (4.18) and (5.44)

are satisfied. In other words, at least one solution to (4.18) and (5.44) should lie in

the region (l1, · · · , ln) : li > 0. The following theorem gives the necessary and

sufficient condition for kis such that the control law is feasible.

Theorem 2. For any γi > 0 and 0 < βi ≤ 1, with a set of positive kis, at least

one solution to (4.18) and (5.44) lie in the region L ∈ (l1, · · · , ln) : li > 0 iff the

kis satisfy the following set of inequalities:

k1 > 0,

k2 >a21p1+

k1β2

p2

p2= (1 + γ1 + 1

β2)k1,

· · ·

ki >∑i−1

j=1aijpj+

ki−1

βipi

pi, ∀i < n

· · ·

kn > uM−λM

uM−(λM +dM )·∑n−1

j=1anjpj+

kn−1

βnpn

pn,

(4.31)

where, the coefficients pis are computed recursively using the formula

p1 = 1

pi = −∑i−1

j=1 ai−1jpj

(4.32)

Proof 2. =⇒ (Necessary condition):

Assume there exists a solution satisfying (4.18) and (5.44), which lies in the region

L ∈ (l1, · · · , ln) : li > 0. Denote it by

L′ ∈ (l′1, · · · , l′n) : l′i > 0 (4.33)

Let us first prove the following claim using mathematical induction.

Claim 3.

pil′i−1 > pi−1l

′i

pi > 0,(4.34)

for all i ≤ n.

Proof of Claim 3.

(i) Looking at the first row of AL′ < D, we have −k1l′1 + l′2 < 0. From the definition

Page 89: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

83

of p1 and p2 given by (5.48), and noting that a11 = −k1, we have p2l′1 > p1l

′2. Since

l′1, l′2 > 0 and p1 > 0, we obtain p2 > 0. Thus, the claim is true for j = 2.

(ii) Suppose the claim is true for all 2 ≤ j < i. Then, we have

l′j >pjl

′j+1

pj+1>pjpj+1l

′j+2

pj+1pj+2> · · · > pjl

′i−1

pi−1(4.35)

Now, looking at the i− 1 row of AL′ < D, and using aij > 0, ∀i > j, we get

i−1∑

j=1

ai−1jl′j + l′i < 0

⇒i−1∑

j=1

ai−1jpj l′i−1

pi−1+ l′i < 0

⇒ −pil′

i−1

pi−1+ l′i < 0, (using (5.48))

⇒ pil′i−1 > pi−1l

′i.

(4.36)

Since l′i−1, pi−1, l′i > 0, we get pi > 0. Thus, the claim is also true for j = i.

This procedure can be continued until i = n, as the R.H.S of the first n− 1 rows

of AL′ < D are all zeros. This completes the proof of the claim.

Next, it will be shows that the condition on gains, given by (5.47), follows from

(5.44) through the use of claim 3.

For each i < n, from the i-th row of AL′ < D we have

i−1∑

j=1

aijl′j +(

ki−1

βi− ki

)

l′i + l′i+1 < 0

⇒i−1∑

j=1

aijpj l′ipi

+(

ki−1

βi− ki

)

l′i + l′i+1 < 0, (l′j >pj l′ipi, ∀j < i using claim 3)

⇒i−1∑

j=1

aijpj l′ipi

+(

ki−1

βi− ki

)

l′i < 0

⇒[

i−1∑

j=1

aijpj

pi+(

ki−1

βi− ki

)

]

l′i < 0

⇒ ki >∑i−1

j=1aijpj+

ki−1

βipi

pi

(4.37)

From the last row of AL′ < D, we get

n−1∑

j=1

anjl′j +(

kn−1

βn− kn

)

l′n < −dM)

⇒n−1∑

j=1

anjpj l′npn

+(

kn−1

βn− ki

)

l′n < −dM

⇒[

kn −n−1∑

j=1

anjpj

pn− kn−1

βn

]

l′n > dM

(4.38)

Page 90: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

84

Applying (4.18), which is uM − λM ≥ knl′n, to the above inequality and eliminating

l′n:(

kn −n−1∑

j=1

anjpj

pn− kn−1

βn

)

[uM − λM ] ≥ kndM

⇒ kn >uM−λM

(uM−(λM +dM ))·∑n−1

j=1anjpj+

kn−1

βnpn

pn,

(4.39)

which completes the proof of necessary condition.

⇐= (Sufficient condition):

Suppose that kis satisfy inequalities (5.47). Then, for i ≤ n− 1,

ki >∑i−1

j=1aijpj+

ki−1

βipi

pi

⇒∑i−1

j=1aijpj+

ki−1

βipi−kipi

pi< 0

⇒ −pi+1

pi< 0, ( using definition of pi and aii = −ki + ki−1

βifrom(5.46))

⇒ pi+1

pi> 0.

(4.40)

Since p1 = 1 > 0, we have pi > 0, ∀i ≤ n.

Next, we are interested in finding the solution of the linear matrix equation AL′ =

D. Using the first n− 2 algebraic equations of AL′ = D, we can represent all the l′is

with i > 1 in terms of l′1:

l′1 = p1l′1

l′2 = p2l′1

· · ·l′n = pnl

′1.

(4.41)

Substituting them into the last equation of AL′ = D, we have:

n−1∑

j=1

anjpjl′1 + (kn−1

βn− kn)pnl

′1 = −dM

⇒ l′1 = −dMn−1∑

j=1

anjpj+(kn−1

βn−kn)pn

.(4.42)

From the last equation of (5.47), since 0 < (uM − (λM + dM)) < uM − λM , we have

kn >∑n−1

j=1anjpj+

kn−1

βnpn

pn

⇒n−1∑

j=1

anjpj + (kn−1

βn− kn)pn < 0.

(4.43)

Page 91: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

85

Thus, l′1 > 0. As a result, all the l′is are greater than zero. Multiplying both sides of

(5.47) by (uM − (λM +dM))pn, and rearranging the terms to factor out uM , we obtain

knl′n = knpnl

′1 =

−dMknpn

n−1∑

j=1

anjpj + (kn−1

βn− kn)pn

< uM − λM (4.44)

Thus, we have shown that the only solution to the linear matrix equation AL′ = D

lie in the region (l′1, · · · , l′n) : l′i > 0, knl′n < uM which is an open set. Since A is a

n by n matrix, the solution to inequality AL < D must exist and is a a sector region

with the acme being the point L′ which solves equation AL′ = D. Then, we can

always pick a solution of AL < D to be arbitrarily close to L′ such that L also lies in

the open region (l′1, · · · , l′n) : l′i > 0, knl′n < uM − λM. After choosing that L, we

have

(i) AL < D;

(ii) knln < uM − λM ;

(iii) all the lis are greater than zero.

Thus, there exists at least one solution satisfying (4.16)-(4.17), which lies in the

region (l′1, · · · , l′n) : l′i > 0. This proves the sufficiency condition of the theorem.

Remark 3. Condition (5.47) plays an important role in the selection of controller

parameters kis, and can be explained as follows. It can be easily shown that the

closed-loop transfer function from d(t) to z1 (which is also the tracking error y − yd)

inside the linear unsaturated region Ω11 is

[

z1(t)

d(t)

]

=1

sn + knsn−1 + knkn−1sn−2 + · · ·+n∏

j=2

kj · s +n∏

j=1

kj

. (4.45)

Due to the conditions imposed by (5.47), the closed-loop poles of (5.61) cannot be

assigned arbitrarily. However, as pi depends only on kjs for j < i, from (5.47) we can

easily work out a recursive way of choosing the controller parameters. Specifically, we

first choose k1, and then choose k2 large enough such that the first inequality of (5.47)

is satisfied. Continuing in this fashion, we can get a set of kis that are permissible.

In the next subsection, we clearly outline such a technique for choosing kis such that

Page 92: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

86

all closed-loop requirements are met, without violating the constraints imposed by

(5.47).

Controller Gain Selection: A Recursive Root-Locus Design

As mentioned above, the closed-loop poles of (5.61) cannot be arbitrarily assigned

because of the constraints (5.47) on the controller parameters ki. However, (5.47)

implies that ki can be chosen arbitrarily large, which makes it possible to satisfy the

closed-loop performance requirement. In the linear unsaturated region, the desired

closed-loop performance can be achieved by placing the poles sufficiently far to the

left of imaginary axis e.g., if the required steady-state error is δ, and the transient

response criteria dictates that the slowest closed-loop pole be to the left of p0, then

it is sufficient to place all the poles to the left of pcl = min−√

dM

δ,−p0. In the

following, we propose a recursive root locus design to meet the above requirement.

• Step 1: Select k1 > −pcl, then the root of the equation s+ k1 = 0 lie to the left

of s = pcl.

• Step 2: Let the virtual open-loop system be k2s+k1

s2 , then the virtual closed-loop

characteristic equation is s2 + k2s + k1k2 = 0. To determine k2, draw the root

locus of k2s+k1

s2 . This open-loop system has two poles at origin and one zero at

−k1, left to s = pcl. From the general guidelines for drawing a root-locus, there

exists a k2 large enough such that: (a) the first inequality of (5.47) is satisfied

and, (b) all the roots of s2 + k2s+ k1k2 = 0 lie to the left of s = pcl, on the real

axis.

.............

.............

.............

• Step n: Let the virtual open-loop system be kn

sn−1+kn−1sn−2+···+n−1∏

j=1

kj

sn , then the

virtual closed-loop characteristic equation is exactly the same as that of the

Page 93: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

87

actual system, i.e., knsn−1 + knkn−1s

n−2 + · · · +n∏

j=1

kj. To determine kn, draw

the root locus of kn

sn−1+kn−1sn−2+···+n−1∏

j=1

kj

sn . This open-loop system has n poles

at origin and n− 1 zeros to the left of s = pcl. As the difference in the number

of poles and zeros is one, there always exists an asymptote along the negative

real axis. Thus, for sufficiently large gain kn, there is a branch of the root-locus

on the negative real axis. This implies (a) the last two inequalities of (5.47) are

satisfied and (b) all the closed-loop poles lie to the left of s = pcl on the real

axis.

Thus, we can choose the controller gains such that the desired closed-loop perfor-

mance is achieved, as well as the conditions imposed for the existence of a feasible

control law given by (5.47) are also satisfied simultaneously.

4.4 Simulation Example: Third Order Integrator Chain

A 3rd order chain of integrators is used to demonstrate the effectiveness of the

proposed scheme. The proposed scheme is compared with a transformation based

controller design technique [50] to underline the superior achievable performance. In

the first set of simulations, it is assumed that there is no disturbance present in the

system. The results show that the achievable convergence rate is better than that

obtained in [50]. Simulation studies conducted in presence of disturbances - second

and third set of simulations, demonstrate the robustness of the proposed scheme

against input disturbance.

Simulation 1: The goal of this simulation study was to investigate the convergence

rate of the proposed scheme against a transformation based approach for large initial

Page 94: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

88

conditions in absence of input disturbance. The third order chain of integrator given

by

x1

x2

x3

=

0 1 0

0 0 1

0 0 0

x1

x2

x3

+

0

0

1

u (4.46)

was used in [50] and thus, provides a level platform to compare the two approaches.

The initial conditions are same as used in the cited paper x(0) = [2,−2, 3]. We use

the control law proposed in Theorem 3 of [50] for comparison. The controller param-

eters for the proposed scheme are: [k1, k2, k3] = [0.2, 0.9, 20], [l1, l2, l3] = [4, 0.8, 0.05],

[β1, β2, β3] = [0.909, 0.909, 1], [γ1, γ2, γ3] = [0.01, 0.01, 0]. The parameter selection

scheme proposed in the previous section was used as a starting point, and then it

was tuned to improve the performance. As seen from fig. (5.2), we achieve slightly

faster convergence with the proposed scheme. This shows that the performance of

the proposed technique in terms of achievable convergence rate is at least as good as

that of a transformation based approach. However, the true strength of the proposed

controller and its robustness against disturbances is demonstrated in the next set of

simulations when large disturbances are considered.

Simulation 2: The purpose of this simulation study was to compare the conver-

gence rate and steady-state error for a third order integrator chain in presence of

input disturbance. The input disturbance and initial conditions were chosen to be

d(t) = 0.1 sin(πt2) and x(0) = [0.2,−0.2, 0.3] respectively. The initial conditions were

chosen to be smaller than the previous case in order to highlight the effect of distur-

bance on steady-state error. Controller parameters are same as used in the previous

simulation. The disturbance attenuation capability of the proposed scheme over a

coordinate transformation based approach is evident from fig. (5.3) and can be ex-

plained as follows. First, note that there is no transparent and easy way of tuning the

controller proposed in [50] to attenuate the effect of disturbances on the states. In

fact, as the stability analysis in the aforementioned paper is performed in the trans-

formed coordinates, there is no convenient way of studying the effect of disturbance

Page 95: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

89

on the original states. Second, in our design it is easy to understand and account

for the effect of disturbances. The closed loop characteristic equation of the system

in the linear unsaturated region is given by s3 + 20s2 + 18s + 3.6 = 0. Thus, using

the final value theorem, we know the disturbance should be attenuated by a factor

of 3.6, which matches the simulation result. In fact, the effect of disturbance can

be attenuated to any desired extent by choosing high controller gains, as outlined in

the previous section. Also, note that the convergence of states depend on two fac-

tors: first, how fast the states reach the unsaturated region, and second, how fast the

states go to zero once within the unsaturated region (or a residual ball in presence

of disturbances). Thus, if we choose large li, we can make the convergence to the

unsaturated region faster, but then we cannot choose very high controller gains once

within the unsaturated region, as inequalities (4.16-4.18) must be satisfied. In fact, in

the context of bounded control, such a trade off is expected. Due to the transparency

of these trade-offs in the proposed scheme and the available freedom in choosing the

controller parameters, a better controller can be designed when the initial conditions

are known e.g., when the initial conditions are not too large, we can select a relatively

small li and choose high gains to attenuate the effect of disturbances and vice-versa.

Simulation 3: In this case, we present an example where the conventional ap-

proach fails to yield a stable controller. In the problem formulation, we had al-

ready shown that when uM = 10, the transformation makes it impossible to de-

sign a stable controller with dM = 4. In contrast, using the proposed technique,

we can not only design a stable controller, but also achieve desired disturbance

attenuation. For this set of simulations, the desired trajectory is a filtered a 0.5-

step command and the objective is to track it with a steady-state error less than

0.01 in presence of large input disturbance d(t) = 4 sin(πt2). Initial conditions are

set at x(0) = [0, 0, 0.2]. Controller parameters used for the tracking problem are:

[k1, k2, k3] = [1, 8, 50], [l1, l2, l3] = [0.12, 0.1, 0.2], [β1, β2, β3] = [0.909, 0.909, 1], and

[γ1, γ2, γ3] = [0.01, 0.01, 0]. Note that the closed loop characteristic equation of the

system in the linear unsaturated region is give by s3 + 50s2 + 400s+ 400 = 0, which

Page 96: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

90

ensures that ess ≤ 0.01. As can be seen from fig. (4.4), the steady-state error is less

than 0.01 i.e., the desired performance is achieved. This example clearly illustrates

the effectiveness of the proposed scheme in reducing the design conservativeness over

a transformation based approach.

Page 97: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

91

0 10 20 30 40 500

2

4

6

8

10

12

time (sec)

norm

(x(t)

)

Proposed designMarchand and Hably

0 10 20 30 40 50−1

−0.5

0

0.5

time (sec)

u

Proposed designMarchand and Hably

0 10 20 30 40 50−2

0

2

4

6

8

10

12

time (sec)

x 1

Proposed designMarchand and Hably

0 10 20 30 40 50−2

−1

0

1

2

3

time (sec)

x 2

Proposed designMarchand and Hably

0 10 20 30 40 50−1

−0.5

0

0.5

1

1.5

2

2.5

3

time (sec)

x 3

Proposed designMarchand and Hably

Figure 4.2. Comparative results for stabilization in absence of distur-

Page 98: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

92

0 10 20 30 40 500

0.1

0.2

0.3

0.4

0.5

time (sec)

norm

(x(t)

)

Proposed designMarchand and Hably

0 10 20 30 40 50−1

−0.8

−0.6

−0.4

−0.2

0

0.2

0.4

time (sec)

u

Proposed designMarchand and Hably

0 10 20 30 40 50−0.05

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

time (sec)

x 1

Proposed designMarchand and Hably

0 10 20 30 40 50−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

time (sec)

x 2

Proposed designMarchand and Hably

0 10 20 30 40 50−0.1

−0.05

0

0.05

0.1

0.15

0.2

0.25

0.3

time (sec)

x 3

Proposed designMarchand and Hably

Figure 4.3. Comparative results for stabilization in presence of distur-

Page 99: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

93

0 10 20 30 40 50 60 70 800

0.1

0.2

0.3

0.4

0.5

0.6

0.7

time (sec)

x 1 and

x1d

x1

x1d

0 10 20 30 40 50 60 70 80−0.05

0

0.05

0.1

0.15

time (sec)

e(t)

0 10 20 30 40 50 60 70 80−12.0

−8.0

−4.0

0.0

4.0

8.0

time (sec)

u

0 10 20 30 40 50 60 70 80−0.1

0

0.1

0.2

0.3

0.4

0.5

0.6

time (sec)

x 2

0 10 20 30 40 50 60 70 80−1

−0.5

0

0.5

1

time (sec)

x 3

Figure 4.4. Tracking in presence of large disturbance

Page 100: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

94

4.5 Conclusion

The main contribution of the present work lies in proposing a conceptually dif-

ferent approach to solve the problem of global stabilization for a chain of integrators

in presence of input disturbance with desired performance in the unsaturated region.

Based on Teel’s work, many modifications have been proposed in the literature to

improve the performance of the controller. In our analysis, it was clearly shown that

all such schemes exhibit poor robustness properties with respect to input disturbance

and leads to conservative design. In fact, a quantitative analysis revealed that even

when the magnitude of the disturbance is less than that of the available control input,

the coordinate transformation can render the design of a stabilizing controller impos-

sible. These limitations cannot be overcome by any modification based on Teel’s

work, as coordinate transformation is an essential step in all such designs. In order

to remove these limitations, we take a fundamentally different viewpoint and propose

a scheme which does not rely on coordinate transformation, and is based on back-

stepping design. The resulting controller is easy to implement and tune, as we only

deal with the original coordinates. Furthermore, the proposed design can also cap-

ture the trade-offs which are inherent to any input-saturated control design problem

- choosing high controller gains versus enlarging the unsaturated region of linear con-

troller operation. Comparative studies have been performed on a third order chain

of integrator to show the superior performance of the proposed technique. The first

set of studies, performed in absence of disturbances, revealed that the convergence

rate of the proposed scheme is at least as good as that proposed by Marchand and

Hably, which is based on Teel’s work. In presence of disturbances, however, signif-

icant differences could be seen in terms of disturbance attenuation and convergence

of the states. A tracking problem with large disturbance, where a stable controller

could not be designed due to coordinate transformation, was also solved to show the

effectiveness of the proposed scheme.

Page 101: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

95

5. SATURATED ADAPTIVE ROBUST ACTUATOR FAULT-TOLERANT

CONTROL FOR FEEDBACK LINEARIZABLE SYSTEMS

5.1 Introduction

In this chapter, an adaptive robust fault-tolerant scheme is proposed to deal with

unknown actuator faults in presence of disturbances and input magnitude constraints.

Among various approaches, adaptive control based fault-tolerant schemes have found

popularity among researchers, as they have the ability to learn the change in system

parameters due to actuator faults. However, input magnitude constraint - one of the

most important factors which can limit the performance of any control system, has

largely been overlooked in the literature.

The harmful effect of actuator faults on the system response increases manifold in

presence of actuator saturation. A control system which does not explicitly take into

account input saturation could generate large control signals to suppress the tran-

sients. However, as all actuators have limited authority, this could lead to actuator

saturation. Furthermore, direct adaptive scheme, which are preferred over indirect

scheme due to their lower dynamic order, may generate unreliable parameter esti-

mate once the actuators saturate. This could further degrade the performance of an

actuator fault-tolerant scheme.

In this chapter, the backstepping and saturation function based approach devel-

oped in the last chapter is combined with a least-squares estimator to develop an

indirect saturated adaptive robust fault-tolerant scheme. x-swapping lemma is used

to design the least-squares estimator. The bounded feedback control generated using

saturation functions compromise the performance when error-variables are far away

from zero, and lays more emphasis on bringing the error-variables in a region where

the controller is unsaturated. A carefully chosen set of controller gain ensures that

Page 102: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

96

desired closed-loop performance is achieved, once within the unsaturated region. Ad-

ditionally, the indirect scheme ensures that parameter estimation is unharmed despite

actuator saturation and the proposed controller is shown to ISS with respect to para-

metric uncertainties and disturbances in the linear region. This results in improved

tracking performance, and guarantees asymptotic stability in presence of parametric

uncertainties only.

Controller design with actuator constraints is a challenging problem in itself. It

is worth mentioning that in the present work, we focus our attention to only those

class of nonlinear systems which are feedback linearizable with matched uncertain-

ties. Feedback linearization [52] based control techniques have become a useful tool

in the repertoire for flight control systems design e.g., [53], [54] and [55]. One of the

main reasons for the popularity of this technique is that it allows linear control design

techniques to be applied to nonlinear systems, and relieves the burden of nonlinear

systems stability analysis. In the present work, we add another powerful tool to this

collection in the form of adaptive robust actuator fault-tolerant controller, which ex-

plicitly considers the input magnitude constraint. Note that we do not claim that the

proposed scheme can accommodate faults of any magnitude. Furthermore, for general

nonlinear systems, it is not possible to design a globally stable bounded controller.

Thus, the initial conditions for the system trajectories and the extent of parametric

uncertainties and disturbance will be assumed to be bounded by known constants.

To summarize, three main contributions of the proposed work are - (i) design of a

control law that accommodates unknown actuator faults with desired closed-loop per-

formance in the unsaturated region, (ii) explicit consideration of actuator limits in

the controller design, which allows the controller to pull out of saturation phase after

faults and, (iii) uninterrupted adaptation regardless of saturation.

The performance of the fault-tolerant controller proposed in this chapter is com-

pared with the one developed in the second chapter, using the nonlinear hypersonic

aircraft model. The results clearly indicate the harmful effect of saturation on the

closed-loop stability. The controller which does not consider input constraints in

Page 103: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

97

the design leads to unbounded signals, where as the proposed scheme goes into the

saturation mode temporarily, but returns to the unsaturated region quickly.

5.2 Problem Formulation

In this work, we consider systems which can be written as

x = f(x) + g(x)[b1u1 + · · ·+ bquq] + w(x)d(t)

y = h(x), |uj(t)| ≤ uM , j = 1, ..., q (5.1)

where w(x) is the distribution matrix for the disturbance. It will be assumed that

f(x), g(x), w(x) and h(x) are smooth functions. The plant parameters bi, which

are assumed to be unknown, belong to a known region i.e., bi ∈ [bi,min, bi,max]. In the

present analysis, we focus our attention to the class of systems for which the following

assumption holds.

A1: System (5.1) has a well-defined relative degree m with respect to the output

y = h(x) such that there exists a diffeomorphism T (x) which transforms the x to

[ζ, η]′ coordinates as follows

ζ

η

=

ζ1

ζ2...

ζm

η1

...

ηn−m

= T (x) =

h(x)

Lfh(x)...

Lm−1f h(x)

T1(x)...

Tn−m(x)

(5.2)

Page 104: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

98

Thus, the dynamics can be rewritten as

ζ1 = ζ2

ζ2 = ζ3...

ζm = Lmf h(x) + LgL

m−1f h(x)[b1u1 + · · ·+ bquq] + LwL

m−1f h(x)d(t)

η = fη(ζ, η, d(t)) (5.3)

where it is assumed that the disturbance distribution matrix is such that LwLjf = 0

for all j = 1, ..., m− 2 and LwLm−1f 6= 0, and Lm

f h(x) 6= 0, for all x ∈ Rn.

In the present work, we will also make the following assumption about the zero-

dynamics.

A2: The η-dynamics is input to state stable (ISS) with respect to ζ, η, d(t).

Assumption A2 guarantees the boundedness of all closed-loop signals when a stable

controller can be designed for the ζ-dynamics.

In this work, we will consider actuator faults which can be modeled as

uj(t) =

uj, ∀t ≥ Tf ,

if jth actuator gets stuck at Tf

ηjju∗j(t), ∀t ≥ Tf ,

if jth actuator loses efficiency at Tf

(5.4)

where u∗j(t) represents the control command to the jth actuator, uj is an unknown

constant value at which the actuator gets stuck, Tf is the unknown instant of failure

and ηjj represents actuator loss in efficiency with ηjj ∈ [(ηjj)min, 1], (ηjj)min ≥ 0. It

will be assumed that uj belongs to a known interval i.e., uj ∈ [uj,min, uj,max].

For the system described by (5.1), subjected to unknown actuator faults (5.4)

and bounded disturbances, the objective is to design a control law such that the

output tracking error converges to a sufficiently small neighborhood of zero where the

controller is unsaturated, and has desired closed-loop performance and disturbance

attenuation properties within the unsaturated region. It is also desired that the

tracking error asymptotically converges to zero in absence of disturbances.

Page 105: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

99

5.3 Adaptive Robust Actuator Fault-Tolerant Control

Denote fζ(x) = Lmf h(x), gζ(x) = LgL

m−1f h(x) and wζ = LwL

m−1f h(x). In the

present work, we will assume that control commands to all the actuators are the

same i.e., u∗1 = · · · = u∗q = u0. With this control input, and fault model described by

(5.4), the healthy and faulty actuators can be parameterized in the following way

uj(t) = ηjj(1 − σjj)u0(t) + σjj uj

⇒q∑

j=1

bjuj(t) =

(

q∑

j=1

bjηjj(1 − σjj)

)

u0(t) +

(

q∑

j=1

bjσjjuj

)

⇒q∑

j=1

bjuj(t) = κu0(t) + µ (5.5)

where

σjj =

0 before jth actuator gets stuck

1 after jth actuator gets stuck

ηjj =

1 before jth actuator loses efficiency

[(ηjj)min, 1] after jth actuator loses efficiency

κ =

q∑

j=1

bjηjj(1 − σjj), µ =

q∑

j=1

bjσjj uj

We will make the following practical assumption regarding the extent of uncertainties

present in the system.

A3: The unknown parameters θ , [κ, µ]T and disturbance satisfy,

θ ∈ Ωθ , θ : θmin ≤ θ ≤ θmax (5.6)

wζ(x)d(t) ∈ Ωd , wζ(x)d(t) : |wζ(x)d(t)| ≤ dM (5.7)

Now, that we have established a parametric fault-model, we are ready to present

the bounded control laws to be used in the backstepping based design. We will use

Page 106: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

100

the coordinate transformation xi = ζi − y(i−1)d to simplify the analysis. The designed

virtual control law αi and the corresponding error zi are given by

zi = xi − αi−1(zi−1),

αi = −σi(zi),

˙xi = zi+1 + αi, i = 1, ..., n− 1 (5.8)

where σi(zi) are saturation functions (see figure (5.1)). The saturation functions used

in this work play a significant role in guaranteeing the overall system stability. In fact,

it is only through a careful and rigorous analysis of the effects of saturation function

parameters on system dynamics that a set of conditions will be derived, which will

ensure all desired objectives can be achieved. The saturation functions are defined as

(a) ziσi(zi) ≥ 0, ∀zi (5.9)

(b) σi(zi) = kizi, ∀|zi|≤li,

σi(zi) = Mi, ∀|zi|≥Li (5.10)

(c) |σi(zi)| ≤Mi, ∀zi (5.11)

(d)∂σi

∂zi

≤ ki, ∀zi (5.12)

Also, li = βiLi with βi ≤ 1, and Mi = kili(1 + γi) with γi > 0. The interval for

zi is divided into three different regions - Ωi1 = zi : |zi| ≤ li, Ωi

2 = zi : |zi| ≤Li and Ωi

3 = zi : |zi| > Li. Note that the nonlinear transition region of the

saturation function (Ωi2\Ωi

1) needs to be at least second order differentiable, as the

backstepping design involves taking derivatives of σi. σm, however, need not have a

smooth transition region, as this appears in the last step. The lm and Lm parameters

used in the definition of σm depend on the extent of uncertainties, and will be defined

later.

Page 107: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

101

Figure 5.1. Saturation function

Page 108: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

102

Now, substituting (5.8) in (5.3), the error dynamics can be written as

z1 = z2 − σ1(z1)

...

zi = zi+1 − σi(zi) +

i−1∑

j=1

[

j∏

r=1

∂σi−r

∂zi−r

]

(zi−j+1 − σi−j(zi−j))

(5.13)

...

zm = fζ(x) + κgζ(x)u0 + µgζ(x) + wζ(x)d(t) − y(m)d

+m−1∑

j=1

[

j∏

r=1

∂σm−r

∂zm−r

]

(zm−j+1 − σm−j(zm−j))

(5.14)

Remark 1. Note that the form of the virtual control laws designed is signifi-

cantly different, and at the same time much simpler than the typical “cancellation”

backstepping design. In cancellation backstepping design, the detrimental effect of

the virtual control law of step i − 1 i.e., αi−1 is completely canceled at step i, by

incorporating the appropriate terms in the design of αi. This results in a simple

linear dynamics for the error variables, such that the closed-loop poles can be placed

at any desired location. But, this simplicity comes at the expense of a complicated

control law, with exponentially growing terms as the order of the system increases.

In the context of bounded control, it is extremely difficult to show the boundedness

of such a control law. For chain of integrators, as we shall see, the error dynamics can

be stabilized with a much simple control law σi(zi), and without canceling the terms

resulting from the derivative of αi−1. Although linear, the resulting error-dynamics

without cancellation leads to a set of constraints that the controller parameters must

satisfy for global stabilization. This makes the parameter selection process slightly

more complicated than the cancellation backstepping design. Next, we present the

adaptation mechanism used in the present work.

Page 109: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

103

5.3.1 Parameter Estimation

We will use x-swapping lemma (Ch.6, [31]) to implement least-squares estimation

scheme, and then use discontinuous projection to ensure that the parameters stay

within a known region in presence of disturbances. Note that as all unknown param-

eters appear in the mth channel in the ζ coordinates, we need only ζm dynamics for

estimation purposes.

The ζm-dynamics can rewritten as

ζm = fζ(x) + κgζ(x)u0 + µgζ(x) + wζ(x)d(t)

= fζ(x) + φ(x, u)Tθ + wζ(x)d(t) (5.15)

where θ , [κ, µ]T and φ(x, u)T = [gζ(x)u0, gζ(x)]. Following the standard steps of

x-swapping, we define the following filters

Ω0 = A(Ω0 + ζm) − fζ(x) (5.16)

ΩT = AΩT + φ(x, u)T (5.17)

Define the prediction error as ǫ = ζm + Ω0 − ΩT θ, which is calculable. It is shown

in [31] that ǫ can be rewritten as

ǫ = ΩT θ + ǫ (5.18)

where ǫ is governed by ˙ǫ = Aǫ, which exponentially converges to zero. Thus, we have

a static model (5.18), that is linearly parameterized in terms of θ with an additional

term ǫ which exponentially decays to zero. With this static model, various estimation

algorithms can be used to estimate the unknown parameters. In the following, we

present the least-squares estimation scheme which will be used in the present work.

˙θ = Γτ

τ =Ωǫ

1 + νtrΩT ΓΩ

Γ = −ΓΩΩT

1 + νtrΩT ΓΩΓ, Γ(0) = Γ(0)T > 0, ν ≥ 0 (5.19)

Page 110: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

104

In order to guarantee certain desired properties, we will use the following discon-

tinuous projection algorithm.

˙θ = Projθ(Γτ) (5.20)

Projθi=

0 if θi = θi,max and •i > 0

0 if θi = θi,min and •i < 0

•i otherwise

(5.21)

This guarantees that the parameters do not drift away and stay within known bounded

region even in presence of disturbances. Additionally, we will use the following rate-

limiting projection scheme

satθM(ξ) = s0ξ, s0 =

1, ‖ξ‖ ≤ θM

θM

‖ξ‖‖ξ‖ > θM

(5.22)

where θM is a pre-set rate limit. Thus, the final parameter update law used takes the

following form

˙θ = satθM

(Proj(Γτ)) , θ(0) ∈ Ωθ (5.23)

Using the properties of the projection operator in Lemma E.1 in [31], and noting

that s0 is a positive scalar, it is to verify that the following desirable properties hold

P1 θ ∈ Ωθ = θ : θmin ≤ θ ≤ θmax (5.24)

P2 θT (Γ−1Projθ(Γτ) − τ) ≤ 0, ∀τ (5.25)

P3 ‖ ˙θ‖ ≤ θM , ∀t (5.26)

Property P1 guarantees that the estimates stay within a known bounded region, and

P3 guarantees that the parameter update rate is uniformly bounded.

5.3.2 Controller design

In this section, we present the proposed adaptive robust fault-tolerant controller

and prove the overall stability of the system using the following steps:

Page 111: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

105

1. In the first step, it is shown that for any set of initial conditions zi(0), all

error-variables can be driven to an invariant region where the controller is un-

saturated, as long as a set of controller parameters exist which satisfy certain

inequalities.

2. Next, sufficient and necessary conditions for the existence of the controller pa-

rameters is proposed and proved.

3. In this step, a recursive root-locus design is proposed such that once the con-

troller is unsaturated, desired closed-loop performance e.g., disturbance atten-

uation and desired transients can be achieved.

4. In the last step, asymptotic convergence of the adaptive system is proved in

absence of disturbances within the unsaturated region.

Define ua = −fζ(x) − µgζ(x). Control law to be used

u0 = σm

[

1

gζ(x)κ(σa(ua) + y

(m)d − kmzm)

]

(5.27)

where

σa(ua) =

ua, for |ua|≤Ma

sign(ua)Ma, for |ua| > Ma

and

σm(u0) =

u0, for |u0|≤uM

sign(u0)uM , for |u0| > uM

Note that as system (5.3) may not necessarily be stable, the model-compensation

component ua can easily become unbounded. In the context of bounded control,

some assumptions must be made on the growth rate of the nonlinearities to make the

stabilization/tracking problem feasible.

A4: In the present work, it will be assumed that the nonlinearities and the esti-

mated parameters are such that

(i) when the model-compensation is saturated i.e., σa(ua) = Ma, the difference can

Page 112: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

106

be bounded above by a known constant i.e., |σa(ua) − ua| ≤ uaM

(ii) the discrepancy due to parameter estimation mismatch is bounded by a known

constant i.e., |φT (x, u)θ| ≤ hM .

(iii) gζ(x) is such that gζ,min ≤ |gζ(x)|≤gζ,max.

Consider the ζ−dynamics of system (5.3) along with assumptions (A1-A4). The

following theorem states that in spite of unknown actuator faults (5.5), the error

dynamics can be driven to small neighborhood around zero, where the controller is

unsaturated and desired closed-loop performance can be recovered.

Theorem 1. Consider the error-dynamics represented by (5.14). With the control

law given by (5.27), and the chosen parameter update law (5.23), if a set of controller

parameters can be chosen such that

kili > li+1 + ki−1Ni, i = 1, ..., m− 1 (5.28)

kmlm > km−1Nm + hM + dM + uaM , (5.29)

kmlm ≤ κmingζ,minuM − (Ma + λM) (5.30)

where

Ni = Li +Mi +

i−2∑

j=1

[

(

j∏

r=1

ki−1−r)(Li−j −Mi−1−j)

]

and k0 = 0 then, the error variables zi reach a region where the controller is unsat-

urated in a finite time (i.e., z ∈⋂m

j=1 Ωj1), for any set of initial conditions and any

fault pattern.

Proof. Consider the following claims

Claim 1. u0 is unsaturated for

|zm| ≤ lm ,gζ,minκminuM − (Ma + λM)

km

(5.31)

and saturated for

|zm| ≥ Lm ,gζ,maxκmaxuM + (Ma + λM)

km

(5.32)

Page 113: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

107

Claim 2. For any initial condition zm(0),

(a) if zm(0) ∈ Ωm3 i.e., |zm(0)| > Lm, then it reaches Ωm

2 in finite time.

(b) if zm(0) ∈ Ωm2 \Ωm

1 i.e., lm < |zm(0)|≤Lm, then it reaches Ωm1 in finite time.

(c) if zm(0) ∈ Ωm1 , then zm(t) ∈ Ωm

1 , ∀t > 0.

Claim 3. Assume zi+1≤li+1, then

(a) if |zi(t0)| > Li, then it reaches Ωi2 in a finite time

(b) if zi(t0) ∈ Ωi2\Ωi

1, then it reaches Ωi1 in a finite time

(c) if zi(t0) ∈ Ωi1, then zi(t) ∈ Ωi

1 ∀ time t > t0

Assume all the claims are true. The first claim establishes the bounds for zm such

that the control law given by (5.27) is unsaturated/saturated. Then, from the second

claim, we have zm≤lm after a finite time, and it can never leave this set. Theorem 1

then follows from the recursive application of claim 2 to the intermediate states zj ,

j = m− 1, ..., 1.

Proof of Claim 1. Assume w.l.o.g that zm > 0. Then, for zm≤lm, from (5.31)

we have the following inequalities

kmzm ≤ kmlm ≤ gζ,minκminuM − (Ma + λM) ≤ gζ(x)κuM − (Ma + λM)

⇒ kmzm +Ma + λM

gζ(x)κ≤ uM

⇒ kmzm − σa(ua) + y(m)d

gζ(x)κ≤ uM

⇒ u0 ≤ uM

Thus, the controller is unsaturated. In order to show that the controller is saturated

for zm > Lm, consider the following inequalities

kmzm ≥ kmLm ≥ gζ,maxκmaxuM + (Ma + λM)

⇒ kmzm − (Ma + λM) ≥ uMgζ(x)κ

⇒ kmzm − σa(ua) + y(m)d

gζ(x)κ≥ uM

Thus, the control law u = u0 is saturated.

Page 114: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

108

Proof of Claim 2, part (a). Assume w.l.o.g zm(0) > 0. Then, we have the

following inequalities for zm ∈ Ωm3

˙xm = fζ(x) + κgζ(x)u0 + µgζ(x)

−y(m)d + φT (x)θ + wζ(x)d(t)

= −ua − y(m)d + φT (x)θ + wζ(x)d(t) + κgζ(x)u0

≤ Ma + λM + hM + dM − κmingζ,min(x)uM

⇒ xm(t) ≤ xm(0) − (κmingζ,min(x)uM

−(Ma + λM + hM + dM))t

⇒ zm(t) + αm−1(t) ≤ zm(0) + αm−1(0) − (κmingζ,min(x)uM

−(Ma + λM + hM + dM))t

Using |αm−1(t)| ≤ Mm−1 and substituting zm(t) = Lm, an upperbound for the time

taken to reach Ωm1 can be found as

tm32 ≤zm(0) − Lm + 2Mm−1

(κmingη,min(x)uM − (Ma + λM + hM + dM))(5.33)

From (5.29) and (5.30) we also have

κmingζ,minuM − (Ma + λM) ≥ km−1Nm + hM + dM + uaM

⇒ κmingζ,minuM ≥ hM + dM +Ma + λM .

Thus, zm reaches Ωm2 in a finite time.

Proof of Claim 2, part (b). Note that when zm ∈ Ωi2\Ωi

1, the control law

can be saturated or unsaturated depending on the relative magnitudes of the various

uncertainties (e.g., φT θ, wζ(x)d(t)), the model-compensation (ua) and the robust

component (−kmzm). However, the error-variables can be made to converge in this

region, as long as conditions (5.29) and (5.30) are satisfied. Both the cases i.e.,

Page 115: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

109

|u0| > uM and |u0|≤uM , will be considered separately. Assume w.l.o.g zm > 0. The

derivative of zm can be written as

zm = fζ(x) + µgζ(x) − ymd + κgζ(x)u0

+ φT (x)θ + wζ(x)d(t) +

m−1∑

j=1

[

j∏

r=1

∂σm−r

∂zm−r

]

(zm−j+1 − σm−j(zm−j))

≤ −ua − y(m)d + κgζ(x)u0 + km−1Nm + hM + dM (5.34)

Case A. u0 is not saturated.

In this case, substituting (5.27) in (5.34) and using condition (5.29), we obtain

zm ≤ σa(ua) − ua − kmzm + hM + dM + km−1Nm

≤ −kmlm + km−1Nm + hM + dM + uaM

< 0

Thus, zm is negative in this case.

Case B. u0 is saturated.

Substituting u0 = uM , we get the following inequalities from (5.34)

zm ≤ −ua + κgζ(x)u0 + km−1Nm + hM + dM + λM

≤ Ma − κmingζ,minuM + km−1Nm + hM + dM + λM

For zm to be negative, we must have

uM >km−1Nm +Ma + hM + dM + λM

κmingζ,min(5.35)

Note that we have following inequalities from (5.29) and (5.30)

κmingζ,minuM − (Ma + λM) > kmlm > km−1Nm + hM + dM + uaM

⇒ uM >km−1Nm + hM + dM + λM +Ma

κmingζ,min(5.36)

Thus, condition (5.35) is automatically satisfied. Thus, zm is negative for zm ∈ Ωi2\Ωi

1,

irrespective of whether the control input is saturated or not.

Page 116: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

110

Proof of Claim 2, part (c). For Ωm1 to be invariant, zmzm should be negative

at zm = lm. Note that from (5.29), there exists a δm such that

km(lm − δm) = km−1Nm + hM + dM + uaM (5.37)

Now, substituting zm = lm in zm, and using (5.37), we get

zm = −kmzm + σa(ua) − ua + φT (x)θ + wζ(x)d(t)

+m−1∑

j=1

[

j∏

r=1

∂σm−r

∂zm−r

]

(zm−j+1 − σm−j(zm−j))

⇒ zmzm ≤ |zm|(−kmlm + km−1Nm + uaM + hM + dM)

≤ −km|zm|δm < 0 (5.38)

Thus, any trajectory starting in Ωm1 will remain in this set.

Proof of Claim 3, part (a). When zi ∈ Ωi3 i.e., |zi| > Li, we have αi = −Mi

(w.l.o.g assume zi > 0) from the definition of σi. Using the assumption of claim 2

i.e., |zi+1|≤li+1 and substituting αi = −Mi in (5.8), we obtain

˙xi = zi+1 −Mi

≤ li+1 −Mi = −(Mi − li+1)

⇒ xi(t) ≤ xi(t0) − (Mi − li+1)(t− t0)

⇒ (Mi − li+1)(t− t0) ≤ xi(t0) − xi(t) + αi−1(t0) − αi−1(t)

⇒ ti32 ≤ xi(t0) − Li + 2Mi−1

Mi − li+1(5.39)

where in the last step we substituted zi(t) = Li and ti32 , t − t0, the time taken to

reach Ωi2 from Ωi

3. This proves the first part of the claim.

Proof of Claim 3, part (b). To prove the next two parts, we proceed as done

in the proof of first claim. From (5.28), there exists a δi such that

ki(li − δi) = li+1 + ki−1Ni (5.40)

Page 117: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

111

Without loss of generality, assume zi(t0) > 0. Then, from the zi-dynamics (5.13), we

have

zi = zi+1 − σi(zi) +i−1∑

j=1

[

j∏

r=1

∂σi−r

∂zi−r

]

(zi−j+1 − σi−j(zi−j))

≤ li+1 − kili +i−1∑

j=1

[

(

j∏

r=1

ki−r)(Li−j+1 +Mi−j)

]

≤ li+1 − kili + ki−1Ni (5.41)

= −kiδi

⇒ ti21 ≤ Li − likiδi

(5.42)

where we have used the fact that Li − li≥|zi(t) − li|, ∀ zi(t) ∈ Ωi2\Ωi

1. Thus, there

exists a finite time-interval ti21, within which any trajectory starting in Ωi2\Ωi

1 reaches

Ωi1. This completes the proof of second part of the claim.

Proof of Claim 3, part (c) In order to show that any trajectory starting in

Ωi1(t0) stays there ∀ t > t0, we need to show that zizi < 0 at zi = ±li. For zi = li, we

know σi(zi) = kili and using the assumption |zi+1|≤li+1, from (5.13) the zi-dynamics

can be written as

zizi = zi

(

zi+1 − σi(zi) +

i−1∑

j=1

[

j∏

r=1

∂σi−r

∂zi−r

]

(zi−j+1 − σi−j(zi−j))

)

≤ |zi|(

li+1 − kili +

i−1∑

j=1

[

(

j∏

r=1

ki−r)(Li−j+1 +Mi−j)

])

≤ −ki|zi|δi (5.43)

This shows that Ωi1 is positively invariant, and also completes the proof of the theorem

1.

5.3.3 Controller Parameter Selection

There are two important questions which need to be answered next. First, when

can the existence of a solution to inequalities (5.28)-(5.30) be guaranteed. Note that

even when such a solution exists, it only guarantees the convergence of all intermediate

Page 118: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

112

error variables to the linear unsaturated region (an invariant set in the neighborhood

of zero). The system response within the linear unsaturated region is governed by the

gains ki, and leads to the second important question - how to select the gains such

that the desired closed-loop performance can be achieved, without violating (5.28)-

(5.30). It is natural to first select a set of gains ki such that desired performance

can be achieved, and then seek a set of lis such that inequalities (5.28)-(5.30) can

be satisfied. However, such an approach may not work as after fixing the gains ki,

solving the inequalities (5.28)-(5.30) may result in negative lis, which will make the

control law unrealizable. Hence, in the following section, we first state and prove

the main result regarding the necessary and sufficient condition for the existence of a

solution to the inequalities. Then, a systematic way of choosing the controller gains in

accordance with this theorem is proposed, such that desired closed-loop performance

can be achieved.

Necessary and Sufficient Conditions for the Existence of Controller Pa-

rameters

After a series of derivations, (5.28)-(5.29) can be rewritten in a matrix form

AL < D, (5.44)

where L = [l1, l2, · · · , lm−1, lm]T , D = [0, 0, · · · , 0,−(hM + dM + uaM)]T . And A is a

function of ki given by

A =

−k1 1 0 · · · 0

k21(1 + γ1)

(

k1

β2− k2

)

1 · · · 0

· · · · · · · · · · · · · · ·am1 am2 am3 · · ·

(

km−1

βn− km

)

, (5.45)

Page 119: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

113

where

aij = ki−1kj

i−j−1∏

r=1

ki−r−1(1 + γj) + ki−1

i−j∏

r=1

ki−r−11βj, ∀i > j.

aij = −ki + ki−1

βi, ∀i = j

aij = 1, ∀j = i+ 1

aij = 0, ∀j > i+ 1

(5.46)

For any γi > 0 and 0 < βi ≤ 1, if we fix a set of positive kis, then the control

law is feasible if and only if the there exist l1, l2,..., lm > 0 such that (inequality3)

and (5.44) are satisfied. In other words, at least one solution to (inequality3) and

(5.44) should lie in the region (l1, · · · , lm) : li > 0. The following theorem gives

the necessary and sufficient condition for kis such that the control law is feasible.

Theorem 2. For any γi > 0 and 0 < βi ≤ 1, with a set of positive kis, at least

one solution to (inequality3) and (5.44) lie in the region L ∈ (l1, · · · , lm) : li > 0iff the kis satisfy the following set of inequalities:

k1 > 0,

k2 >a21p1+

k1β2

p2

p2= (1 + γ1 + 1

β2)k1,

· · ·

ki >∑i−1

j=1aijpj+

ki−1

βipi

pi, ∀i < m

· · ·

kn > uM−(Ma+λM )uM−(Ma+hM+uaM+λM+dM )

·∑n−1

j=1anjpj+

kn−1

βnpn

pn,

(5.47)

where, the coefficients pis are computed recursively using the formula

p1 = 1

pi = −∑i−1j=1 ai−1jpj

(5.48)

Proof. =⇒ (Necessary condition):

Assume there exists a solution satisfying (5.44) and (5.30), which lies in the region

L ∈ (l1, · · · , lm) : li > 0. Denote it by

L′ ∈ (l′1, · · · , l′m) : l′i > 0 (5.49)

Let us first prove the following claim using mathematical induction.

Page 120: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

114

Claim 3.

pil′i−1 > pi−1l

′i

pi > 0,(5.50)

for all i ≤ n.

Proof of Claim 3.

(i) Looking at the first row of AL′ < D, we have −k1l′1 + l′2 < 0. From the definition

of p1 and p2 given by (5.48), and noting that a11 = −k1, we have p2l′1 > p1l

′2. Since

l′1, l′2 > 0 and p1 > 0, we obtain p2 > 0. Thus, the claim is true for j = 2.

(ii) Suppose the claim is true for all 2 ≤ j < i. Then, we have

l′j >pjl

′j+1

pj+1>pjpj+1l

′j+2

pj+1pj+2> · · · > pjl

′i−1

pi−1(5.51)

Now, looking at the i− 1 row of AL′ < D, and using aij > 0, ∀i > j, we get

i−1∑

j=1

ai−1jl′j + l′i < 0

⇒i−1∑

j=1

ai−1jpj l′i−1

pi−1+ l′i < 0

⇒ −pil′

i−1

pi−1+ l′i < 0, (using (5.48))

⇒ pil′i−1 > pi−1l

′i.

(5.52)

Since l′i−1, pi−1, l′i > 0, we get pi > 0. Thus, the claim is also true for j = i.

This procedure can be continued until i = m, as the R.H.S of the first m− 1 rows

of AL′ < D are all zeros. This completes the proof of the claim.

Next, it will be shows that the condition on gains, given by (5.47), follows from

(5.44) through the use of claim 3.

For each i < n, from the i-th row of AL′ < D we have

i−1∑

j=1

aijl′j +(

ki−1

βi− ki

)

l′i + l′i+1 < 0

⇒i−1∑

j=1

aijpj l′ipi

+(

ki−1

βi− ki

)

l′i + l′i+1 < 0, (l′j >pj l′ipi, ∀j < i using claim 3)

⇒i−1∑

j=1

aijpj l′ipi

+(

ki−1

βi− ki

)

l′i < 0

⇒[

i−1∑

j=1

aijpj

pi+(

ki−1

βi− ki

)

]

l′i < 0

⇒ ki >∑i−1

j=1aijpj+

ki−1

βipi

pi

(5.53)

Page 121: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

115

From the last row of AL′ < D, we get

m−1∑

j=1

amjl′j +(

km−1

βm− km

)

l′m < −(uaM + hM + dM)

⇒m−1∑

j=1

amjpj l′mpm

+(

km−1

βm− ki

)

l′m < −(uaM + hM + dM)

⇒[

km −m−1∑

j=1

amjpj

pm− km−1

βm

]

l′m > (uaM + hM + dM)

(5.54)

Applying (5.30), which is uM − (Ma + λM) ≥ kml′m, to the above inequality and

eliminating l′m:

(

km −m−1∑

j=1

amjpj

pm− km−1

βn

)

[uM − (Ma + λM)] ≥ km(uaM + hM + dM)

⇒ km > uM−(Ma+λM )(uM−(Ma+hM+uaM+λM +dM ))

·∑m−1

j=1amjpj+

km−1

βmpm

pm,

(5.55)

which completes the proof of necessary condition.

⇐= (Sufficient condition):

Suppose that kis satisfy inequalities (5.47). Then, for i ≤ m− 1,

ki >∑i−1

j=1aijpj+

ki−1

βipi

pi

⇒∑i−1

j=1aijpj+

ki−1

βipi−kipi

pi< 0

⇒ −pi+1

pi< 0, ( using definition of pi and aii = −ki + ki−1

βifrom(5.46))

⇒ pi+1

pi> 0.

(5.56)

Since p1 = 1 > 0, we have pi > 0, ∀i ≤ m.

Next, we are interested in finding the solution of the linear matrix equation AL′ =

D. Using the first m− 2 algebraic equations of AL′ = D, we can represent all the l′is

with i > 1 in terms of l′1:

l′1 = p1l′1

l′2 = p2l′1

· · ·l′m = pml

′1.

(5.57)

Page 122: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

116

Substituting them into the last equation of AL′ = D, we have:

m−1∑

j=1

amjpjl′1 + (km−1

βm− km)pml

′1 = −(uaM + hM + dM)

⇒ l′1 = −(uaM+hM+dM )m−1∑

j=1

amjpj+(km−1

βm−km)pm

.(5.58)

From the last equation of (5.47), since 0 < (uM − (Ma + uaM + hM + λM + dM)) <

uM − (Ma + λM), we have

km >∑m−1

j=1amjpj+

km−1

βmpn

pm

⇒m−1∑

j=1

amjpj + (km−1

βm− km)pm < 0.

(5.59)

Thus, l′1 > 0. As a result, all the l′is are greater than zero. Multiplying both sides of

(5.47) by (uM − (Ma + uaM +hM +λM +dM))pm, and rearranging the terms to factor

out uM , we obtain

kml′m = kmpml

′1 =

−(uaM + hM + dM)knpn

m−1∑

j=1

amjpj + (km−1

βm− km)pm

< uM − (Ma + λM) (5.60)

Thus, we have shown that the only solution to the linear matrix equation AL′ = D

lie in the region (l′1, · · · , l′m) : l′i > 0, kml′m < uM which is an open set. Since A

is a m by m matrix, the solution to inequality AL < D must exist and is a a sector

region with the acme being the point L′ which solves equation AL′ = D. Then, we

can always pick a solution of AL < D to be arbitrarily close to L′ such that L also lies

in the open region (l′1, · · · , l′m) : l′i > 0, kml′m < uM − (Ma + λM). After choosing

that L, we have

(i) AL < D;

(ii) kmlm < uM − (Ma + λM);

(iii) all the lis are greater than zero.

Thus, there exists at least one solution satisfying (5.28)-(5.28), which lies in the

region (l′1, · · · , l′m) : l′i > 0. This proves the sufficiency condition of the theorem.

Page 123: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

117

Remark 3. Condition (5.47) plays an important role in the selection of controller

parameters kis, and can be explained as follows. It can be easily shown that the

closed-loop transfer function from d(t) to z1 (which is also the tracking error y − yd)

inside the linear unsaturated region Ω11 is

[

z1(t)

d(t)

]

=1

sm + knsm−1 + kmkm−1sm−2 + · · ·+m∏

j=2

kj · s +m∏

j=1

kj

. (5.61)

Due to the conditions imposed by (5.47), the closed-loop poles of (5.61) cannot be

assigned arbitrarily. However, as pi depends only on kjs for j < i, from (5.47) we can

easily work out a recursive way of choosing the controller parameters. Specifically, we

first choose k1, and then choose k2 large enough such that the first inequality of (5.47)

is satisfied. Continuing in this fashion, we can get a set of kis that are permissible.

In the next subsection, we clearly outline such a technique for choosing kis such that

all closed-loop requirements are met, without violating the constraints imposed by

(5.47).

Controller Gain Selection: A Recursive Root-Locus Design

As mentioned above, the closed-loop poles of (5.61) cannot be arbitrarily assigned

because of the constraints (5.47) on the controller parameters ki. However, (5.47)

implies that ki can be chosen arbitrarily large, which makes it possible to satisfy the

closed-loop performance requirement. In the linear unsaturated region, the desired

closed-loop performance can be achieved by placing the poles sufficiently far to the

left of imaginary axis e.g., if the required steady-state error is δ, and the transient

response criteria dictates that the slowest closed-loop pole be to the left of p0, then

it is sufficient to place all the poles to the left of pcl = min−√

dM

δ,−p0. In the

following, we propose a recursive root locus design to meet the above requirement.

• Step 1: Select k1 > −pcl, then the root of the equation s+ k1 = 0 lie to the left

of s = pcl.

Page 124: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

118

• Step 2: Let the virtual open-loop system be k2s+k1

s2 , then the virtual closed-loop

characteristic equation is s2 + k2s + k1k2 = 0. To determine k2, draw the root

locus of k2s+k1

s2 . This open-loop system has two poles at origin and one zero at

−k1, left to s = pcl. From the general guidelines for drawing a root-locus, there

exists a k2 large enough such that: 1) the first inequality of (5.47) is satisfied

and 2) all the roots of s2 + k2s + k1k2 = 0 lie to the left of s = pcl, on the real

axis.

.............

.............

.............

• Step m: Let the virtual open-loop system be km

sm−1+km−1sm−2+···+m−1∏

j=1

kj

sm , then

the virtual closed-loop characteristic equation is exactly the same as that of the

actual system, i.e., kmsm−1 + kmkm−1s

m−2 + · · ·+m∏

j=1

kj . To determine kn, draw

the root locus of km

sm−1+km−1sm−2+···+m−1∏

j=1

kj

sm . This open-loop system has m poles

at origin and m− 1 zeros to the left of s = pcl. As the difference in the number

of poles and zeros is one, there always exists an asymptote along the negative

real axis. Thus, for sufficiently large gain km, there is a branch of the root-locus

on the negative real axis. This implies (i) the last two inequalities of (5.47) are

satisfied and (ii) all the closed-loop poles lie to the left of s = pcl on the real

axis.

Thus, we can choose the controller gains such that the desired closed-loop perfor-

mance is achieved, as well as the conditions imposed for the existence of a feasible

control law given by (5.47) are also satisfied simultaneously.

5.3.4 Asymptotic Stability

In the linear unsaturated region, the dynamics can be represented as

Page 125: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

119

z1...

zi

...

zm

=

−k1 1 · · · 0...

... · · · ...

−ki−1...k21 −ki−1...k2(k2 − k1) · · · 0

...... · · · ...

−ki−1...k21 −km−1...k2(k2 − k1) · · · −(km − km−1)

z1...

zi

...

zm

+

0...

0

1

(φ(x, u)T θ + wζ(x)d(t))

⇒ z = Aclz +Bθθ +Bdwζ(x)d(t) (5.62)

where

Bθ =

0 0...

...

0 0

φ1(x, u) φ1(x, u)

, Bd =

0...

0

1

From the preceding discussion, we know Acl is stable and ‖φ(x, u)T‖ is bounded, as

|gζ(x)| is bounded below and above by known positive constants, |u| ≤ uM . Thus, it

can be easily verified using a Lyapunov function Vz = zT z that z(t) is Input to State

Stable (ISS) with respect to the inputs θ and wζ(x)d(t). Now we are ready to state

an important result which shows that the proposed controller preserves the desired

property of an adaptive controller - asymptotic tracking in presence of parametric

uncertainties only.

Theorem 3. In presence of parametric uncertainties only i.e., when d(t) = 0, by

using the control law given by (5.27), and the parameter update law (5.23) along with

filters (5.16-5.17), asymptotic output tracking is also achieved i.e., z → 0 as t→ ∞.

Proof. The proof follows from the fact that the controller is ISS w.r.t θ and θ ∈ L∞

for least-square estimation using x-swapping lemma (see (Ch.6, [31])). This, in turn

implies that z ∈ L∞, and from Barbalat’s lemma, we obtain z → 0 as t→ ∞.

Page 126: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

120

5.4 Simulation Example: Nonlinear Hypersonic Aircraft Model

An adaptive robust control (ARC) based approach was recently proposed to solve

the unknown actuator fault accommodation for linear [56] and nonlinear [57] systems.

The superior performance of an ARC based approach in achieving desired transient

response, as well as small steady-state tracking error over a robust adaptive control

based design was demonstrated through comprehensive simulation studies. In the

present work, we compare the performances of the proposed scheme and the recently

developed ARC based fault-tolerant controller in presence of saturation. A nonlinear

longitudinal model of hypersonic aircraft cruising at a velocity of 15 mach, at an

altitude of 110, 000 feet is used to test the effectiveness of the proposed scheme.

Nominal model of the system is

α = q − γ

q =Myy

Iyy

γ =L+ T sinα

mV− (µ− V 2r) cos(γ)

V r2(5.63)

where

α = angle of attack, rad

γ = flight-path angle, rad

V = velocity, ft/sec

q = pitch rate, rad/sec

T = thrust, lbf

L = lift, lbf

Myy = pitching moment, lbf.ft

Iyy = moment of inertia, slug.ft2

Further, details of the model can be found in [37]. Note that the nominal model of

the system does not take into account any unstructured modeling uncertainties and

Page 127: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

121

external disturbances. As modeling uncertainties are inherent to any realistic system

model, unmatched uncertainties will be introduced in order to make the simulation

studies more meaningful. The state-space representation is given by

x1 = x2 + a1y + a2 sin(y) + a3y2 sin(y) + a4 cos(x3)

x2 = a5y2 + a6y + (a7 + a8y + a9y

2)x2 + b1u1 + b2u2 + ∆(t)

x3 = a10 cos(x3) − a1y − a2 sin(y)

y = x1 (5.64)

where [x1, x2, x3] = [α, q, γ] and ∆(t) = 0.02 sin(3t) represents the input disturbance.

The nominal plant parameters are

a1 = −0.0427, a2 = −3.4496×10−4, a3 = 5×10−5, a4 = 0.0014,

a5 = −4.2006, a6 = 1.0821, a7 = −3.6896, a8 = 0.1637,

a9 = −0.1242, a10 = 0.0014, b1 = 0.8, b2 = 0.8

The initial conditions are set to x(0) = [0, 0.01, 0]T .

In ARC based technique, once the controller saturates, it cannot return to the un-

saturated region of controller operation. On the other hand, in the proposed design,

the controller saturates temporarily, but returns to the unsaturated region. This can

be explained as follows. Following an actuator fault, the performance requirements

necessitate that the transients be suppressed using large feedback action. Note that

undesirable transients can increase the demanded control input on two accounts: first,

it induces a large feedback action due to the proportional type controller i.e., −kizi,

and second, since the model-compensation depends on z as well i.e., ϕ(x) = ϕ(xd+z),

as z increase, ϕ(x) may increase correspondingly. However, as the control input is

limited, the transients cannot be suppressed effectively after model-compensation,

which further increases the transient error. This, in turn, demands larger control

input, leading to a controller saturation scenario, which is unsalvageable. In the pro-

posed approach, on the other hand, when the error is large, we sacrifice the model

compensation to certain extent, and use the available control input to supply maxi-

Page 128: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

122

mum possible feedback action −kizi, such that zi can be made negative, and the error

can be reduced to an extent which allows the controller to be unsaturated. This can

also reduce the required model-compensation.

As can be seen from fig. 1, in the ARC based approach, once the controller gets

saturated, the adaptation mechanism breaks down, and the estimated parameter

assumes the minimum or maximum value used in the projection algorithm. This is

not surprising, as the adaptation mechanism is driven by the error, which has now

two sources - mismatch in estimated parameters, and error due to the saturation

effect. But, as in the proposed approach an indirect scheme is used, the model

structure does not change whether the controller is saturated or not, and this results

in accurate estimation.

Page 129: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

123

0 20 40 60 80 100 120−20

−15

−10

−5

0

5

time (sec)

y an

d y

d (rad

)

Commanded versus actual angle of attack

RefARC

20 40 60 80 100−0.5

0

0.5

0 20 40 60 80 100 120

−0.2

−0.1

0

0.1

0.2

0.3

time (sec)

u(t)

(rad

)

Control signal (elevator angle): ARC

u1(t)

u2(t)

max(u)

min(u)

Fault

0 20 40 60 80 100 1200

0.5

1

1.5

2

2.5

time (sec)

κ est

Estimated controller gain ( κest

) vs. time

κestmax(κ)

min(κ)

0 20 40 60 80 100 120−1.5

−1

−0.5

0

0.5

1

1.5

time (sec)

µ est

Estimated fault ( µest

) vs. time

µest

min(µ)

max(µ)

Figure 5.2. Comparative results for stabilization in absence of distur-bances

Page 130: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

124

0 20 40 60 80 100 120−0.2

−0.15

−0.1

−0.05

0

0.05

0.1

0.15

time (sec)

y an

d y

d (rad

)

Commanded versus actual angle of attack

ReferenceOutput

0 20 40 60 80 100 120

−0.2

−0.1

0

0.1

0.2

0.3

time (sec)

u(t)

(rad

)

Control signal (elevator angle): Saturated ARC

u1(t)

u2(t)

min(u)

max(u)

Saturation

Fault

0 20 40 60 80 100 1200

0.5

1

1.5

2

2.5

time (sec)

κ est

Estimated controller gain ( κest

) vs. time

κest

min(κ)

max(κ)

0 20 40 60 80 100 120−0.1

−0.05

0

0.05

0.1

time (sec)

µ est

Estimated fault ( µest

) vs. time

µest

Figure 5.3. Comparative results for stabilization in presence of distur-bance

Page 131: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

125

5.5 Conclusion

In this chapter, an indirect adaptive robust scheme was proposed to accommodate

unknown actuator faults in presence of actuator magnitude constraints. Feedback

linearizable systems with matched uncertainties were considered in the present work.

Actuator saturation disrupts the functioning of an actuator fault-tolerant controller

in two ways - first, the working actuators can saturate due to the undesired transients

following an actuator fault and second, once the actuators saturate, the adaptation

may become unreliable. A backstepping based approach was proposed to deal with

the first issue, which explicitly takes into account the actuator saturation limits. In

the proposed approach, performance requirements were relaxed once the actuator

saturates, and more emphasis was put on returning the actuator to the unsaturated

mode. Second, it was shown that an indirect adaptive scheme, with controller and

estimator separation, would yield reliable parameter estimates despite actuator satu-

ration. Another important feature of the proposed scheme is that once all the states

are within a pre-determined region where the controller is unsaturated, certain desired

closed-loop properties, like disturbance attenuation to any extent could be recovered.

Furthermore, it was shown that the error would asymptotically converge to zero in

in spite of actuator faults, in presence parametric uncertainties only. Comparative

simulation studies showed that in presence of actuator saturation limits, which would

make an otherwise stable fault-tolerant control scheme unstable, the proposed design

could achieve desired control objectives.

Page 132: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

126

6. CONCLUSIONS AND FUTURE WORK

6.0.1 Conclusions

In this dissertation, an output feedback based Adaptive Robust Fault-Tolerant

Control (ARFTC) scheme was presented for unknown actuator fault accommoda-

tion. The proposed scheme is applicable to a class of uncertain linear and nonlinear

systems and addresses a large class of actuator faults. A critical review of the existing

literature revealed two problems associated with actuator fault-tolerant control - (i)

undesired transients and (ii) unacceptably large steady-state tracking errors. Both

the problems cannot be addressed by any adaptive control or robust control based

fault-tolerant scheme when used individually. In the proposed approach, adaptation

and robust feedback were used simultaneously to achieve desired transients and main-

tain tracking performance in face of large parametric uncertainties introduced due to

failing actuators. For the linear case, comparative simulation studies were done us-

ing a linearized model for lateral motion of Boeing 747. For nonlinear systems, a

Hypersonic aircraft model was used to evaluate and compare the performance of the

proposed design with robust backstepping based design. It could be concluded based

on the comprehensive simulation studies that the proposed scheme can achieve su-

perior performance over a conventional robust MRAC or robust backstepping based

based technique.

One of the assumption made in the first part of work was that the healthy ac-

tuators always had sufficient control authority to accommodate the actuator faults.

In fact, this has become a standard assumption in the adaptive control based fault-

tolerant control designs. However, ignoring actuator saturation can lead to a disas-

trous outcome, as it can easily disrupt the functioning of an adaptive control based

fault-tolerant scheme in the following ways. First, the working actuators can satu-

Page 133: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

127

rate due to the undesired transients following an actuator fault and second, once the

actuators saturate, the adaptation may become unreliable. As the first step towards

solving this problem, we solved the problem of global stabilization of an integrator

chain using a conceptually different approach. This problem was first solved by Teel in

1992. Based on Teel’s work, many modifications had been proposed in the literature

to improve the performance of the controller. In our analysis, it was clearly shown

that all such schemes exhibit poor robustness properties with respect to input distur-

bance and leads to conservative design. In fact, a quantitative analysis revealed that

even when the magnitude of the disturbance is less than that of the available control

input, the coordinate transformation can render the design of a stabilizing controller

impossible. These limitations could not be overcome by any modification based on

Teel’s work, as coordinate transformation is an essential step in all such designs. In

order to remove these limitations, we took a fundamentally different viewpoint and

proposed a scheme which did not rely on coordinate transformation, and was based

on backstepping design. Comparative studies were performed on a third order chain

of integrator to show the superior performance of the proposed technique. The first

set of studies, performed in absence of disturbances, revealed that the convergence

rate of the proposed scheme is at least as good as that proposed by Marchand and

Hably, which is based on Teel’s work. In presence of disturbances, however, signif-

icant differences could be seen in terms of disturbance attenuation and convergence

of the states. A tracking problem with large disturbance, where a stable controller

could not be designed due to coordinate transformation, was also solved to show the

effectiveness of the proposed scheme.

Finally, we combined the adaptive robust and input saturated control designs to

propose a saturated adaptive robust actuator-fault tolerant controller. Feedback lin-

earizable systems with matched uncertainties were considered in the present work.

Controller design using feedback linearization is a popular technique for flight control

applications, as it makes it possible to apply linear control theory to a nonlinear sys-

tem. As mentioned earlier, there were two chief problems associated with actuator

Page 134: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

128

saturation - saturation of healthy actuators due to undesired transients and unreliable

adaptation. A backstepping based approach was proposed to deal with the first issue,

which explicitly takes into account the actuator saturation limits. In the proposed

approach, performance requirements were relaxed once the actuator saturates, and

more emphasis was put on returning the actuator to the unsaturated mode. Sec-

ond, it was shown that an indirect adaptive scheme, with controller and estimator

separation, would yield reliable parameter estimates despite actuator saturation. An-

other important feature of the proposed scheme is that once all the states are within a

pre-determined region where the controller is unsaturated, certain desired closed-loop

properties, like disturbance attenuation to any extent could be recovered. Further-

more, it was shown that the error would asymptotically converge to zero in in spite

of actuator faults, in presence parametric uncertainties only. Comparative simulation

studies performed on a Hypersonic aircraft model showed that in presence of actuator

saturation limits, which would make an otherwise stable fault-tolerant control scheme

unstable, the proposed design could achieve desired control objectives.

6.0.2 Future Work

1. In the present work, an adaptive robust approach was proposed to accommodate

unknown actuator faults. A similar approach can also be developed to address

a large of sensor faults. Some common type of sensor faults include constant

bias failures, drift or additive-type sensor failures and multiplicative-type sensor

failures [58]. As the effect of these faults can be captured by a parametric

model, a similar framework as developed in this thesis can be used to address

these faults.

2. The backstepping based approach was proposed in this work to deal with in-

put saturation, which differs significantly from the conventional “cancellation”

backstepping design. This resulted in a simpler control law, which was shown to

be bounded for an integrator chain, and for a class of feedback linearizable non-

Page 135: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

129

linear system with some assumptions on the nonlinearities and uncertainties. A

natural extension of the proposed technique would be the design of a bounded

controller for systems in parametric feedback form with bounded uncertainties

and nonlinearities. Furthermore, based on the bounds of the nonlinearities, ap-

proximation schemes can be developed for the region of attraction, which would

be a valuable addition to existing literature on saturated control.

3. Based on the work proposed in this dissertation and the previous work on

adaptive robust fault-detection [59], an active fault-tolerant design seems like a

natural extension. The proposed fault-tolerant control accommodates the fault

without an active FDI module. As in many practical applications it may be

necessary to identify and isolate the fault, an adaptive robust fault-tolerant

design with room for exchanging information with an FDI module should be

investigated.

Page 136: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

LIST OF REFERENCES

Page 137: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

130

LIST OF REFERENCES

[1] A. R. Teel. Global stabilization and restricted tracking for multiple integratorswith bounded controls. Systems & Control Letters, 18(3):165–171, 1992.

[2] J.M. Maciejowski and C.N. Jones. Mpc fault-tolerant flight control case study:Flight 1862. In IFAC Safeprocess Conference, 2003.

[3] G. Tao, S. Chen, and S. M. Joshi. An adaptive actuator failure compensationcontroller using output feedback. IEEE Trans. on Automat. Contr., 47(3):506–511, 2002.

[4] X. Tang, G. Tao, and S.M. Joshi. Adaptive actuator failure compensationfor nonlinear MIMO systems with an aircraft control application. Automatica,43(11):1869–1883, 2007.

[5] X. Tang, G. Tao, and S.M. Joshi. Adaptive output feedback actuator failurecompensation for a class of non-linear systems. International Journal of AdaptiveControl and Signal Processing, 19(6), 2005.

[6] J. D. Boskovic, S. H. Yu, and R. K. Mehra. Stable adaptive fault-tolerant con-trol of overactuated aircraft using multiple models, switching and tuning. InAIAA Guidance, Navigation, and Control Conference and Exhibit, pages 739–749, Boston, MA, 1998.

[7] X. Tang, G. Tao, and S.M. Joshi. Adaptive actuator failure compensation forparametric strict feedback systems and an aircraft application. Automatica,39(11):1975–1982, 2003.

[8] Jovan D. Boskovic, S. E. Bergstrom, and Raman K. Mehra. Robust integratedflight control design under failures, damage and state-dependent disturbances.AIAA Jl. of Guidance, Control and Dynamics, 28(5).

[9] M. Krstic, P.V. Kokotovic, and I. Kanellakopoulos. Transient performance im-provement with a new class of adaptive controllers. Systems & Control Letters,21(6):451–461, 1993.

[10] A. Datta and PA Ioannou. Performance analysis and improvement in model ref-erence adaptive control. IEEE Transactions on Automatic Control, 39(12):2370–2387, 1994.

[11] ML Corradini, G. Orlando, and G. Parlangeli. A fault tolerant sliding modecontroller for accommodating actuator failures. In 44th IEEE Conference on De-cision and Control, 2005 and 2005 European Control Conference. CDC-ECC’05,pages 3091–3096, 2005.

[12] D. Kim and Y. Kim. Robust variable structure controller design for fault tolerantflight control. Journal of Guidance Control and Dynamics, 23(3):430–437, 2000.

Page 138: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

131

[13] H. Alwi and C. Edwards. Application of fault tolerant control using slidingmodes with on-line control allocation on a large civil aircraft. In Proc. of theIEEE Multi-conference on Systems and Control, Singapore, 2007.

[14] Z. Gao and S.X. Ding. Actuator fault robust estimation and fault-tolerant controlfor a class of nonlinear descriptor systems. Automatica, 43(5):912–920, 2007.

[15] J. Zhang and P. A. Ioannou. Stable adaptive control for performance improve-ment. In Proc. of the 45th IEEE Conf. on Decision and Control, pages 1087–1092,San Diego, California, 2006.

[16] B. D. O. Anderson, T. Brinsmead, D. Liberzon, and A. S. Morse. Multiplemodel adaptive control with safe switching. Int. J. of Adaptive Control andSignal Processing, 15:445–470, 2001.

[17] Bin Yao and M. Tomizuka. Adaptive robust control of MIMO nonlinear systemsin semi-strict feedback forms. Automatica, 37(9):1305–1321, 2001. Parts of thepaper were presented in the IEEE Conf. on Decision and Control, pp2346-2351,1995, and the IFAC World Congress, Vol. F, pp335-340, 1996.

[18] Bin Yao. High performance adaptive robust control of nonlinear systems: ageneral framework and new schemes. In Proc. of IEEE Conference on Decisionand Control, pages 2489–2494, San Diego, 1997.

[19] P.C. Shrivastava and RF Stengel. Stability boundaries for aircraft with unsta-ble lateral-directional dynamics and control saturation. Journal of Guidance,Control and Dynamics, 12(1):62–70, 1989.

[20] M.A. Dornheim. Report pinpoints factors leading to YF-22 crash. Aviation Weekand Space Technology, 137(19):53–54, 1992.

[21] D.S. Bernstein and A.N. Michel. A chronological bibliography on saturatingactuators. International Journal of Robust and Nonlinear Control, 5(5):375–380,1995.

[22] T. Hu and Z. Lin. Control systems with actuator saturation: Analysis and design.Birkhauser, 2001.

[23] V. Kapila and KM Grigoriadis. Actuator Saturation Control. Actuator Satura-tion Control, pages 1–301, 2002.

[24] SP Karason and AM Annaswamy. Adaptive control in the presence of inputconstraints. IEEE Transactions on Automatic Control, 39(11):2325–2330, 1994.

[25] AM Annaswamy and J.E. Wong. Adaptive control in the presence of saturationnon-linearity. International Journal of Adaptive Control and Signal Processing,11(1):3–19, 1998.

[26] E.N. Johnson and S.K. Kannan. Adaptive flight control for an autonomousunmanned helicopter. In AIAA Guidance, Navigation and Control Conference,volume 11, 2002.

[27] F. Liao, J. L. Wang, and G. H. Yang. Reliable robust flight tracking control: anlmi approach. IEEE Trans. on Control System Technology., 10(1):76–89, 2002.

Page 139: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

132

[28] Jovan D. Boskovic and Raman K. Mehra. Multiple-model adaptive flight controlscheme for accommodation of actuattor failures. AIAA Jl. of Guidance, Controland Dynamics, 25(4).

[29] Bin Yao and Li Xu. Output feedback adaptive robust control of uncertain linearsystems with disturbances. ASME Journal of Dynamic Systems, Measurement,and Control, 128:938–945, 2006. (in press). Part of the paper appeared in theProc. of 1999 American Control Conference, pp556-560.

[30] C. T. Chen. Linear system theory and design. Saunders college publishing,Philadelphia, PA, USA, 1984.

[31] Miroslav. Krstic, I. Kanellakopolous, and Petar V. Kokotovic. Nonlinear andAdaptive Control Design. New York: Wiley, 1995.

[32] Bin Yao and M. Tomizuka. Adaptive robust control of SISO nonlinear systems ina semi-strict feedback form. Automatica, 33(5):893–900, 1997. (Part of the paperappeared in Proc. of 1995 American Control Conference, pp2500-2505, Seattle).

[33] Y. Liu and G. Tao. An adaptive disturbance rejection algorithm for mimo sys-tems with an aircraft flight control application. In Proc. of AIAA Guidance,Navigation, and Control Conf., South Carolina, USA, 2007.

[34] G. F. Franklin, J. D. Powell, and A. Emami-Naeni. Feedback Control of DynamicsSystems. Prentice Hall, 2002.

[35] S. S. Mulgund and R. F. Stengel. Aircraft flight control in wind shear using se-quential dynamic inversion. Jl. of Guidance, Control and Dynamics, 18(5):1084–1091, 1995.

[36] P. Pisu and A. Serrani. Attitude tracking with adaptive rejection of rate gyros.IEEE Trans. on Automat. Contr., 52(12):2374–2379, 2007.

[37] H. Xu and M. Mirmirani. Robust adaptive sliding mode control for a class ofmimo nonlinear systems. In Proc. of the 2001 AIAA Guidance, Navigation andControl Conference, 2001.

[38] M.V. Kothare, P.J. Campo, M. Morari, and C.N. Nett. A unified framework forthe study of anti-windup designs. Automatica, 30:1869–1869, 1994.

[39] F. Wu, KM Grigoriadis, and A. Packard. Anti-windup controller design us-ing linear parameter-varying control methods. International Journal of Control,73(12):1104–1114, 2000.

[40] A.R. Teel. Anti-windup for exponentially unstable linear systems. InternationalJournal of Robust and Nonlinear Control, 9(10):701–716, 1999.

[41] F. Tyan and D.S. Bernstein. Anti-windup compensator synthesis for systems withsaturating actuators. International Journal of Robust and Nonlinear Control,5(5):521–538, 1995.

[42] D.Q. Mayne, J.B. Rawlings, C.V. Rao, and P.O. Scokaert. Constrained modelpredictive control: Stability and optimality. Automatica, 36:789–814, 2000.

[43] F. A. Cuzzola, J. C. Geromel, and M. Morari. An improved approach for con-strained robust model predictive control. Automatica, 38(7):1183–1189, 2002.

Page 140: AN ADAPTIVE ROBUST APPROACH TO ACTUATOR FAULT …

133

[44] F. Mazenc and L. Praly. Adding integrations, saturated controls, and sta-bilization forfeedforward systems. IEEE Transactions on Automatic Control,41(11):1559–1578, 1996.

[45] R. Freeman and L. Praly. Integrator backstepping for rounded controls andcontrol rates. IEEE Transactions on Automatic Control, 43(2):258–262, 1998.

[46] F. Mazenc and A. Iggidr. Backstepping with bounded feedbacks. Systems &Control Letters, 51(3-4):235–245, 2004.

[47] G. Kaliora and A. Astolfi. Nonlinear control of feedforward systems with boundedsignals. IEEE Transactions on Automatic Control, 49(11):1975–1990, 2004.

[48] L. Marconi and A. Isidori. Robust global stabilization of a class of uncertainfeedforward nonlinear systems. Systems & control letters, 41(4):281–290, 2000.

[49] J.Q.Gong and Bin Yao. Global stabilization of a class of uncertain systems withsaturated adaptive robust controls. In IEEE Conf. on Decision and Control,pages 1882–1887, Sydney, 2000.

[50] N. Marchand and A. Hably. Global stabilization of multiple integrators withbounded controls. Automatica, 41(12):2147–2152, 2005.

[51] B. Zhou, G.R. Duan, and Z.Y. Li. On improving transient performance in globalcontrol of multiple integrators system by bounded feedback. Systems & ControlLetters, 57(10):867–875, 2008.

[52] A.J. Krener. Approximate linearization by state feedback and coordinate change.Systems & Control Letters, 5(3):181–185, 1984.

[53] D. Enns, D. Bugajski, R. Hendrick, and G. Stein. Dynamic inversion: an evolvingmethodology for flight control design. International Journal of Control, 59(1):71–91, 1994.

[54] S.N. Singh, M. Steinberg, and R.D. DiGirolamo. Nonlinear predictive controlof feedback linearizable systems and flight control system design. Journal ofGuidance, Control, and Dynamics, 18(5):1023–1028, 1995.

[55] S.N. Singh and M. Steinberg. Adaptive control of feedback linearizable nonlinearsystems with application to flight control. Journal of Guidance Control andDynamics, 19(4):871–877, 1996.

[56] S. Gayaka and B. Yao. An adaptive robust scheme for multiple actuator faultaccommodation. In Proc. of ASME Dynamic Systems and Control Conference,Ann Arbor, Michigan, 2008.

[57] S. Gayaka and B. Yao. An output feedback based adaptive robust fault tolerantcontrol scheme for a class of nonlinear systems. In Proc. of ASME DynamicSystems and Control Conference, Hollywood, California, 2009.

[58] G. Heredia, A. Ollero, M. Bejar, and R. Mahtani. Sensor and actuator faultdetection in small autonomous helicopters. Mechatronics, 18(2):90–99, 2008.

[59] P. Garimella. An Adaptive Robust Framework for Model Based Fault Detection.PhD thesis, Purdue University, West Lafayette, USA, 2005.