You are on page 1of 153

ADAPTIVE CONTROL WITH AEROSPACE APPLICATIONS

by

Ross Gadient

A Dissertation Presented to the


FACULTY OF THE USC GRADUATE SCHOOL
UNIVERSITY OF SOUTHERN CALIFORNIA
In Partial Fulfillment of the
Requirements for the Degree
DOCTOR OF PHILOSOPHY
(ELECTRICAL ENGINEERING)

May 2013

Copyright 2013 Ross Gadient


Acknowledgments

First I would like to thank my advisor Dr. Petros Ioannou, who has been generous

with his time, his guidance, and his wisdom. I am grateful for our many discussions

which led to the exploration of topics contained within. His optimism and enthusiasm

helped fuel my research efforts even at times when hope seemed lost. I consider myself

fortunate to have had an advisor who provided such expertise coupled with sincere

interest and care.

The members of my research committee deserve thanks for providing insight and

constructive criticism, along with enriching my knowledge in the classroom. Professor

Michael Safonov, Professor Henryk Flashner, and Professor Edmond Jonckheere all

contributed in significant ways to the construction of my dissertation and the direction of

my research.

I would like to acknowledge those who originally stoked my interest in the area of

control theory during my days at the University of Illinois. Professor Daniel Metz,

Professor Carolyn Beck, and Professor Juraj Medanic all played pivotal roles in my

introduction to control theory and subsequent continuing study. Special considerations go

to Professor Francesco Bullo, who helped to broaden my research horizons while serving

as my M.S. thesis advisor.

I am indebted to professional colleagues and friends who have helped me better

understand both control theory and application, and I hope my research has positively

contributed to these areas. I would like to express my gratitude to Irene Gregory,

Professor Gary Balas, Professor Anuradha Annaswamy, Professor Naira Hovakimyan,

ii
Brian Whitehead, Travis Gibson, Jason Levin, Ryan Ratliff, Joseph Brinker, and many

others for their contributions.

Special thanks go to Eugene Lavretsky and Kevin Wise for their encouragement, for

their continuing demonstration of excellence in research, application, and teaching of

control theory, and for their friendship.

Most of all, I would like to thank my family: My grandparents Margaret & John and

Bette & Richard for instilling an early love of learning, the denizens of Onanguisse for

sharing positive energy, my brother John for providing inspiration and countless laughs,

my parents Denise and Jeff for their unconditional love and support, and my wife Kat for

her grace, her patience, and her love.

iii
Table of Contents

Acknowledgments............................................................................................................... ii

List of Figures ..................................................................................................................... v

Abstract ............................................................................................................................. vii

Introduction ......................................................................................................................... 1

Chapter 1: Model Following Using Dynamic Inversion.................................................... 7


Chapter 1.1: Baseline Dynamic Inversion Design ....................................................... 8
Chapter 1.2: Aircraft Longitudinal Design ................................................................ 11
Chapter 1.3: Aircraft Directional Design ................................................................... 14

Chapter 2: Model Following Using Dynamic Inversion Controller with State Limiting 18
Chapter 2.1: State Limiter – Motivating Example ..................................................... 20
Chapter 2.2: State Limiter Application to Aircraft Dynamics ................................... 36
Chapter 2.3: Application to X-48B Aircraft Simulation ............................................ 38
Chapter 2.4: Flight Test of the X-48B AOA and AOS Limiting System .................. 49

Chapter 3: Dynamic Inversion with Adaptive Augmentation ......................................... 62


Chapter 3.1: PID Control for Cascaded Systems with Uncertain Dynamics ............. 63
Chapter 3.2: Parameter Adaptation and Closed-Loop System Dynamics ................. 67
Chapter 3.3: Design Example: AOA Tracking .......................................................... 71

Chapter 4: Dynamic Inversion with State Limiting and Adaptive Augmentation ........... 81
Chapter 4.1: Baseline+Adaptive Control+State Limiting for Uncertain Dynamics .. 83
Chapter 4.2: Design Example: AOA Tracking .......................................................... 89

Chapter 5: Adaptive Design with Improved Performance under Input Time-Delays ... 100
Chapter 5.1: Problem Definition and Baseline Optimal Control Design................. 103
Chapter 5.2: Adaptive Augmentation via Classical MRAC Architecture ............... 110
Chapter 5.3: Alternative Adaptive Augmentation via B-SPM Model ..................... 114
Chapter 5.4: Design Example: AOA Tracking ........................................................ 118

Chapter 6: Concluding Remarks and Suggestions for Future Work.............................. 137

Bibliography ................................................................................................................... 142

iv
List of Figures

Figure 1: Limiter Onset Modulation Function 22


Figure 2: Step Response Comparison – Nominal vs. Limited System 24
Figure 3: AOA Limiter Performance in Windup Turn 40
Figure 4: Elevator Response for AOA Windup Turn 42
Figure 5: Reference Model Parameters for AOA Windup Turn 43
Figure 6: AOS Limiter Performance, Step Input 45
Figure 7: Rudder Response for AOA Step Command 46
Figure 8: Reference Model Parameters for AOS Step Command 48
Figure 9: X-48B BWB Configuration 49
Figure 10: X-48B Positive AOA Envelope Limits 51
Figure 11: Windup/Wind-down Turn V-n Diagram 52
Figure 12: Representative Windup/Wind-down Turn V-n Diagram 53
Figure 13: Representative Windup/Wind-down Turn Airspeed/AOA Plot 54
Figure 14: AOA Limiter Results, Slat Retracted/Aft CG 56
Figure 15: AOA Limiter Results, Slat Extended/Fwd CG 57
Figure 16: Representative Sideslip Limiter Maneuvers 59
Figure 17: Sideslip Limiter Results, Slat Ret/Aft CG 60
Figure 18: Sideslip Limiter Results, Slat Ext/Aft CG 61
Figure 19: Representation of the Sets Sr  b  B  SR 70
Figure 20: Baseline Closed-Loop Response to AOA Doublet without Uncertainties 74
Figure 21: Total Matched Uncertainty vs. AOA at q=0 76
Figure 22: Baseline Closed-Loop Response in Presence of Matched Uncertainties 77
Figure 23: Baseline+Adaptive Closed-Loop Response in Presence of Uncertainties 79
Figure 24: Adaptive Parameter Norm vs. Time 80
Figure 25: Baseline Closed-Loop Response to AOA Doublet without Uncertainties 92
Figure 26: Closed-Loop Response to AOA Doublet: No Uncertainties, with Limiter 93
Figure 27: AOA Doublet Response: No Uncertainties, with Limiter (Zoom) 94
Figure 28: AOA Doublet Response: Uncertainties with Adaptive+Limiter 96
Figure 29: AOA Doublet Response: Uncertainties with Adaptive+Limiter (Zoom) 97
Figure 30: Adaptive Parameter Norm vs. Time 98
Figure 31: Time Response of Pure Time-Delay and Padé Approximations 106
Figure 32: Frequency of Pure Time-Delay and Padé Approximations 107
Figure 33: Baseline Closed-Loop Response to AOA Doublet without Uncertainties 120
Figure 34: Baseline Closed-Loop Response in Presence of Matched Uncertainties 122
Figure 35: Baseline+MRAC Closed-Loop Response in Presence of Uncertainties 124
Figure 36: MRAC Adaptive Parameter Norms vs. Time 125
Figure 37: Baseline+MRAC Closed-Loop Response with High Adaptive Gains 126
Figure 38: Baseline+B-SPM Normalized Response in Presence of Uncertainties 128
Figure 39: Baseline+B-SPM Adaptive Parameter Norms vs. Time 129
Figure 40: Comparison of Adaptive Closed-Loop Responses: with Uncertainties 130

v
Figure 41: Comparison of Control Input Power Spectral Densities 131
Figure 42: Baseline Response: Input Time-Delay of 935ms 133
Figure 43: Baseline + B-SPM Normalized Response: Input Time-Delay of 935ms 135

vi
Abstract

Robust and adaptive control techniques have a rich history of theoretical development

with successful application. Despite the accomplishments made, attempts to combine the

best elements of each approach into robust adaptive systems has proven challenging,

particularly in the area of application to real world aerospace systems. In this research,

we investigate design methods for general classes of systems that may be applied to

representative aerospace dynamics. By combining robust baseline control design with

augmentation designs, our work aims to leverage the advantages of each approach.

This research contributes the development of robust model-based control design for

two classes of dynamics: 2nd order cascaded systems, and a more general MIMO

framework. We present a theoretically justified method for state limiting via

augmentation of a robust baseline control design. Through the development of adaptive

augmentation designs, we are able to retain system performance in the presence of

uncertainties. We include an extension that combines robust baseline design with both

state limiting and adaptive augmentations. In addition we develop an adaptive

augmentation design approach for a class of dynamic input uncertainties. We present

formal stability proofs and analyses for all proposed designs in the research.

Throughout the work, we present real world aerospace applications using relevant

flight dynamics and flight test results. We derive robust baseline control designs with

application to both piloted and unpiloted aerospace system. Using our developed

methods, we add a flight envelope protecting state limiting augmentation for piloted

aircraft applications and demonstrate the efficacy of our approach via both simulation and

vii
flight test. We illustrate our adaptive augmentation designs via application to relevant

fixed-wing aircraft dynamics. Both a piloted example combining the state limiting and

adaptive augmentation approaches, and an unpiloted example with adaptive augmentation

show the ability of our approach to retain desired performance in the presence of relevant

system uncertainty. Finally, we present alternative adaptive augmentation design

developed to mitigate time delays at the system input and which demonstrates significant

improvement over an existing widely used adaptive augmentation approach when applied

to fixed wing aircraft dynamics.

viii
Introduction

Demands on safety and performance have necessitated the development of

increasingly sophisticated flight control systems. The design of autopilots for high-

performance aircraft was one of the primary motivations for active research in adaptive

control in the 1950s. In comparison to gain scheduling, adaptive control has a learning

capability which identifies changes in the parameters of the flight dynamics by

processing the input/output data and adjusts its parameters accordingly. Adaptive designs

may offer benefits over fixed-gain counterparts such as improved performance, increased

robustness to uncertainties, decreased design cycle time, and lower cost. These benefits

are a byproduct of their ability to adjust control parameters as a function of online

measurements. Early attempts at adaptive flight control used controllers with unproven

stability properties, which sometimes led to disastrous consequences [5]. Since that time,

significant research in the field of robust and adaptive control has allowed the design of

stable adaptive systems. Various adaptive control methods have been developed for

controlling dynamic systems with parametric and dynamic uncertainties [6]-[17]. These

techniques have been extended and applied to the problem of aircraft control. Recently, a

dynamic inversion baseline controller on the X-36 tailless fighter was augmented by an

adaptive neural network flight control system [18]. In addition, adaptive augmentation

has been used to augment baseline LQR control design in production joint direct attack

munitions (JDAM) unpowered weapons [19]. In this thesis, we take a similar adaptive

augmentation approach to a robust baseline controller and apply to dynamics

representative of an aircraft in all three axes.

1
This thesis considers two model following control system architectures designed to

provide robust baseline with adaptive tracking performance, applied to relevant flight

dynamics. First, given second-order dynamical systems in cascaded form, we design a

baseline model-reference dynamic inversion based tracking control law in both explicit

and implicit forms. Interest in considering this particular class of systems stems from

flight control related applications, where inner-loop controllers for fixed-wing aircraft are

often designed based on simplified models [1]-[4]. These models represent the decoupled

fast responses in aircraft pitch, roll and yaw axes and are given in such cascaded form.

The benefit of such a baseline control design is two-fold. Such a design allows for

favorable closed-loop system characteristics in terms of traditional robustness measures

and disturbance rejection. In addition, the choice of dynamic inversion architecture

allows the designer to satisfy both aircraft performance and handling qualities

requirements in a convenient manner. The latter is of significant importance in the case of

piloted vehicles, where well-known robust control techniques such as LQR may be used

to provide good margins and robustness but may not provide insight into handling

qualities.

An additional challenge addressed is the need to keep the system dynamics in a

prescribed region of the corresponding state space. This region is often referred to as the

“operational flight envelope”. For example, an Angle-of-Attack (AOA) command

tracking controller must include an AOA protection system (often called the “AOA

Limiter”), whose purpose is to maintain the aircraft AOA within a pre-specified range,

outside of which a loss of control is expected. Vehicles with control augmentation

2
systems rely on state limiters to prevent the exceeding of predetermined state limits.

These limiters typically activate once predefined envelope boundaries are exceeded and

introduce sharp changes to control characteristics once active. The resulting

nonlinearities make analysis using conventional control system methods difficult or

invalid. Current state-of-the-art in limiter design consists of ad-hoc limiting schemes

developed specifically for the system under consideration. As a result, such methods

require extensive design and test iterations specific to each individual application. Often,

these designs do not provide stability / performance guarantees. The proposed method

introduces a formal yet numerically efficient limiting technique which gradually modifies

the expected behavior of the system dynamics when user-specified criteria are

approached / exceeded. In essence, such a controller has to blend two subsystems, the

tracker and the limiter, with seamless transitions between the two controllers while

preserving closed-loop stability at all times. In this thesis we present a state limiting

augmentation design via formal methods that keeps the system state in a prescribed

region of the state space while maintaining closed-loop stability.

Another challenge addressed in this thesis is the use of adaptive augmentation to

allow the control system to maintain desired performance in the presence of system

uncertainties. Aircraft operate over a wide range of speeds and altitudes, and their

dynamics are nonlinear and often time-varying. As an aircraft moves through different

flight conditions, the linear design model at an operating point changes. A controller

designed for each operating point can be scheduled to be switched on when the operating

point is reached. Between operating points interpolation or other techniques may be used

3
to modify or mix the controllers of the neighboring operating points. This approach is

referred to as gain scheduling, and does not necessarily possess guarantees of closed-loop

stability when the system undergoes unexpected changes in its dynamics and

environment. We propose the use of adaptive augmentation to provide closed-loop

stability guarantees and improved performance in the presence of system uncertainties.

A resulting development derived in this thesis is the synthesis of a robust dynamic

inversion baseline controller with both state limiting and adaptive augmentation devices.

This design allows for insight into flying qualities design for baseline control of an aerial

vehicle, state limiting to keep the vehicle within a prescribed region of the system state

space, and adaptive augmentation to mitigate matched uncertainties. The combination of

all three elements into an aggregate design yields a flight control design that is both

theoretically justified as well as relevant to real-world flight control problems where

concerns such as permissible flight envelope restrictions are important to maintain safe

piloted flight.

In addition to dynamic inverse based control for a class of cascaded 2nd-order

systems, we also develop robust baseline control for aerial vehicles using LQR-optimal

approach. Again using an augmentation approach, we apply classical MRAC adaptive

augmentation to representative flight dynamics to mitigate matched uncertainties. We

advance existing results by formulating the problem to specifically address and mitigate

input time-delay, which is of crucial importance in aerospace systems due to such factors

as sensor latency and actuator performance uncertainty. After reformulating the classical

MRAC architecture to address matched uncertainties and input time-delay, we apply and

4
extend the B-SPM method [7], [28] to representative aerospace systems and illustrate the

significant improvement in system performance under uncertainty and time delay. In all

chapters, relevant real-world aerospace systems and simulations are employed to

illustrate and validate the theoretical advances derived.

The thesis is organized as follows. Chapter 1 derives baseline model following

control using dynamic inversion, with application to aircraft longitudinal and directional

dynamics. In Chapter 2, we derive state limiting modification to the model following

dynamic inversion controller formed in Chapter 1. The efficacy of the state limiting

component is illustrated via high-fidelity piloted simulation and subsequent flight testing

on the X-48B Blended-Wing-Body (BWB) vehicle. In Chapter 3 we derive adaptive

augmentation and apply the results to flight dynamics corresponding to the longitudinal

axis of a fixed-wing aircraft. We present and discuss simulation data to substantiate our

claims that the adaptive augmentation is able to restore system tracking performance in

the presence of matched uncertainties.

Chapter 4 combines the advances developed in the first three chapters to form a

theoretically justified control design consisting of robust dynamic inversion baseline

control, adaptive augmentation to mitigate matched uncertainties, and state limiting

augmentation to keep the system within a prescribed region in the system state space. We

present simulation results using representative aerospace data to show the efficacy of the

design in keeping an aircraft in its permissible flight envelope while restoring tracking

performance in the presence of matched uncertainties. In Chapter 5, we derive robust

LQR-optimal baseline control law and apply classical MRAC-based architecture

5
extended to explicitly address time-delay at the system input. We then apply and extend

the B-SPM adaptive approach to the same problem and show significantly improved

performance in the presence of input time-delay while restoring tracking performance in

the presence of matched uncertainties. Finally, Chapter 6 presents topics for further

investigation.

6
Chapter 1: Model Following Using Dynamic Inversion

This chapter considers a class of model following control systems designed to provide

tracking performance. Given second-order dynamical systems in cascaded form, we

design a baseline model-reference dynamic inversion based tracking control law in both

explicit and implicit forms. Interest in considering this particular class of systems stems

from flight control related applications, where inner-loop controllers for fixed-wing

aircraft are often designed based on simplified models [1]-[4]. These models represent

the decoupled fast responses in pitch, roll and yaw axes and are given in cascaded form.

The dynamic inversion approach is particularly useful in flight control applications

where the designer may explicitly include vehicle handling qualities requirements in the

selection of reference model desired damping and frequency to meet handling qualities

requirements. After deriving dynamic inversion tracking control for the general case, we

include specific design for aircraft longitudinal and directional axes. The results of the

chapter contribute robust baseline flight control architecture in which both aircraft

performance and handling qualities requirements may be achieved in a convenient

manner. The dynamic inversion design presented will serve as robust baseline control

architecture which to be further improved in the following chapters via adaptive

augmentation and state limiting augmentation components.

7
Chapter 1.1: Baseline Dynamic Inversion Design

This section formulates the system dynamics, poses the control problem, and derives

baseline Dynamic Inversion (DI) Proportional + Integral + Derivative (PID) control

architecture for flight dynamics.

Consider 2nd order dynamical systems in the cascaded form:

 x1  F1  x1 , z   B1 x2
 0

 (1.1)
 x2  F2  x1 , x2 , z   x2
0 cmd

where x   x1 x2  is the system state vector, z is the known bounded external signal,
T

 F , F  are known state-dependent functions, B


0
1
0
2
1 is a known nonzero constant, and x2cmd

is the system control input. The control goal is to design the control input x2cmd so that the

system 1st state component x1 tracks any given bounded time-varying command x1cmd  t  ,

while keeping all the signals in the closed-loop system bounded, uniformly in time.

This approach will employ a model reference based design framework. The reference

model is driven by a bounded possibly time-varying reference command x1cmd , and is

chosen to be 2nd order, with desired damping ratio  and natural frequency  :

 2  cmd
x  2
m
2 1
x  x1m  2   x1m   2  x1cmd  x1m  (1.2)
 s  2 s   
1

We note that in aerospace applications the desired damping ratio and natural frequency

are selected by the designer to achieve performance and, if applicable, handling qualities

for piloted aircraft.

Differentiating the first state component in (1.2) yields:

8
F10 F10
x1  x1  z  B1 x2 (1.3)
x1 z

Substituting the first equation in (1.1) into (1.3) results in:

 F 0 F 0 
x1   1  F10  B1 x2   1 z  B1 F20   B1 x2cmd (1.4)
 x1 z 
f  x1 , x2 , z , z 

or, equivalently:

x1  f  x1 , x2 , z, z   B1 x2cmd (1.5)

One may immediately note that in (1.5) the function f  x1 , x2 , z, z  and the constant

B1 are known. Dynamic Inversion (DI) based Proportional + Integral + Derivative (PID)

control is introduced in the form:

  x  xm  
 
x2cmd  B11  x1m  f  K D xˆ1  x1m  K P  x1  x1m   K I  1 1  
 s 
(1.6)

Because system state derivative is often not available as a measurement, the quantity:

xˆ1  F10  x1 , z   B1 x2 , xˆ1  0   x1  0  (1.7)

in (1.6) is estimated / predicted 1st state derivative. Let e1  x1  x1m be the reference

model tracking error signal. Substituting (1.6) into (1.5) results in the closed-loop

tracking error dynamics:

e1
e1  K D e1  K P e1  K I 0 (1.8)
s

Note that the error dynamics characteristic polynomial is given by:

 3  KD  2  KP   KI  0 (1.9)

9
As a result, exponential stability of the error dynamics in (1.8) can be achieved with

any appropriate PID gains selected via Routh tables or other methods. If we define

desired error characteristic polynomial as:

 2
 2     2    k   0 (1.10)

where  ,   are the desired damping ratio and natural frequency while k  0 is a fixed

positive constant, we may expand (1.10) and compare to (1.9) to yield PID feedback

gains:

KD  2   k

KP  2  k  
2
(1.11)

KI   k
2

Using PID gains (1.11), closed-loop tracking error dynamics (1.8) become globally

exponentially stable. This implies that system state component x1  t  tracks reference

signal x1m  t  exponentially fast.

Under control (1.6), the closed-loop dynamics of the second state component in (1.1)

become:

x2  F20  x1 , x2 , z   x2cmd
 e 
 F20  x1 , x2 , z   B11  x1m  f  K D e1  K P e1  K I 1  (1.12)
 s
 F 0
F 0
e 
 B11  x1m  1  F10  B1 x2   1 z  K D e1  K P e1  K I 1 
 x1 z s

Assuming zero initial conditions, error dynamics (1.8) and the PID gains (1.11) yields:

s 2
 2   s   2   s  k  x1   s  k   s 2  2   s   2  x1m (1.13)
 2 x1cmd

10
Hence, the closed-loop transfer function from the model reference external command

x1cmd to the system state x1 has the desired 2nd order form:

x1 2
 (1.14)
x1cmd s 2  2   s   2

In order to show that the control in (1.6) solves the tracking problem, note that since

the selected PID gains yield asymptotically stable error dynamics, the error signal e1  t 

and its time derivative e1  t  are bounded. At the same time, the model reference signal

x1m  t  and its time derivative x1m  t  are bounded by design. Consequently, the system

state components x1  t  and x1  t  are bounded. Using the first equation in system

dynamics (1.1) implies that x2 is bounded, and therefore together with (1.8) the tracking

problem is solved in the sense that all signals are bounded and the tracking error

converges to zero exponentially fast.

Chapter 1.2: Aircraft Longitudinal Design

Assuming small angle-of-sideslip (AOS)    0 , negligible lift due to elevator

 L  0  , and negligible thrust effects, fixed-wing aircraft short-period dynamics can be

written in the form of (1.1) [1], [4]:

   L   Qgrav  q

 (1.15)
q  M    M q q  M IC  qcmd

11
where  is aircraft angle-of-attack (AOA), q is the angular pitch rate, L is the known

lift curve slope, Qgrav is the known gravity term, M  is the known static stability

(pitching moment), M q is the known constant pitch damping, M IC is the known pitching

moment increment due to inertial cross-coupling effects, and qcmd is the commanded

pitch acceleration (control input).

The AOA reference model dynamics is chosen in the form of (1.2):

 2 
m   2    m  2   m   2  cmd   m 
2  cmd
(1.16)
 s  2 s   

To perform the design, one needs to match the required components against the

corresponding AOA/pitch rate dynamics terms. Comparing (1.15) against generic

cascaded dynamics (1.1) yields:

x1   , x2  q, z   Qgrav M IC  , x2cmd  qcmd


T

(1.17)
F1   L   Qgrav , B1  1, F2  M    M q q  M IC

In this case, the baseline PID feedback gains are:

K K D  2    k1
K K P     2  k1  (1.18)
KI K I   2 k1

where the integrator pole k1 is chosen as:

k1  L (1.19)

Based on the AOA equation in (1.15) and the reference dynamics (1.16), define

reference model pitch rate signal:

12
qm  m  L m  Qgrav (1.20)

Substituting (1.17)-(1.20) into DI PID control (1.6) yields explicit model following pitch

acceleration command:

   qm  q  
qcmd  qm  M    M q q  M IC  2     qm  q  (1.21)
 2 s 

Alternatively, if it is desired to generate control (1.6) without explicit implementation

of reference model dynamics (1.16), an implicit model following pitch acceleration

command may be formed via (1.21), (1.20), and (1.16) as:

qcmd   M    M q q  M IC  2   

  s  L   
    q  Qgrav   Qgrav (1.22)
 2 
cmd
s 

We verify that control input (1.22) results in desired closed-loop dynamics for system

(1.15) as follows. Differentiating the first state equation in (1.15) and substituting the

second state equation yields:

   L   Qgrav  M   M q q  M IC  qcmd (1.23)

Employing control signal (1.22) in (1.23) gives closed-loop dynamics:

  s  L 
s  s  L    2     cmd     q  Qgrav  (1.24)
 2 s 

Noting from the first state equation in system dynamics (1.15) that:

q   s  L    Qgrav (1.25)

and substituting into (1.24) yields:

   cmd   
s  s  L     s  L  2      (1.26)
 2 s 

13
Cancelling terms and reducing (1.26) gives:

2
s   2       2  cmd (1.27)
s

Finally, differentiating (1.27) produces the desired 2nd order closed-loop response for the

vehicle angle-of-attack dynamics:

 2
 2 (1.28)
 cmd s  2   s   2

Based on system dynamics (1.15), the resulting closed-loop response for the pitch rate

dynamics to an angle-of-attack command is:

q  2  s  L 
 2 (1.29)
 cmd s  2  s   2

Note that in this case, if L  0 then the pitch rate dynamics would be non-minimum

phase, but the pitch rate dynamics would remain bounded in time.

Chapter 1.3: Aircraft Directional Design

Assuming small angles, fixed-wing aircraft AOS / yaw rate dynamics in stability axis

can be written in the form of (1.1) [1], [4]:


   Y   Rgrav  r
 (1.30)
r  N    N r r  N p p  N IC  rcmd

where  is aircraft AOS, r is the angular yaw rate, p is angular roll rate, Y is the

sideforce slope due to AOS, Rgrav is the known yaw rate component due to gravity,

14
N  , N r , N p , N IC  represent dimensional yawing moment terms, and rcmd is the

commanded yaw acceleration (control input).

The AOS reference model dynamics is chosen in the form of (1.2):

 2 
m   2    m  2    m   2   cmd   m 
2  cmd
(1.31)
 s  2 s   

In this case, comparing (1.30) against generic cascaded dynamics (1.1) yields:

x1   , x2  r , z   Rgrav N IC  , x2cmd  rcmd


T

(1.32)
F1  Y   Rgrav , B1  1, F2  N    N r r  N p p  N IC

Similar to AOA case, the baseline PID feedback gains are:

K K D  2    k1
K K P     2  k1  (1.33)
K I K I   2 k1

where the integrator pole k1 is chosen as:

k1  Y (1.34)

Based on the AOS equation in (1.30) and the reference dynamics (1.31), define

reference model yaw rate signal:

rm Y m  Rgrav  m (1.35)

Substituting (1.32)-(1.35) into DI PID control (1.6) yields explicit model following

yaw acceleration command:

   rm  r  
rcmd  rm  N    N p p  N r r  N IC  2     rm  r  (1.36)
 2 s 

15
Alternatively, if it is desired to generate control (1.6) without explicit implementation of

reference model dynamics (1.31), an implicit model following yaw acceleration

command may be formed via (1.36), (1.35), and (1.31) as:

rcmd   N    N p p  N r r  N IC  2   

   s  Y   
    r  Rgrav   Rgrav (1.37)
 2 
cmd
s 

Similar to pitch dynamics, we verify that control input (1.37) results in desired

closed-loop dynamics for system (1.30) as follows. Differentiating the first state equation

in (1.30) and substituting the second state equation yields:

  Y   Rgrav  N    Nr r  N p p  N IC  rcmd (1.38)

Employing control signal (1.37) in (1.38) gives closed-loop dynamics:

    s  Y  
s  s  Y    2     cmd     r  Rgrav  (1.39)
 2  s 

Noting from the first state equation in system dynamics (1.30) that:

r   s  Y    Rgrav (1.40)

and substituting into (1.39) yields:

  cmd   
s  s  Y     s  Y  2     (1.41)
 2 s 

Cancelling terms and reducing (1.41) gives:

2
s   2       2  cmd (1.42)
s

Finally, differentiating (1.42) produces the desired 2nd order closed-loop response for the

vehicle angle-of-sideslip dynamics:

16
 2
 2 (1.43)
cmd s  2   s   2

Based on system dynamics (1.30), the resulting closed-loop response for the yaw rate

dynamics to an angle-of-sideslip command is:

r  2  s  Y 
 (1.44)
cmd s2  2  s   2

Note that in this case, if Y  0 then the yaw rate dynamics would be non-minimum

phase, but the yaw rate dynamics would remain bounded in time.

17
Chapter 2: Model Following Using Dynamic Inversion

Controller with State Limiting

While the controller developed in Chapter 1 was shown to solve the tracking problem,

in practical application there are additional requirements that must be considered during

the design process. Vehicles with control augmentation systems rely on state limiters to

prevent the exceeding of predetermined state limits. These limiters typically activate once

predefined envelope boundaries are exceeded and introduce sharp changes to control

characteristics once active. The resulting nonlinearities make analysis using conventional

control system methods difficult or invalid. Current state-of-the-art in limiter design

consists of ad-hoc limiting schemes developed specifically for the system under

consideration. As a result, such methods require extensive design and test iterations

specific to each individual application. Often, these designs do not provide stability /

performance guarantees.

An aircraft flight controller must keep the vehicle in a pre-specified region of the

corresponding state. This region is referred to as the operational flight envelope. For

example, an AOA protection system (often called AOA Limiter) is typically employed to

maintain the aircraft AOA within an allowable range outside of which loss of control is

expected. The control challenge in such cases is to blend two subsystems, the tracking

controller and the limiter, with seamless transition between the two controllers while

preserving closed-loop stability at all times. We solve this problem by introducing state

limiting augmentation to the baseline dynamic inversion controllers introduced in

Chapter 1.

18
The proposed method introduces a limiting technique which gradually modifies the

expected behavior of the system dynamics when user-specified criteria are approached /

exceeded. Given a baseline model following dynamic inversion control law designed to

meet robustness and performance requirements throughout a flight envelope, a state-

limiting augmentation component protects the system trajectories from leaving an

allowable subset in the system state space. The proposed design is applied to the X-48B

blended-wing-body (BWB) aircraft in both the longitudinal and lateral/directional axes

with both piloted simulation and flight test results included. The results of this chapter

contribute theoretically justified state limiting to a robust baseline control design,

allowing for aircraft performance as well as protecting the system from exiting a pre-

specified region of the state space. The limiter design will be further improved in

subsequent chapters by including adaptive augmentation to mitigate matched

uncertainties.

19
Chapter 2.1: State Limiter – Motivating Example

In order to illustrate the main idea of the design, first consider scalar dynamics:

xu (2.1)

where x is the system state and u is the control input. The control goal is once again to

choose u such that x tracks the state of a user-defined reference model xm whose

dynamics are:

xm  2   xm   2  xcmd  xm  (2.2)

where  and  represent the desired damping ratio and natural frequency, respectively.

Note that we are dealing with a simplified version of the problem presented in Chapter 1.

Furthermore assume that it is required that the system state x1 does not exceed pre-

specified limits:

 xlim  x1  t   xlim (2.3)

In order to ensure feasibility of a control solution, we assume that the commanded

signal xcmd satisfies the same limits, and for clarity we assume that the state derivative x

is available on-line. For simplicity, limits (2.3) are defined as symmetric but we note that,

without loss of generality, this need not be the case. Choosing control solution as:

u  2   x   2  xcmd  x  (2.4)

yields closed-loop system:

x  2   x   2  xcmd  x  (2.5)

Let

20
e  x  xm (2.6)

denote the tracking error. Subtracting the reference dynamics from the system dynamics

gives:

e  2   e   2 e (2.7)

This relation implies global exponential stability of the origin. Consequently, the

system state tracks the state of the reference model exponentially fast starting from any

initial conditions. However, the desired state limits (2.3) are not guaranteed.

We propose a formal design modification to enforce the state limits. The main idea is

to gradually change the reference model if the system state approaches any of its limits.

The modified reference dynamics in turn alter the control input to the system in such

regions. Resulting control commands maximize vehicle state-limited performance within

existing control law architecture. Specifically in this case, the proposed state limiter logic

introduces state-dependent damping ratio and natural frequency. Consider modified

damping ratio and natural frequency:

 x  x       x  
(2.8)
x  x       x  

In (2.8),   x  is the modulation function described in [20] and based on work in

[21], and x is the aggregate state vector containing the degrees of freedom of the system.

This is a continuous state-dependent map which allows smooth transition of the desired

damping ratio and natural frequency. The modulation function allows the designer to

define the region sufficiently close to the state limits for the modified reference model

parameters (2.8) to become active. This ensures that for all nominal system conditions

21
(away from state limits), the desired baseline control system behavior and system

dynamics are achieved. Construction of the modulation function is presented below.

Suppose that x0 is the center point of the sphere R   x  x0  R . Also suppose

that the set  R represents the allowable state domain. Let   0 be a small positive

constant and define R    x  x0  R    . The multidimensional modulation function

is defined as:

0, x   R 

  x   0    x   1, x   R   R  (2.9)
1, x  R

Formally, the modulation function can be written as:

  x  x0  xlim  
  x   max 0, min 1,1   (2.10)
    

where xlim is a vector whose components represent the limit value for each system degree

of freedom, and the limiter remains inactive in nominal conditions until system state

approaches defined limits within user-defined tolerance  . The limiter is then gradually

applied as the system state approaches the state limits, avoiding undesirable behavior.

Graphical representation of the modulation function is shown in Figure 1.

  x
1

x  x0
0 xlim   xlim

Figure 1: Limiter Onset Modulation Function

22
Including modification (2.8) results in modified reference dynamics:

xm  2  x  x  x  x  xm  x2  x   xcmd  xm  (2.11)

Note that in (2.8) the quantities  and  are user-defined incremental damping

ratio and natural frequency tables, respectively. These two tables are chosen to meet the

desired tracking performance and system robustness requirements. For example, the

reference model can be modified such that its state has no overshoot while tracking a

step-input command. Hence, if the latter is within the limits then the system state and the

reference model state will remain within the desired limits as well.

To illustrate the limiter concept, consider system (2.1) and reference model (2.2) with

baseline parameters   0.3 and   4 rad/sec. Obviously, in this case the baseline step

response will contain significant overshoot since the system is underdamped. Select

limiter parameters:

   4.5,    4.5,   0.1 (2.12)

Parameters (2.12) indicate that once the system response comes within 10% of the

defined state limit, the damping and natural frequency will increase and reach values of

  4.8 and   8.5 rad/sec once the limiter is fully active, i.e. the system reaches the

limit. Consider a step command with magnitude 2, where the state limit for system state

x1  2 . The step response for the baseline and state limited systems is shown in Figure 2.

The response shows that the limiter is able to arrest the system response and keep the

state response within the specified limit. The modulation parameter   x  is inactive until

the state response enters the modulation region defined by  , at which point it

23
increments from zero to one as the state approaches the limit. The modified system

damping ratio and frequency stay at their respective baseline design values until the

system near the state limit, at which point both evolve to their modified values defined by

increments (2.12).

xcmd
4
limit area
x (no limit)
x

2
x (limit)
0
0 0.5 1 1.5 2 2.5 3

1
(x)

0.5

0
0 0.5 1 1.5 2 2.5 3

5
x

0
0 0.5 1 1.5 2 2.5 3

10
x

0
0 0.5 1 1.5 2 2.5 3

time, sec

Figure 2: Step Response Comparison – Nominal vs. Limited System

24
In order to investigate the efficacy of the limiter design, we must consider both the

closed-loop stability of the system as well as determine whether the limit design truly

guarantees that the system state remains within the imposed limit for all time. To simplify

the analysis we define the system in state space form:

x1  x2
(2.13)
x2  u

and apply control:

u  2  x  x  x  x  x2  x2  x   xcmd  x1  (2.14)

yielding closed-loop dynamics:

 x1   0 1   x1   0 
 x     2  x  2   x    x    x    2  x  xcmd (2.15)
 2  x x x  2  x 

Similarly we may define reference dynamics (2.11) in state space form:

 xm1   0 1   xm1   0 
 x     2  x  2   x    x    x    2  x  xcmd (2.16)
 m2   x x x   m2   x 

Based on relations (2.15) and (2.16) with error signal definition (2.6) we can now

derive the closed-loop error dynamics in state space form:

e   0 1  e 

e    2  x  2   x    x   e  (2.17)
   x x x  
E A E

alternately expressed as:

E  A x E (2.18)

25
It is sufficient to analyze system (2.17) to show closed-loop stability of the system.

Based on definition (2.8) and limiter modulation function (2.10) we find three regions of

interest:

Case 1: x  x0  xlim   . In this region, the modulation function is inactive, and the

error dynamics may be expressed in terms of the nominal damping  and natural

frequency  . Error dynamics (2.18) are no longer a function of x and instead satisfy

autonomous linear time invariant (LTI) relationship:

 0 1 
E 2
2   
E (2.19)
 

The closed-loop poles of (2.19) are given by:

1 , 2       2  1 (2.20)

Because the nominal damping  and natural frequency  are selected by the designer,

they may be chosen to ensure that error dynamics (2.19) are exponentially stable.

Case 2: x  x0  xlim . In this region, the modulation function is completely active,

and the error dynamics may be expressed in terms of the full incremented damping

    and natural frequency     . Error dynamics (2.18) are no longer a function of

x and instead satisfy autonomous linear time invariant (LTI) relationship:

 0 1 
E E (2.21)
       2          
2

The closed-loop poles of (2.21) are:

1 , 2                       1
2
(2.22)

26
Once again, since the nominal damping and natural frequency is selected by the designer

and so are the incremental damping and natural frequency tables, they may be chosen to

ensure that the error dynamics (2.21) are exponentially stable. We note that although the

design goal of the limiter is to ensure that the system dynamics never enter a region

where x  x0  xlim , in practical application external disturbances may drive the system

into this region. One example of this behavior is a gust acting on an aircraft. This analysis

indicates that even should a system exceed its limit, the error dynamics remain

exponentially stable.

Case 3: xlim    x  x0  xlim . In this region, the modulation function is active and

likely changing due to the evolution of the system states. The error dynamics become:

 0 1 
E 2 E (2.23)
 x  x  2  x  x  x  x  

We immediately note that, as shown in the analyses of cases 1 and 2, the designer may

choose the nominal damping and natural frequency as well as the incremental tables for

each parameter, so that the error dynamics (2.23) could be made exponentially stable for

any value of the modulation parameter  . However, since dynamics (2.23) now represent

a time-varying system, one cannot determine the closed-loop stability by simply

inspecting its eigenvalues. In addition, since dynamics (2.23) are not autonomous as

posed, stability results such as LaSalle‟s Theorem do not apply and we must use an

alternate approach to show stability.

We demonstrate three stability results for dynamics (2.23). Since each of the results is

a sufficient condition for stability, they may be conservative in some cases.

27
Method 1: Vanishing Perturbation

Re-write error dynamics (2.23):

E  A0 E  g  t , E  (2.24)

where A0 represents Hurwitz LTI component of error dynamics, while g  t , E 

represents time varying component due to the activity of the modulation function and

satisfies g  t , 0   0 . Since the modulation function operates between the values of zero

and one, we know that g  t , E  2   E 2


for all t  0 and all E  n
where  is a

nonnegative constant. Let Q  QT  0 and solve the Lyapunov equation:

P A0  A0T P  Q (2.25)

for P . The quadratic Lyapunov function V  E   ET P E will then satisfy:

min  P  E 2  V  E   max  P  E
2 2
2

V
A0 E   E T Q E  min  Q  E 2
2
(2.26)
E
V
 2 E T P  2 P 2 E 2  2 max  P  E
E 2 2 2

The derivative of V  E  along the trajectories of the perturbed system (2.24) satisfies:

V  E   min  Q  x 2  2 max  P   x
2 2
2
(2.27)

Hence, the origin is globally exponentially stable if   min  Q  / 2 max  P  . This ratio is

maximized with the choice Q  I , and therefore the origin is globally exponentially

stable if:

28
1
 (2.28)
2 max  P 

We note that (2.28) places a bound on the size of the perturbation permissible while

retaining stability. In this case, it is a restriction on the amount of change permitted in the

damping and frequency from nominal values. The result is likely to be extremely

conservative because the vanishing perturbation analysis is a worst case analysis.

Method 2: Input-to-State Stability

Consider error dynamics (2.23) as:

E  f  t , E, u  (2.29)

where input u  t   g  t , E  is a piecewise continuous, bounded function of t for all

t  0 . If we again model the system as in (2.24), we see that the unforced system:

E  A0 E (2.30)

has a globally uniformly asymptotically stable equilibrium point at the origin. Due to the

modulation function, we may model the external bounded input as:

g  t , E     t   A1  A0  E (2.31)

where A1 corresponds to the fully active modulation system (2.21) and A0 corresponds to

the nominal system (2.19). We can write the system dynamics solution as:

E t   e E  t0    e g  , E    d
A0  t t0  t A0  t  
(2.32)
t0

and use the bound:

A0  t t0     t t0 
e ke (2.33)

29
to estimate the solution by:

E t   k e E  t0    k e g  , E    d
   t t0  t    t  
(2.34)
t0

Using relation (2.31), we may reduce (2.34):

E t   k e E  t0    k e   t   g  , E    d
   t t0  t

t0
(2.35)
E  t0    k e    A1  A0 E   d
   t t0  t    t  
ke
t0

Noting the modulation parameter satisfies 0    t   1 and that quantity A1  A0 now

represents a constant as written in (2.35), we may use the Bellman-Gromwell Lemma to

bound error dynamics as:

E  t   k E  t0  e 
  k A1  A0  t
(2.36)

Therefore, for stability we require that:


A1  A0  (2.37)
k

Relationship (2.37) gives a bound on the maximum singular value of the matrix  A1  A0 

in relation to the slowest eigenvalue of the unperturbed system and corresponding

eigenvectors. Again, this result will be conservative but provides another sufficient

stability result.

Method 3: Second Method of Lyapunov

Rewrite error dynamics (2.23) as an unforced second-order differential equation as a

function of time:

E  f t  E  g t  E  0 (2.38)

where:

30
f  t   2  x  t  x  t 
(2.39)
g  t   x2  t 

Because of the relation (2.8), it is apparent that:

0  f min  f  t   f max
(2.40)
0  g min  g  t   g max

From dynamics (2.38) or (2.23), we have:

e1  e2
(2.41)
e2   g  t  e1  f  t  e2

Following the analysis presented in [22], define Lyapunov function:

1 2
V  t , E   e12  e2 (2.42)
g t 

Differentiating (2.42) along the system trajectories yields:

2 e2 e2 d  1 
V  t , E   2 e1 e1   e22  
g t  dt  g  t  
 g t   2 f t  g t   2 (2.43)
   e2
g t 
2
 
K t 

Using results from [22], if:

g t   2 f t  g t 
 (2.44)
g t 
2

and f  t   f max   the solution is asymptotically stable.

31
Using (2.44), we may now investigate what the requirement of K  t   0 indicates for

error dynamics with our specific structure (2.39). Using (2.44) while letting  1, for

stability we require:

g  t   2 f  t  g  t  (2.45)

Substituting structure (2.39) into (2.45) yields stability requirement:

x  t   2  x  t  x2  t  (2.46)

We immediately note that in the case where damping and frequency are constant and

satisfy bounds (2.40) that (2.46) is always satisfied, as should be expected. Rearranging

(2.46) yields:

x  t 
 2  x  t  (2.47)
x2  t 

Integrating both sides of (2.47):

t
x   t

0 x2   d  2 0  x   d (2.48)

Perform change of variables:

y      dy     d
(2.49)
y  0    0 , y t    t 

Using new variable of integration, (2.48) becomes:

 t  t
dy
 2
 2     d (2.50)
  0 y 0

Performing integration on the left hand side of (2.50) yields:

32
 t 
1
t
 2     d
y    0
(2.51)
0

Rearranging (2.51) yields condition for stability:

  0
 t   t
(2.52)
1  2   0      d
0

We now note that it is not necessary to perform the integration in (2.52), because due

to conditions (2.39) and (2.40), we have that:

t
1  2   0      d  1, t  0 (2.53)
0

As a result, stability requirement (2.52) can always be satisfied by using (2.40) and

selecting:

  0   min  gmin (2.54)

We must now show that the state limiter solution guarantees that the system does not

exceed its pre-specified state limit in addition to providing stable tracking response. First,

we return to Lyapunov function (2.42) which as shown via (2.54) confirms the stability of

error dynamics (2.38) in the region where the modulation function is active and time-

varying. Since Lyapunov analysis only provides sufficient conditions for stability, we

begin by ensuring that Lyapunov function (2.42) is also valid in the cases when the

modulation function is inactive. In these cases, the limiter is either completely active or

inactive, and stability has already been established by virtue of the fact that all

eigenvalues have negative real components. In the time-invariant case, we have:

33
1 2
V  E   e12  e2 (2.55)
g

In (2.55) we have either g   2 or g       based on (2.19) and (2.21), but in


2

either case g is strictly positive. Differentiating (2.55) along the system trajectories

yields:

2 e2 e2
V  E   2 e1 e1 
g
(2.56)
2f 2
 e2  0
g

Now, using LaSalle‟s method, from (2.56) we have:

2 f 2
S  e, e : V  E   0  e2  0  e2  0 (2.57)
g

Substituting e2  0 into error dynamics (2.41) yields:

e1  0  e1  t   e1  0  (2.58)

Additionally,

e2  0  e2  0   g e1  0  e1  0 (2.59)

This indicates that the origin is the only point in the set S where V  0 , which indicates

that the error dynamics are asymptotically stable. Therefore, Lyapunov function (2.55) is

sufficient for use in cases when the modulation function is not varying.

Using error definition (2.6), suppose that system dynamics satisfy:

e k (2.60)

for some k  0 . In terms of system state, we may rewrite (2.60) as:

34
x  k  xmmax (2.61)

where xmmax is user-defined maximum permissible value of the reference model position

state, which must be less than the absolute system position limit xlim . Redefine:

k   xmmax (2.62)

for some   0 . Substituting (2.62) into (2.61) yields:

x  1    xmmax (2.63)
xlim

Therefore, if error dynamics satisfy (2.60) system position will stay within the limit as

shown in (2.63). Solving (2.63) for  , we get:

xlim  xmmax
 (2.64)
xmmax

Substituting (2.64) into (2.62) yields:

k  xlim  xmmax (2.65)

Returning to Lyapunov function (2.42) along with error constraint (2.60), we have:

1
V  e, e   k 2 
2
emax (2.66)
g t 
 k , g  t , emax 

where emax is the maximum error rate achievable/allowable for specific system dynamics.

Lyapunov function (2.66) will form an elliptical level set with the specific shape of the

ellipse defined by user-selected value for emax . Because of results proved in [22], the

system trajectories will always be negative definite and therefore dynamics will evolve

35
toward the origin. In other words, once system dynamics begin in a Lyapunov level set

the dynamics must evolve in the negative (shrinking) direction. Therefore, we may

formally state:

e  0   k  V  e, e     k , g  t  , emax   x  t   xlim , t  0 (2.67)

Recalling that xmmax is the user-defined maximum permissible value of the reference

model, we see that (2.65) provides a tuning knob by which the designer may define the

maximum permissible error for which the system dynamics are guaranteed to remain

within system limits.

Chapter 2.2: State Limiter Application to Aircraft Dynamics

In order to apply the limiter augmentation to the aircraft pitch dynamics described in

Chapter 1.2, we note the implicit pitch acceleration command (1.22) may alternatively be

expressed as:

 L 
qcmd   M    M q q  M IC   2  cmd       cmd     
 s  (2.68)
2    q  Qgrav   Qgrav

Including limiter modification (2.8) to (2.68) yields:

2  
qcmd   M    M q q  M IC      x     cmd       cmd     
L
 s  (2.69)
2     x        x     q  Qgrav   Qgrav

Pitch acceleration command (2.69) now takes the form:

qcmd  qcmd
bl
   x   qcmd (2.70)

36
bl
where qcmd is baseline pitch acceleration command (2.68), and

2  L 
 qcmd   2         cmd       cmd     
 s  (2.71)
2                q  Qgrav 

is the state limiting augmentation component.

Similarly, returning to the aircraft yaw dynamics described in Chapter 1.3, we note

the implicit yaw acceleration command (1.37) may be formed without explicit integral as:

 
rcmd   N    N p p  N r r  N IC  2        cmd      Y   Rgrav (2.72)
 2 

Or, alternatively:

 
rcmd   N    N p p  Nr r  N IC   2    cmd   2    Y   Rgrav (2.73)

Including limiter modification (2.8) to (2.73) yields:

rcmd   N    N p p  N r r  N IC      x         cmd  
2

(2.74)
 2     x         x      Y    R
 grav

Yaw acceleration command (2.74) now takes the form:

rcmd  rcmd
bl
   x   rcmd (2.75)

bl
where rcmd is baseline yaw acceleration command (2.73), and

 rcmd   2           cmd   2              
2
(2.76)

is the state limiting augmentation component.

We immediately note from pitch acceleration command (2.71), yaw acceleration

command (2.75), and construction of the modulation function (2.9) that when the system

37
is in nominal (away from state limits) region the baseline tracking control law provides

the entire command. As the system approaches state limits within user-defined tolerance

 , the state limiting augmentation component gradually activates and modifies the

reference model dynamics. This in turn modifies the closed-loop system dynamics, and

based on user-defined incremental damping ratio and natural frequencies may be used to

modify system behavior near state limits. For example, the user may define the

incremental damping ratio such that in regions near the state limits the closed-loop

behavior becomes overdamped, thus preventing overshoot of state limits.

Chapter 2.3: Application to X-48B Aircraft Simulation

The X-48B is an 8.5% scale version of a full-scale blended-wing-body (BWB)

aircraft designed to investigate the stability and control (S&C) characteristics of this

aircraft configuration. The flight test vehicle has successfully completed more than 75

flights at NASA Dryden since July 2007. Information on the vehicle and flight test results

may be found in [23]-[25]. As shown in [24], the latter portion of the first phase of flight

testing was high risk departure limiter assaults intended to investigate the ability of the

aircraft to prevent entry into uncontrolled flight regimes. The AOA and AOS limiters

presented in previous sections were specifically designed to address assault maneuvers

where representative reference model dynamics would experience overshoot into

undesirable regions outside of the flight envelope. We include simulation results in this

chapter but note that the results presented were subsequently taken to flight testing, where

a total of four successful limiter assault flights took place. Details of the flight test

38
approach along with presentation and discussion of the results may be found in the

following section and in more detail in [26].

The baseline longitudinal control system for the X-48B is defined most simply by

(2.68), with some additional components such as thrust and control allocation effects

neglected for the purposes of this work. The X-48B baseline control law originally

contained an AOA limiter that was satisfactory for large portions of the flight regime but

was found to insufficiently arrest AOA during aggressive assault maneuvers. As a result,

limiter augmentation component was included and pitch acceleration command modified

to (2.70). High-fidelity nonlinear piloted simulation results for slats retracted windup turn

maneuver comparing the original limiter with the proposed limiter augmentation are

shown in Figure 3.

39
AOA Response, Wind-Up Turn

 (deg)

Command
Baseline + Aug.
Baseline

Time (sec)

Figure 3: AOA Limiter Performance in Windup Turn

The results show that the limiter augmentation is able to arrest the AOA as the

response approaches the command, which in this case is equal to the limit. The original

limiter is not able to arrest the response and the vehicle would enter an undesirable flight

regime. It should be noted that there is some minor overshoot in the AOA response as the

40
windup turn maneuver is continued past the initial transient. In this specific case, this is

due to a modification of the limiter presented above. The portion of (2.71) due to the

AOA error is only active once the AOA limit is exceeded in implementation – this was

done to avoid the alpha rate decreasing prior to the limit being exceeded. The X-48B

program was willing to allow a small overshoot with graceful return to the limit in

exchange for improved pitch tracking (handling qualities) approaching the limit. We note

that the augmentation could provide strict adherence to the AOA limit if so desired, as

demonstrated in the lateral/directional axes.

The elevator surface position and rate responses to the wind-up turn maneuver are

shown in Figure 4. The elevator responses show that the addition of limiter augmentation

does not adversely affect the surface position or rate, both of which remain within

desirable ranges with smooth dynamics. Per design, the surface response remains dictated

completely by the baseline pitch acceleration command when the modulation function

  x  is not active (zero).

41
Elevator Activity
Baseline + Aug.
e
e dot Baseline

Baseline + Aug.
Baseline

Baseline + Aug.

Time (sec)

Figure 4: Elevator Response for AOA Windup Turn

Finally, the reference model parameters for the AOA maneuver are shown in Figure

5. Per design, the reference damping and natural frequency remain static when the

modulation function   x  is not active. As the modulation function activates, the

42
damping of the reference model increases, causing the closed-loop system to move

toward overdamped characteristics and preventing overshoot of the state limits.

AOA Augmentation Parameters


Baseline + Aug.

Baseline + Aug.

Baseline + Aug.

Time (sec)

Figure 5: Reference Model Parameters for AOA Windup Turn

In the lateral/directional axes, the baseline control system for the X-48B is defined by

(2.72). The original baseline control system did not include an AOS limiter, so the

43
augmentation scheme described in this paper was designed specifically to address the

limiter assault flight test maneuvers. Limiter augmentation component was included and

yaw acceleration command was modified to (2.75). High-fidelity nonlinear piloted

simulation results for slat extended level flight sideslip step input maneuver comparing

the baseline unlimited control law against the proposed limiter augmentation are shown

in Figure 6.

44
AOS Step Response

 (deg)

Command
Baseline + Aug.
Baseline

Time (sec)

Figure 6: AOS Limiter Performance, Step Input

The results show that in the lateral/directional axes, where the theory presented above

was implemented without modification, the limiter is able to fully arrest the closed-loop

dynamics and prevent the AOS from exceeding the limit – which is equal to the

command in this case. We again point out that the nominal AOS response remains intact

45
until such time that the AOS is sufficiently close to the limit to activate the limiter

augmentation. This allows the design to satisfy limiter requirements while still meeting

performance and handling qualities requirements. The rudder surface response is shown

in Figure 7.

Rudder Activity
r

Baseline + Aug.
Baseline

Baseline + Aug.
Baseline
r dot

Baseline + Aug.

Time (sec)

Figure 7: Rudder Response for AOA Step Command

46
Just as in the longitudinal axis, the rudder response remains smooth and within

desirable ranges for surface deflection and rate with the inclusion of the limiter

augmentation component. The acceleration command and surface response is entirely

dictated by the baseline command until the limiter augmentation component becomes

active, which is illustrated in the similar surface response until the modulation function

becomes nonzero. The reference parameters for the AOS step maneuver are shown in

Figure 8.

47
AOS Augmentation Parameters
Baseline + Aug.

Baseline + Aug.

Baseline + Aug.

Time (sec)

Figure 8: Reference Model Parameters for AOS Step Command

Once again, the reference model parameters remain constant until the modulation

function   x  becomes active, at which point the parameters change due to the limiter

augmentation. The damping ratio increases, corresponding to a smooth transition to an

overdamped system in order to prevent overshoot in the AOS response.

48
Chapter 2.4: Flight Test of the X-48B AOA and AOS Limiting System

The X-48B‟s Blended Wing Body (BWB) aircraft configuration, presented in Figure

9, represents a design departure from the conventional “tube and wing” shape of

traditional transport aircraft. This novel configuration offers the potential for

revolutionary improvement in performance and efficiency over current day airframe

configurations. A blended-wing configuration is characterized by an overall aircraft

design that provides minimal distinction between wings and fuselage and fuselage and

tail. It closely resembles a flying wing configuration, but concentrates more volume in

the center section of the aircraft than a traditional flying wing.

Figure 9: X-48B BWB Configuration

The X-48B Low Speed Vehicle (LSV) is an 8.5% scale version of a full scale blended

wing body aircraft designed to investigate stability and control characteristics of this

aircraft configuration. Early analysis of blended wing-body aerodynamic characteristics

49
identified the potential for sustained spins and nose-up „tumble‟ post-departure modes.

As a result, one of the goals of the X-48B flight test program was to demonstrate AOA

and sideslip limiters that would provide departure resistance and allow aggressive

maneuvering up to CLmax and a sideslip limit equivalent to a full-scale normal landing in a

35kt. crosswind. In the final phase of X-48B handling qualities tests the airplane was

taken to its limit of controlled flight. In the third and final phase, “departure limiter

assaults” were performed to challenge the ability of the aircraft‟s flight control system to

prevent entry into uncontrolled flight regimes and to validate the software algorithms

employed in the computerized flight control system to prevent such occurrence. Through

simulation analysis and flight test the full-envelope AOA and sideslip limits that would

meet program goals and prevent departures from controlled flight were established. The

positive-AOA range of these limits is presented in Figure 10.

The X-48B currently uses a modified model-following control scheme with dynamic

inversion control mixing. The longitudinal and directional reference models are standard

2nd order equivalent systems with frequency and damping selected for handling qualities

requirements and stability margins. Tuning for handling qualities in general provides

predictable initial response and tracking characteristics; however, if the desired damping

characteristics are less than critically damped (<1) will result in an overshoot beyond the

desired steady-state response. If this response, for example, was an aggressive pull to the

AOA limit (coincident with CLmax ) the ideal system would experience an overshoot of up

to 2-4 deg AOA for X-48 representative reference model dynamics. Similar performance

would be experienced in the directional axis in response to an aggressive pedal input.

50
Flight testing indicated that this would result in AOA excursions into a region

characterized by an abrupt uncontrollable right wing drop, an undesirable flight

characteristic during maneuvers requiring maximum performance and minimum altitude

loss. Sideslip excursions beyond the recommended limit were not intentionally tested.

 limit variation with 

Slat Extended Limit


 - deg

Slat Retracted Limit

 - deg

Figure 10: X-48B Positive AOA Envelope Limits

In order to reduce limit overshoots while maintaining excellent handling qualities

away from the limiter, the limiter developed in Section 2.2 was developed and applied to

the X-48B vehicle.

AOA limiter testing was conducted using the standard windup/wind-down turn

method [27]. A windup turn is a constant airspeed, increasing AOA (and thus normal

acceleration) turn at fixed power, where altitude is traded to maintain airspeed. Above

51
corner speed, the lowest airspeed at which the normal acceleration limit can be attained, a

windup turn will increase AOA until the normal acceleration limit is reached; below

corner speed the windup turn increases AOA and normal acceleration until the AOA limit

is reached. Once the limiting condition (AOA or normal acceleration) is reached, the

windup turn transitions to: a wind-down turn, a constant normal acceleration deceleration

to corner speed, and then a constant AOA deceleration below corner speed. Deceleration

rate, typically specified in knots per second, is modulated by increasing or decreasing the

rate of descent of the airplane. This is a high workload closed-loop test; however, test

tolerances are typically very broad, and usable test data can be obtained with wide

variations in airspeed during the windup portion and in deceleration rate during the wind-

down portion. A windup/wind-down turn progression is presented graphically in Figure

11.

NZ
AOA Limit

Nz Limit Vc

WUT
1 WDT
0 Airspeed

Figure 11: Windup/Wind-down Turn V-n Diagram

52
To illustrate the interpretation of X-48B flight test data a representative windup turn

(WUT) / Wind-down turn (WDT) maneuver with annotations is presented in Figure 12.

All WUT/WDT maneuvers presented in this section adhere to this example.

D
E
Nz

B
F
G A

Vc

Figure 12: Representative Windup/Wind-down Turn V-n Diagram

The same maneuver, presented in terms of airspeed and AOA with identical phases

annotated, is presented in Figure 13.

53
E
F

AOAcf
D C

B
A
Vc

Figure 13: Representative Windup/Wind-down Turn Airspeed/AOA Plot

The segments of the maneuver presented in Figure 12 and Figure 13 are as follows:

A. Initial Condition. This corresponds to the start of data collection and is typically just

after a roll-in to a descending turn is initiated but before AOA has increased

significantly above 1g AOA at the test conditions.

B. Windup segment. During this portion of the maneuver altitude is sacrificed to

maintain airspeed as AOA and normal acceleration are increased. The goal is a

near-constant airspeed, increasing normal acceleration turn up to the test normal

acceleration limit

C. Point C corresponds to the limit normal acceleration for the test. At this point the

maneuver is continued; however, airspeed is allowed to decrease so that AOA

will increase.

54
D. Constant Normal Acceleration Wind-down segment: During this portion of the

maneuver the descending turn is shallowed such that airspeed decreases. AOA is

increased to maintain normal acceleration. The goal is a near-constant normal

acceleration, increasing AOA, decreasing airspeed turn up to the test AOA limit.

E. “Corner Speed”. This corresponds to the point at which the airplane achieves the limit

normal acceleration at the limit AOA. Below this speed the airplane is AOA-

limited; above this speed, normal acceleration limited.

F. Constant AOA Wind-down segment: During this portion of the maneuver the

descending turn is continued with decreasing airspeed and normal acceleration,

while normal acceleration decreases.

The maneuver is intended to be flown as a seamless transition from initial condition

through the windup segment and transitioning to the wind-down segments until

termination at some proscribed minimum airspeed. Ideally once at a limiting condition

(segments C-G) the airplane can “ride the limiter” and transition through each maneuver

phase with no pilot compensation to avoid exceeding a limit. Limiter functionality was

evaluated during all phases, and included limiter performance in the presence of high

approach rates, changing limits while limited, and changing flight conditions while

limited. AOA limiter test results for slat retracted/aft CG and slate extended/forward CG

are presented in Figure 14 and Figure 15.

55
AOA-limited WDT

g-
lim
ite
d
W
DT
AOAcf

WUT

Vc

g-limited WDT

DT
dW
mite
A-li
O
Nz

A WUT

Vc

Figure 14: AOA Limiter Results, Slat Retracted/Aft CG

56
- l i mit ed WDT
g

DT
dW
ite
WUT

lim
Nz

A-
AO

Vc

AOA-limited WDT
g-
lim
ite
AOAcf

d
W
DT

WUT

Vc

Figure 15: AOA Limiter Results, Slat Extended/Fwd CG

57
The AOA limiter limited overshoots to less than 2 deg in all cases tested, and

generally less than approximately 1 deg for all configurations and conditions tested,

including up to 5 deg/sec AOA rates and 5 kt/sec deceleration rates. In the slat

extended/forward CG case approaching the limiter from above corner speed resulted in

an approximate 1 deg steady-state error below corner speed (i.e. the airplane settled

approximately 1 deg above the AOA limit rather than at the limit) with minimal

overshoot. Limiter approaches from below corner speed converged to the limit AOA

condition with minimal overshoot. Limiter performance was acceptable in all cases

evaluated, and the limiter demonstrated functionality in the presence of changing limits.

It should be noted that the X-48B g-limiter is actually an AOA limiter, with the AOA

limit estimated as that AOA that will result in the normal acceleration limit at the current

flight condition (also known as the „AOA-for-g limiter‟). The AOA-for-g function did

not account for lift differences with control surface deflection; consequently, the limiter

with trailing edge up elevon during a windup turn typically underestimates g for any

given AOA, resulting in less than the normal acceleration limit when on the g-limit

portion of the V-n diagram. G-limiter performance was noted to improve somewhat

during aft CG testing, when trailing edge up control deflections were reduced. This

phenomenon is characterized by the „g-limited WDT‟ lines in Figure 14 and Figure 15.

The sideslip limiter was tested by abrupt full pedal inputs from a stabilized flight

condition. Buildup was conducted by initially performing these inputs from steady-

heading sideslips, resulting in smaller pedal inputs and sideslip command changes,

eventually progressing to step inputs from zero sideslip, the worst-case condition. Inputs

58
were tested in both directions to establish symmetry of the limiting function. To illustrate

the interpretation of X-48B flight test data representative step pedal input maneuvers with

annotations are presented in Figure 16. All sideslip limiter test maneuvers presented in

this report adhere to this example.

 limit variation with 

Slat Ret Limit

C2

C1
A1
 - deg

B1
A2
B2

 - deg

Figure 16: Representative Sideslip Limiter Maneuvers

A. Initial Condition (Sideslip may be nonzero)

B. Sideslip Change due to Step Pedal Input, Near-Constant AOA and Airspeed.

C1 A1  C 2  A2 C1 A1  C 2 A2


C. Test Sideslip Limit and Maneuver Completion

In this example the maneuver with the smaller sideslip increment (A1-C1) was flown

prior to the larger sideslip input (A2-C2), thus building up in time rate of change of

59
sideslip. Sideslip limiter test results for slat retracted/aft CG and slat extended/aft CG are

presented in Figure 17 and Figure 18 respectively.

 limit variation with 

Slat Ret Limit

Step Input from Stabilized Sideslip


 - deg

Step Input from Stabilized Sideslip

 - deg

Figure 17: Sideslip Limiter Results, Slat Ret/Aft CG

60
 limit variation with 

Slat Ext Limit

 - deg Step Input from Stabilized Sideslip

 - deg

Figure 18: Sideslip Limiter Results, Slat Ext/Aft CG

Sideslip limiter performance was considered excellent at all conditions tested, with

typically less than 0.5 deg overshoot. The limiter also demonstrated excellent

compensation during sideslip limit changes as AOA varied (Figure 18).

The AOA and sideslip limiter system developed for the X-48B demonstrated

acceptable performance at all conditions tested, including slat extended and retracted,

forward and aft CG, and low and high assault rates. The incorporation of state-dependent

damping improved the limiter performance by decreasing the AOA and/or sideslip

overshoot at high assault rates. The limiter is now a part of the baseline X-48B control

laws and will be used for envelope protection in subsequent flight tests.

61
Chapter 3: Dynamic Inversion with Adaptive Augmentation

The baseline dynamic inversion control solution presented in Chapter 1 assumed

complete knowledge of the system dynamics. In practical applications, the inevitable

presence of system uncertainties requires additional control design. We propose using

baseline dynamic inversion control presented in Chapter 1 along with adaptive

augmentation to provide tracking performance in the presence of system uncertainties.

The adaptive augmentation component provides tracking performance in the presence of

the system uncertainties.

As in the baseline control design, the control architecture including adaptive

augmentation will be developed for a class of second-order systems in the cascaded form.

This particular class of systems is chosen primarily to clarify and expose key features of

the design process. These systems also naturally appear in flight dynamics and control

problems, which constitute the primary focus and motivation for the control

development. One may immediately note that since the design is based on the dynamic

inversion method, the developed controller can be extended to a generic class of feedback

linearizable MIMO systems with cascade-connected dynamics.

The results of this chapter contribute theoretically justified adaptive augmentation

design for such a generic class of feedback linearizable MIMO systems in cascade form.

The design will be further improved in the next chapters by including the state limiter

previously developed to form an aggregate system containing the combined benefits of

each of the design components.

62
Chapter 3.1: PID Control for Cascaded Systems with Uncertain Dynamics

Similar to Chapter 1.1, we consider 2nd order dynamical systems in the cascaded

form, but now include system uncertainties:

 x1  F1  x1 , z   B1 x2  f1  x1 , z 
 0

 (3.1)
 x2  F2  x1 , x2 , z   x2  f 2  x1 , x2 , z 
0 cmd

where components  f1 , f 2  are unknown continuously differentiable functions that

represent the system uncertainties. The control goal remains to design the control input

x2cmd so that the system 1st state component x1 tracks any given bounded time-varying

command x1cmd  t  while keeping all the signals in the closed-loop system bounded,

uniformly in time, now in the presence of system uncertainties  f1 , f 2  .

As in (1.5), we differentiate the first system state component to yield:

x1  f  x1 , x2 , z, z   B1 x2cmd  d  x1 , x2 , z, z  (3.2)

where d  x1 , x2 , z, z  now represents the unknown system uncertainty. We now introduce

modified Dynamic Inversion (DI) based Proportional + Integral + Derivative (PID)

control is introduced in the form:

  x  xm  
 
x2cmd  B11  x1m  f  K D xˆ1  x1m  K P  x1  x1m   K I  1 1   v 
 s  
(3.3)

where v represents adaptive augmentation component of baseline dynamic inversion

control law first introduced in (1.6). Substituting (3.3) into (3.2) yields the updated

tracking error dynamics:

63
e1
e1   K D e1  K P e1  K I   d  K D f1   v (3.4)
s
D  x1 , x2 , z , z 

Via Introduction of aggregate state x   x1 , x2 , z, z  , (3.4) may be expressed as:


T

e1
e1   K D e1  K P e1  K I  Dx  v (3.5)
s

Select adaptive augmentation control signal v to dominate the system uncertainties

online:

v Dˆ  x1 , x2 , z, z   ˆDT  D  x1 , x2 , z, z  (3.6)

We immediately note that an online approximation of the system uncertainty is not

required in the proposed design, only the ability to dominate the system uncertainty. This

is an important facet of the design, since approximation would require persistency of

excitation conditions [6], [28] that are not verifiable in real-time system operation.

Linear-in-parameters online representation of the uncertain function D  x  in (3.5) is

performed on a compact x region  , using an N-dimensional regressor vector

D  x   N
, with radial basis functions (RBFs) [29]:

Dˆ  x   ˆDT  D  x  (3.7)

where ˆD  N
is the vector of online estimated parameters. It is assumed that the

number of RBF components N is large enough, and the components are chosen so that the

uncertainty D  x  can be represented within the prescribed tolerance  Dmax on  :

D  x    D*   D  x    D  x 
T
(3.8)

64
In (3.8),  D* denotes the vector of true unknown constant parameters, and  D is the

unknown bounded representation error, with a known upper bound:

 D  x    Dmax (3.9)

Subtracting (3.8) from (3.7), the function representation error can be expressed in terms

of the parameter estimation error:

 
T
eD Dˆ  D  ˆD   D*  D   D   DT  D   D (3.10)
 D

Returning to the tracking error dynamics, substituting (3.6) into (3.5) and using

relation (3.10) yields:

e1
e1   K D e1  K P e1  K I  eD (3.11)
s

Using previously derived PID gains (1.11) and substituting into (3.11) yields:

e1    2    k  e1   2  2   k  e1   2 k
e1
 eD (3.12)
s

Regrouping terms in (3.12) gives:

 ke 
e1  k e1  2    e1  k e1    2  e1  1   eD (3.13)
 s 

We now introduce the filtered tracking error:

k e1  s  k 
e1f  e1   e1 (3.14)
s s

Using (3.13) and (3.14) yields filtered tracking error dynamics:

e1f  2   e1f   2 e1f  eD (3.15)

Filter tracking error dynamics may be expressed in matrix form as:

65
 e1f   0 1   e1f  0
 f  2      eD (3.16)
 e1    2     e1f  1
ef Aref ef Bref

The main idea we wish to exploit now is that by controlling the filtered tracking error,

we will in turn ensure control of the original tracking error, or more specifically:

e1f  0  e1f  0  e1  0 (3.17)

We show (3.17) holds as follows. First, if the time derivative of the filtered tracking error

e1f  t  is driven to become small then the original tracking error signal e1  t  will also

become small. This statement directly follows from the definition (3.14), which could

alternatively be written as:

e1  k e1  e1f (3.18)

The explicit solution to (3.18) is:

e1  t   e e1  t0    e e1f   d
 k  t t0  t  k  t  
(3.19)
t0

Consequently, if T such that e1f  t   1 , t  t0  T , then as t   we have:

e1  t   e e1  t0   1 
 k  t t0 T 
e  k t   d
t

t0 T
(3.20)
1 1
e
 k  t t0 T 
e1  t0  
k
1  e  k  t t0 T 
 k
1
In other words, the absolute value of the original tracking error e1  t  approaches
k

exponentially fast.

66
Second, if the filtered tracking error e1f  t  is driven to become small then the original

tracking error signal e1  t  will also become small. Integrating (3.19) by parts yields:

e1  t   e e1  t0    e e1f   d
 k  t  t0  t  k  t  
t0

e1  t0   e  k t   e1f    k  e  k t  e1f   d
 k  t  t0  t t
e (3.21)
t0 t0

 k  t  t0 
 e t   e t   e t   k   k  t  
e1f   d
t
e 1 0 1
f
1
f
e
t0

Consequently, if T such that e1f  t   1 , t  t0  T , then as t   we have:

e1  t   e e1  t0  T   e1f  t0  T   1  k 1 
 k  t t0 T 
e  k t   d
t

t0 T
(3.22)
e
 k  t t0 T 
e1  t0  T   e1f  t0  T   1  1 1  e   k  t t0 T 
  21

In other words, if e1f  t   1 , then as t   the following asymptotic relation takes

place:

e1  t   2 1  o 1 (3.23)

Chapter 3.2: Parameter Adaptation and Closed-Loop System Dynamics

Choose a symmetric positive-definite matrix Q  0 and solve the following algebraic

Lyapunov equation:

P Aref  Aref
T
P  Q (3.24)

Since Aref is Hurwitz by design, the Lyapunov equation has unique positive-definite

symmetric solution P , which is used to introduce a Lyapunov function candidate in the

form:

67
V  e f ,  D   eTf P e f    DT D1  D (3.25)

where symmetric positive-definite matrix D defines the rate of adaptation.

Differentiating (3.25) along the trajectories of the system (3.16) yields:

V  eTf Q e f  2 eTf P Bref eD  2   DT  D1 ˆD (3.26)

Regrouping terms and using (3.10) gives:


V  eTf Q e f  2 eTf P Bref  D  2  DT  D eTf P Bref   D1 ˆD  (3.27)

To make the time derivative of V in (3.27) negative outside of a compact  e,   

subset of  choose the following parameter adaptation laws:


ˆD   D proj ˆD ,  D eTf P Bref  (3.28)

In (3.28) proj denotes the projection operator [30], which forces the adaptive parameters

to evolve in a pre-specified   D region. Substituting (3.28) into (3.27) yields:

V  eTf Q e f  2 eTf P Bref  D (3.29)

Using (3.29), we must now show closed-loop stability and boundedness of the error

dynamics.

Assumption 1: The command x1cmd for reference model (1.2) is chosen such that

 x t  x2m  t  z  t  z  t   
m T
1 (3.30)

forward in time.

We begin by noting that (3.29) implies that

68
V  eTf Q e f  2 eTf P Bref  D
(3.31)
 min  Q  e f
2
 2 e f P Bref  Dmax

where min  Q  is the minimum eigenvalue of Q and  Dmax max x  D  x  . We also note

that because of the projection operator the norm of the parameter estimation error will

stay uniformly bounded; that is,

   t    max   (3.32)

Using (3.31) we can now establish uniform ultimate boundedness (UUB) [8] of the

closed-loop system trajectories. Toward that end, define the following compact subset in

the e f region:


 2 P Bref  Dmax 

Sr  ef  r  (3.33)


min  Q  

Also define a minimal level set b  eTf P e f  b that contains S r . Since

min  P  e f  eTf P e f  max  P  e f


2 2
(3.34)

then choosing

b  max  P  r 2 (3.35)

implies that for all e f  r ,

eTf P e f  max  P  e f  max  P  r 2  b


2
(3.36)

Hence, the set S r is contained in the level set b .

69
Suppose that all initial values of the filtered tracking error e f  t0  belong to a

compact set S R e f 
 R . Let B  eTf P e f  B be the maximal level set which

belongs to S R . To maintain closed-loop system stability, a specific relation between the

boundaries for the sets b , B , Sr , and S R must be imposed. These sets will be used to

prove that the closed-loop system trajectories are UUB. Graphical representation of the

four sets is shown in Figure 19.

Figure 19: Representation of the Sets Sr  b  B  SR

Choose:

B  min  P  R 2 (3.37)

Then if eTf P e f  B then using (3.34) yields:

min  P  e f  eTf P e f  B  min  P  R 2


2
(3.38)

Consequently, e f  R ; that is, the filtered tracking error is in S R . Because of (3.31) and

(3.33), the time derivative V is negative outside of S r . Therefore, the filtered tracking

70
error e f will enter level set b in finite time, and will remain in the set from then on.

Therefore, the closed-loop system trajectories are UUB. Moreover, due to the use of the

projection operator, all the estimated parameters are bounded. Hence, the tracking

problem is solved. In summary, the corresponding total explicit model following control

signal can be written using (3.3), (3.6), and (3.7):

 PID controller

 m  KI  
 B1  x1  f  x    K D s  K P    x1  x1    B1  D  D  x 
1 1 ˆT
x2cmd m
(3.39)
  s   adaptive augmentation
 
baseline dynamic inversion controller

Chapter 3.3: Design Example: AOA Tracking

This section applies the developed adaptive control methodology to construct an

AOA tracking system for a fixed-wing aircraft, whose short period dynamics including

lift and pitching moment uncertainties can be written as a modified version of (1.15):

   L   Qgrav  q   L  

 (3.40)
q  M    M q q  M IC  qcmd   M  , q 

where now  L   is the lift force uncertainty and  M  , q  represents the pitching

moment uncertainty. The AOA reference model dynamics is chosen in the form of (1.16)

and the corresponding baseline dynamic inversion controller is in the form of (1.21):

q  qm
bl
qcmd  qm  M    M q q  M IC  2    q  qm    2 (3.41)
s

The system dynamics are in the form of (3.1), where:

71
x1   , x2  q, z   Qgrav M IC  , x2cmd  qcmd
T

F1   L   Qgrav , B1  1, F2  M    M q q  M IC (3.42)
f1   L   , f 2   M  , q 

Using (3.15), the filtered tracking error signal becomes:

 s  k1  e   s  L      qm  q
e1f  1  m  (3.43)
s s s

Hence, the filtered tracking error vector is:

 q q
T

ef  e 
T
1
f
e1
f
 m qm  q  (3.44)
 s 

The regressor vector  D is chosen to depend on AOA only. Then parameter

adaptation laws are written based on (3.28):

  qm  q  0
ˆD   D proj  ˆD ,  D    qm  q  P    (3.45)
  s  1

where P  PT  0 is the unique positive definite symmetric solution of the algebraic

Lyapunov equation (3.24) via the Hurwitz reference model matrix Aref as specified in

(3.16). Based on (3.7), adaptive pitch acceleration augmentation command becomes:

ad
qcmd  ˆDT  t   D   (3.46)

In summary, the total pitch acceleration command consists of the baseline dynamic

inversion command (3.41) and adaptive augmentation (3.46):

qcmd  qcmd
bl
 qcmd
ad
(3.47)

The simulation model is chosen to represent longitudinal dynamics of an aerial

vehicle, specifically an F-16 aircraft. Similar to previously derived longitudinal dynamics

72
(1.15), by neglecting the effects of gravity and thrust the short period aircraft dynamics

may be written in matrix form:

Z  Z 
    1      e 
  V  
  q   V  e
(3.48)
 q  M 
Mq   
   M e 

where  is aircraft AOA, q is aircraft pitch rate,  e is elevator deflection (control


input), V is the trimmed (constant) airspeed, and Z , Z q , Ze  
and M  , M q , M e  are

partial derivatives of the aerodynamic vertical force Z and pitching moment M with

respect to  , q,  e  respectively. Numerical values for the vehicle aerodynamic

derivatives were taken from [4] (Example 5.5-3, Table 3.4-3). These data represent an F-

16 aircraft trimmed at:

VT  502ft/sec, alt  0ft, q  300lb/ft 2


c.g.  0.35 c ,   2.11 deg

The resulting open-loop system matrices are:

 1.0189 1   0.0022 
A , B  (3.49)
 0.8223 1.0774   0.1756 

where  is in radians, q is in radians/second, and  e is in degrees.

Baseline flight control is designed for the baseline system without uncertainties

according to (3.41) with selected dynamics   0.6 and   2 rad/sec. The eigenvalues of

the reference dynamics along with their corresponding natural frequencies are shown

below:

73
1  1.2  0.6 j 1  0.6, 1  2
   (3.50)
2  1.2  0.6 j  2  0.6, 2  2

The baseline system response to a series of angle-of-attack command doublets is shown

in Figure 20. The response shows that in the absence of uncertainties, the baseline system

tracks the AOA doublet commands while maintaining elevator deflection and rate that are

well within acceptable limits.

10 cmd
ref
, deg.

0 
-5

-10
0 10 20 30 40 50 60 70 80

0.4
e, deg

0.2

-0.2

-0.4
0 10 20 30 40 50 60 70 80

10
, dps

0
dot

-5
e

-10
0 10 20 30 40 50 60 70 80

Time, sec

Figure 20: Baseline Closed-Loop Response to AOA Doublet without Uncertainties

74
Including system pitching moment uncertainty yields updated system dynamics:

    1.0189 1     0   0 
       qcmd    M  , q  (3.51)
 q   0.8223 1.0774  q   1    
Three types of matched uncertainties are added to the system: 1) linear-in-state

uncertainty K Tpert x , 2) control effectiveness constant uncertainty   0 , and 3) nonlinear-

in-state uncertainty in the form of (3.8). Addition of the uncertainties updates dynamics

(3.51) to:

    1.0189 1     0   0 
        qcmd   K T x  D x  (3.52)
 q   0.8223 1.0774  q   1   pert  

Numerical values for the uncertainties were chosen as:


 c 2
K Tpert   0.411 0.8619  ,   0.5, D  x   D    0.5 e 2 2
(3.53)

where the center of the Gaussian was set to  c  2  /180 and its width was   0.0233 .

This particular selection of numerical values for K Tpert x and  is equivalent to 50%

increase in the static instability M  , 80% decrease in the pitch damping M q , and 50%

decrease in the control input effectiveness. These changes imply that the vehicle became

50% more statically unstable, lost 80% of its pitch damping ability, and the aircraft

controllability decreased by 50%. In fact, this combination causes the open-loop system

to become unstable (eigenvalues enter the right-half plane), and in addition there is the

simultaneous injection of a significant nonlinear-in-state uncertainty. Such drastic

changes were motivated by intent to demonstrate the effectiveness of the proposed design

methodology. This particular example was also selected to be similar to previous work

75
presented in [31], [32] so that relevant performance comparisons may be available. The

system total uncertainty versus AOA was calculated at q  0 and is shown in Figure 21.

0.6

0.5

0.4

0.3

0.2

0.1

-0.1
-5 -4 -3 -2 -1 0 1 2 3 4 5

AOA, deg

Figure 21: Total Matched Uncertainty vs. AOA at q=0

76
With only the baseline controller in operation and with the uncertainties included, the

closed-loop system tracking performance degradation can be clearly observed from the

data that are shown in Figure 22. Although the tracking performance is poor, both the

control input and its rate remain small.

10 
cmd
ref
, deg.

0 
-5

-10
0 10 20 30 40 50 60 70 80

0.5
e, deg

-0.5

-1

-1.5
0 10 20 30 40 50 60 70 80

10
, dps

0
dot

-5
e

-10
0 10 20 30 40 50 60 70 80

Time, sec

Figure 22: Baseline Closed-Loop Response in Presence of Matched Uncertainties

77
To counteract the effects of the uncertainties, adaptive laws (3.45) were constructed

by solving algebraic Lyapunov equation (3.24) with reference matrix Aref as shown in

(3.16). The regressor vector  D   consisted of 11  -dependent and evenly spaced

Gaussians (RBFs). RBF centers were placed at [-10:2:10] degrees of AOA, and all RBF

widths were set to   0.0233 . The rate of adaptation was chosen to be:

 D  100 (3.54)

and the total control (elevator deflection) was formed as shown in (3.47). With the

baseline + adaptive augmentation control active, closed-loop system tracking

performance was recovered with acceptable elevator deflection and rates as shown in

Figure 23.

78
10
cmd
, deg. 5

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

1
e, deg

-1

-2
0 10 20 30 40 50 60 70 80

10
, dps

0
dot

-5
e

-10
0 10 20 30 40 50 60 70 80

Time, sec

Figure 23: Baseline+Adaptive Closed-Loop Response in Presence of Uncertainties

Finally, the norm of the adaptive parameter is shown in Figure 24. In addition to the

projection operator robustness modification in adaptive laws (3.28) employed to keep the

adaptive parameters within a bounded region, standard dead-zone modification is

79
included to keep the adaptive parameters from evolving when the tracking error is small.

The effect of the dead-zone is seen in the adaptive parameter response, where once the

tracking error becomes small the adaptive parameter stops evolving and remains constant.

2.5

1.5
Theta hat

0.5

0
0 10 20 30 40 50 60 70 80
Time, sec

Figure 24: Adaptive Parameter Norm vs. Time

80
Chapter 4: Dynamic Inversion with State Limiting and Adaptive

Augmentation

In Chapter 1 we derived a baseline dynamic inversion control solution which assumed

complete knowledge of the system dynamics. The results of the chapter contribute robust

baseline flight control architecture in which both aircraft performance and handling

qualities requirements may be achieved in a convenient manner. While the controller

developed in Chapter 1 was shown to solve the tracking problem, in practical application

there are additional requirements that must be considered during the design process.

In order to control a system while keeping the dynamics within a permissible subset

of the system state space, in Chapter 2 we introduced state limiting augmentation to the

baseline dynamic inversion controllers introduced in Chapter 1. The proposed method

introduces a limiting technique which gradually modifies the expected behavior of the

system dynamics when user-specified criteria are approached / exceeded. Given a

baseline model following dynamic inversion control law designed to meet robustness and

performance requirements throughout a flight envelope, a state-limiting augmentation

component protects the system trajectories from leaving an allowable subset in the

system state space.

In addition, in practical applications the inevitable presence of system uncertainties

requires additional control design. In Chapter 3, we proposed using baseline dynamic

inversion control presented in Chapter 1 along with adaptive augmentation to provide

tracking performance in the presence of system uncertainties. The adaptive augmentation

component provides tracking performance in the presence of the system uncertainties.

81
In this chapter we propose combining the contributions of Chapters 1-3 into an

integrated design. The resulting closed-loop system contains robust baseline control by

design, state limiting to maintain system trajectories within a prescribed region in the

system state space, as well as adaptive augmentation to maintain tracking performance in

the presence of system uncertainties. Formal stability proofs and analyses are included

for the combined design. The results of this chapter contribute a theoretically justified

robust + adaptive design including state limiting for a generic class of feedback

linearizable MIMO systems in cascade form.

82
Chapter 4.1: Baseline+Adaptive Control+State Limiting for Uncertain

Dynamics

Similar to Chapter 3.1, we consider 2nd order dynamical systems in the cascaded form

including system uncertainties:

 x1  F1  x1 , z   B1 x2  f1  x1 , z 
 0

 (4.1)
 x2  F2  x1 , x2 , z   x2  f 2  x1 , x2 , z 
0 cmd

We again form baseline dynamic inversion PID control:

  x  xm  
 
x2cmd  B11  x1m  f  K D xˆ1  x1m  K P  x1  x1m   K I  1 1   v 
 s  
(4.2)

where v represents adaptive augmentation component of baseline dynamic inversion

control law first introduced in (1.6). We once again form tracking error dynamics:

e1
e1   K D e1  K P e1  K I  Dx  v (4.3)
s

and select adaptive augmentation control signal v to dominate the system uncertainties

online:

v Dˆ  x1 , x2 , z, z   ˆDT  D  x1 , x2 , z, z  (4.4)

Again using radial basis functions in the regressor vector and analyses (3.7)-(3.10) we

can express the updated tracking error dynamics:

e1
e1   K D e1  K P e1  K I  eD (4.5)
s

We now update baseline PID gains (1.11) to include state limiting component developed

in Chapter 2:

83
 K D  2  x  x  x  x   k

 K P  2  x  x  x  x  k  x  x 
2
(4.6)

 K I  x  x  k
2

Using methodology described in (3.12)-(3.15) we have updated filtered error dynamics

(3.16) that now include both the adaptive augmentation and state limiting design

improvements:

 e1f   0 1   e1f  0


 f 
       eD (4.7)
  x  x  2  x  x  x  x    e1f
2
 e1  1
ef Aref ef Bref

The dependence of Aref on system state x is due to the inclusion of state limiter, while

the use of filtered tracking error and parameter estimation error eD is due to the inclusion

of adaptive augmentation.

We immediately note that similar to state limiter results in Chapter 2, there are three

cases for which stability of system (4.7) must be investigated:

Case 1: x  x0  xlim   . In this region, the modulation function is inactive, and the

error dynamics may be expressed in terms of the nominal damping  and natural

frequency  . In this case, the stability results of the LTI system including adaptive

augmentation presented in Chapter 3 directly apply with Aref in (4.7) corresponding to

the nominal damping and frequency.

Case 2: x  x0  xlim . In this region, the modulation function is completely active, and

the error dynamics may be expressed in terms of the full incremented damping    

84
and natural frequency     . Again, in this case the stability results of the LTI system

including adaptive augmentation presented in Chapter 3 directly apply with Aref in (4.7)

corresponding to:

 0 1 
Aref    (4.8)
       2          
2

Case 3: xlim    x  x0  xlim . In this region, the modulation function is active and

likely changing due to the evolution of the system states. Without adaptive augmentation

present, we illustrated the stability of this system using Lyapunov function (2.42):

1 2
V  t , E   e12  e2 (4.9)
g t 

We now note that Lyapunov function (4.9) may alternatively be expressed as:

V  t , E   eT R  t  e (4.10)

where:

1 0 
R  t    1  , R  t   0 t  0 (4.11)
0
 g  t  

due to the properties of g  t  documented in (2.39)-(2.40). Motivated by (4.10) and by

Lyapunov function used in the stability proof for LTI case with adaptive augmentation

(3.25), we propose Lyapunov function:

V  e f ,  D   eTf R  t  e f    DT D1  D (4.12)

85
where R  t  is defined in (4.11). Differentiating (4.12) along the system trajectories (4.7)

yields:

V   K e2f  2 eTf R  t  Bref eD  2   DT D1 ˆD (4.13)

where K is defined as in (2.43). We immediately note that based on analysis shown in

(2.44)-(2.54), term  K e2f in (4.13) ensures asymptotic stability in the case without

uncertainties and adaptive augmentation, and therefore the second and third terms in

(4.13) are of interest here. Rearranging terms in (4.13) yields:


V   K e2f  2 eTf R  t  Bref  D  2  DT  D eTf R  t  Bref   D1 ˆD  (4.14)

To make the time derivative of V in (4.14) negative outside of a compact  e f ,   

subset of  choose the following parameter adaptation laws:


ˆD   D proj ˆD ,  D eTf R  t  Bref  (4.15)

Substituting (4.15) into (4.14) yields:

V   K e2f  2 eTf R  t  Bref  D (4.16)

Using (4.16), we must now show closed-loop stability and boundedness of the error

dynamics. Similar to analysis presented in Chapter 3:

Assumption 1: The command x1cmd for reference model (1.2) is chosen such that

 x t  x2m  t  z  t  z  t   
m T
1 (4.17)

forward in time.

We begin by noting that (4.16) implies that

86
V   K  t  e2f  2 eTf R  t  Bref  D
(4.18)
  K  t  Bref  2 e f R  t  Bref  Dmax
2
ef

We further note that because previous analysis has shown that term  K e2f ensures

asymptotic stability in the case without uncertainties, we may bound (4.18) further by

considering the case where K  t  is minimum and R  t  is maximum, corresponding to

the “worst” case from a Lyapunov stability analysis:

2
V   K min Bref ef  2 e f Rmax Bref  Dmax (4.19)

We note that the minimum and maximum values of matrices K  t  and R  t  are known

due to their definitions and the properties of f  t  , g  t  given in (2.39). We also note that

because of the projection operator the norm of the parameter estimation error will stay

uniformly bounded; that is,

   t    max   (4.20)

Similar to Chapter 3, using (4.19) we can now establish uniform ultimate

boundedness (UUB) of the closed-loop system trajectories. Toward that end, define the

following compact subset in the e f region:


 2 Rmax Bref  Dmax 

Sr  ef  r  (4.21)

 K min Bref 

Also define a minimal level set b  eTf R  t  e f  b that contains S r . Since

 eTf R  t  e f  Rmax
2 2
Rmin ef ef (4.22)

87
then choosing

b  Rmax r 2 (4.23)

implies that for all e f  r ,

eTf R  t  e f  Rmax
2
ef  Rmax r 2  b (4.24)

Hence, the set S r is contained in the level set b . Suppose that all initial values of

the filtered tracking error e f  t0  belong to a compact set S R e f 


 R . Let

B  eTf R  t  e f  B be the maximal level set which belongs to S R . To maintain

closed-loop system stability, a specific relation between the boundaries for the sets

b , B , Sr , and S R must be imposed. These sets will be used to prove that the closed-

loop system trajectories are UUB. Choose:

B  Rmin R 2 (4.25)

Then if eTf R  t  e f  B then using (4.22) yields:

 eTf R  t  e f  B  Rmax R 2
2
Rmin ef (4.26)

Consequently, e f  R ; that is, the filtered tracking error is in S R . Because of (4.19) and

(4.21), the time derivative V is negative outside of S r . Therefore, the filtered tracking

error e f will enter level set b in finite time, and will remain in the set from then on.

Therefore, the closed-loop system trajectories are UUB. Moreover, due to the use of the

projection operator, all the estimated parameters are bounded. Hence, the tracking

problem is solved.

88
Chapter 4.2: Design Example: AOA Tracking

In this section, we apply the combined baseline control + adaptive augmentation

along with state limiting methodology to construct an AOA tracking system for a fixed-

wing aircraft. Similar to the example developed in Chapter 3, we note that fixed-wing

aircraft short period dynamics including lift and pitching moment uncertainties can be

written as a modified version of (1.15):

   L   Qgrav  q   L  

 (4.27)
q  M    M q q  M IC  qcmd   M  , q 

where  L   is the lift force uncertainty and  M  , q  represents the pitching

moment uncertainty. The AOA reference model dynamics is chosen in the form of (1.16)

and the corresponding baseline dynamic inversion controller is in the form of (1.21) but

now including the state limiter improvement:

q  qm
 qm  M    M q q  M IC  2  x  x  x  x  q  qm   x  x 
bl 2
qcmd (4.28)
s

The system dynamics are in the form of (3.1), where:

x1   , x2  q, z   Qgrav M IC  , x2cmd  qcmd


T

F1   L   Qgrav , B1  1, F2  M    M q q  M IC (4.29)
f1   L   , f 2   M  , q 

Using (3.15), the filtered tracking error signal becomes:

 s  k1  e   s  L      qm  q
e1f  1  m  (4.30)
s s s

89
Hence, the filtered tracking error vector is:

 q q
T

ef  e 
T
1
f
e1
f
 m qm  q  (4.31)
 s 

The regressor vector  D is chosen to depend on AOA only. Based on stability

analysis developed in the previous section, we note that the parameter adaptation laws

have changed due to the presence of state limiting component and are now written based

on (4.15):

  qm  q  0
ˆD   D proj  ˆD ,  D    qm  q  R  t     (4.32)
  s  1

where R  t   R  t   0 is defined in (4.11). Similar to Chapter 3, adaptive pitch


T

acceleration augmentation command becomes:

ad
qcmd  ˆDT  t   D   (4.33)

In summary, the total pitch acceleration command consists of the baseline dynamic

inversion command (4.28) and adaptive augmentation (4.33):

qcmd  qcmd
bl
 qcmd
ad
(4.34)

The simulation model is again chosen to represent longitudinal dynamics of an F-16

aircraft. Similar to previously derived longitudinal dynamics (1.15), by neglecting the

effects of gravity and thrust the short period aircraft dynamics may be written in matrix

form:

Z  Z 
    1      e 
  V  
 q   V  e
(4.35)
 q  M M q     M e 
 

90
where  is aircraft AOA, q is aircraft pitch rate,  e is elevator deflection (control


input), V is the trimmed (constant) airspeed, and Z , Z q , Ze  
and M  , M q , M e  are

partial derivatives of the aerodynamic vertical force Z and pitching moment M with

respect to  , q,  e  respectively. The same numerical values for the vehicle aerodynamic

derivatives were employed as in Chapter 3, with resulting open-loop system matrices:

 1.0189 1   0.0022 
A , B  (4.36)
 0.8223 1.0774   0.1756 

where  is in radians, q is in radians/second, and  e is in degrees.

Baseline flight control is designed for the baseline system without uncertainties and

without limiting according to (4.28) with the same selected dynamics of   0.6 and

  2 rad/sec as used in previous examples. Just as shown in Chapter 3, the nominal

system response to a series of angle-of-attack command doublets is shown in Figure 25.

The response shows that in the absence of uncertainties, the baseline system tracks the

AOA doublet commands while maintaining elevator deflection and rate that are well

within acceptable limits.

91
10
cmd
, deg. 5

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

0.4
e, deg

0.2

-0.2

-0.4
0 10 20 30 40 50 60 70 80

10
, dps

0
dot

-5
e

-10
0 10 20 30 40 50 60 70 80

Time, sec

Figure 25: Baseline Closed-Loop Response to AOA Doublet without Uncertainties

We note that because of the selected damping ratio of 0.6, the response does exhibit

some overshoot. In order to arrest the overshoot, we employ the limiter feature developed

92
in Chapter 2. Without uncertainties and assuming an AOA limit of 5 degrees,   0.1 and

  0.5,   0.5 , the response including limiter is shown in Figure 26.

10
cmd
, deg.

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

0.5
e, deg

-0.5
0 10 20 30 40 50 60 70 80

20
, dps

10

0
dot

-10
e

-20
0 10 20 30 40 50 60 70 80

Time, sec

Figure 26: Closed-Loop Response to AOA Doublet: No Uncertainties, with Limiter

93
The response shows that both the reference model and corresponding system response

now remain within the system state limits. A detailed plot of the system response

showing the limiter activation is shown in Figure 27.

10
, deg.

cmd
5
limit area
0 ref
0 1 2 3 4 5 6 7

1
()

0.5

0
0 1 2 3 4 5 6 7

1.5
x

0.5
0 1 2 3 4 5 6 7

2.5
x

2
0 1 2 3 4 5 6 7

time, sec

Figure 27: AOA Doublet Response: No Uncertainties, with Limiter (Zoom)

94
In Chapter 3 we showed that in the presence of significant uncertainties, adaptive

augmentation is able to restore tracking performance for the system. We now show the

total baseline + adaptive augmentation + limiter system is able to retain tracking

performance in the presence of uncertainties while simultaneously keeping the system

within the state limits. Consider the same uncertainties as previously investigated, that is:

50% increase in the static instability M  , 80% decrease in the pitch damping M q , and

50% decrease in the control input effectiveness. These changes imply that the vehicle

became 50% more statically unstable, lost 80% of its pitch damping ability, and the

aircraft controllability decreased by 50%.

The adaptive law implemented in Chapter 3 must be slightly modified to

accommodate the inclusion of the state limiting feature, and is updated to (4.15). The

limiter parameters remain the same: assume an AOA limit of 5 degrees,   0.1 and

  0.5,   0.5 . Although the adaptive law has changed due to the inclusion of R  t 

rather than P , we keep the adaptive architecture the same is previous: the regressor

vector  D   consisted of 11  -dependent and evenly spaced Gaussians (RBFs). RBF

centers were placed at [-10:2:10] degrees of AOA, and all RBF widths were set to

  0.0233 . The rate of adaptation was chosen to be:

 D  200 (4.37)

We increased the adaptation rate to account for the additional deviation from the

reference model that the limiter may induce. The system response is shown in Figure 28:

95
10
cmd
, deg. 5

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

1
e, deg

-1

-2
0 10 20 30 40 50 60 70 80

20
, dps

10

0
dot

-10
e

-20
0 10 20 30 40 50 60 70 80

Time, sec

Figure 28: AOA Doublet Response: Uncertainties with Adaptive+Limiter

The response shows the cumulative benefit of the total design. The baseline design

provides robust control that is able to handle significant uncertainties. In the presence of

extreme uncertainty, the baseline performance degrades but is restored by the inclusion of

96
adaptive augmentation. The limiter modification further adds the ability to keep the

system within state limits while maintaining the benefits of the adaptive augmentation. A

detailed response plot is shown in Figure 29.

5
, deg.

cmd
limit area
0 ref
0 1 2 3 4 5 6 7

1
()

0.5

0
0 1 2 3 4 5 6 7

1.5
x

0.5
0 1 2 3 4 5 6 7

2.5
x

2
0 1 2 3 4 5 6 7

time, sec

Figure 29: AOA Doublet Response: Uncertainties with Adaptive+Limiter (Zoom)

97
The initial deviation of the response from the reference model shown in Figure 29 is

due to the uncertainties, but as shown the adaptive augmentation is able to restore

tracking performance while the limiter keeps the system within limits. The norm of the

adaptive parameter is shown in Figure 30.

1.8

1.6

1.4

1.2
Theta hat

0.8

0.6

0.4

0.2

0
0 10 20 30 40 50 60 70 80
Time, sec

Figure 30: Adaptive Parameter Norm vs. Time

98
In addition to the projection operator robustness modification in adaptive laws (4.32)

employed to keep the adaptive parameters within a bounded region, standard dead-zone

modification is included to keep the adaptive parameters from evolving when the tracking

error is small. The effect of the dead-zone is seen in the adaptive parameter response,

where once the tracking error becomes small the adaptive parameter stops evolving and

remains constant.

99
Chapter 5: Adaptive Design with Improved Performance under Input Time-

Delays

In Chapters 1-4 we derived a model-following control design resulting in a closed-

loop system containing robust baseline control by design, state limiting to maintain

system trajectories within a prescribed region in the system state space, as well as

adaptive augmentation to maintain tracking performance in the presence of system

uncertainties. While the dynamic inversion approach employed for the baseline control

design was derived for general second-order dynamical systems in cascaded form, it is

used extensively in aerospace flight control applications, particularly in the area of

piloted flight vehicles. The popularity of this design approach is largely due to the ability

of the designer to easily define target closed-loop natural frequency and damping ratios

that satisfy flying qualities requirements for piloted aircraft. These requirements can be

found in the FAA Federal Air Regulations (FAR) Part 25: Airworthiness Standards for

Piloted Aircraft and MIL-F-8785C (Military Specification for Flying Qualities of Piloted

Airplanes) for commercial and military piloted aircraft, respectively.

Another popular control architecture that is used heavily in aerospace applications is

optimal control. Application of optimal control methods, specifically Linear Quadratic

Regulator (LQR) techniques, relies on the inherent robustness properties provided by

LQR-optimal controllers. It is well known that proper selection of LQR design

parameters will achieve 6 dB gain margin and at least 60 degrees of phase margin at the

system control input break points. This is of significant importance for aerospace

applications, where uncertainties including latency and actuator dynamics can reduce

100
system robustness at the system control input. Therefore, the inherent robustness

properties of LQR make this technique very attractive for baseline control design in

aerospace applications, particularly in cases such as unmanned vehicles and munitions,

where uncertainties may be more dramatic and design goals often include maximum

performance. While the LQR approach does not offer the immediate insight into closed-

loop frequency and damping that dynamic inversion allows, LQR techniques are

attractive for piloted applications because of the inherent robustness and a good design

will be tailored to also meeting flying qualities requirements.

It has been shown that LQR optimal controllers can tolerate classes of uncertainties

that may exist in the system control channels, also called “matched” uncertainties since

they appear where control inputs exist in the system dynamics. In the presence of such

uncertainties, baseline closed-loop system performance will degrade. While the inherent

robustness properties of LQR are attractive, it is worth noting that these controllers are

designed to be robust to the entire class of matched uncertainties and therefore may

become overly conservative.

In this chapter, we use the LQR-optimal technique to provide a robust baseline

control design. Similar to the philosophy presented in previous chapters, we again use

adaptive augmentation to provide tracking performance in the presence of system

uncertainties. The now-classical model reference adaptive control (MRAC) design

approach is applied and modified to specifically address systems with matched

uncertainties and input time-delay. We then propose an alternative design approach to

address the uncertainties and input time-delay while improving performance under input

101
time delay. Using aerospace simulations consistent with previous chapters, we illustrate

that the design approach is able to restore baseline closed-loop performance in the

presence of both matched uncertainty and input time-delay, while showing improved

performance in the presence of input time-delay compared with classical MRAC design.

102
Chapter 5.1: Problem Definition and Baseline Optimal Control Design

Similar to Section 3.1 we consider dynamical systems including system uncertainties,

but without restricting the problem formulation to 2nd order cascaded form systems.

Instead consider a general class of MIMO uncertain systems:

  
 
x p  Ap x p  B p   I mm  u u  s  u  d  d  x p  
T  T
(5.1)
 
   s   d xp 

 

where x  is the system state vector and u 


np m
is the control input. The term

d  x p   Td  d  x p  (5.2)

N m
represents linear-in-parameters state-dependent matched uncertainty, where d  is

a matrix of unknown constant parameters, and  d  x p   N


is the known N-dimensional

regressor vector whose components are Lipschitz continuous functions of x p . The term

  s   Tu u  s  (5.3)

represents linear-in-parameters control multiplicative dynamic uncertainties, where

u  Nu  m
is a matrix of unknown constant parameters, and u  s   Nu m
is the known

 Nu  m  -dimensional regressor matrix whose components are strictly proper stable

transfer functions. In order to shorten notation, we are going to mix time and frequency

domain variables in one and the same equation, such as in (5.1). This simply allows us to

define a time-domain function whose Laplace transform equals (5.3). Furthermore, there

103
is no loss of generality to assume that components of the regressor are strictly proper. In

fact, suppose u  s  is only proper. Then we can write it as:

u  s    0   u  s  (5.4)

where  0 is a constant matrix and u  s  is a strictly proper transfer function matrix.

Substituting (5.4) into (5.1) gives:

  
x p  Ap x p  B p  I mm  Tu  0   u  s  u  d  x p  
 Ap x p  B p   I mm  Tu  0   Tu   s  u  d  x 
u p (5.5)

  
  
  
 Ap x p  B p   I mm  Tu  0    I mm   I mm  Tu  0  Tu  u  s   u  d  x p  
1

  
 Tu
 
  
   s   

where we assumed that I mm  Tu 0  is invertible. We assume that Bp  nm


is

nn mm
constant and known, Ap  is constant and unknown, and   is a constant,

diagonal, and unknown matrix with positive diagonal elements corresponding to control

input deficiencies in each input channel.

Motivation for the inclusion of (5.3) in the problem statement comes from the desire

to mitigate time delays in the input channel. The transfer function of an input time delay

 is given by

e s u (5.6)

This expression can cause difficulties in control design and analysis because (5.6) is not a

rational transfer function. Therefore in practice, Padé approximations [33] are often used

104
to approximate a pure time delay by a proper rational transfer function. Use of Padé or

other rational function approximation (RFA) techniques to model time delay allows for

application of problem formulation (5.1) by noting that

U  t      I mm    s  u  t  (5.7)

For example consider a time-delay with   0.01 second corresponding to a one-

sample time-delay in a system with a sample rate of 100Hz, which is a typical

representative rate for modern aerospace control systems. Figure 31 below shows the

time response of the pure time-delay, along with that of two Padé approximations of

order one and four respectively:

105
1.4

1.2

0.8
Input

0.6

0.4

u
udelay
0.2
u1st order Pade
u4th order Pade
0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

Time, sec

Figure 31: Time Response of Pure Time-Delay and Padé Approximations

In addition, the frequency response is shown below in Figure 32.

106
1

Magnitude (dB)
0.5

0
udelay
-0.5 u1st order Pade
u4th order Pade
-1
-1 0 1 2 3
10 10 10 10 10

-100
Phase (deg)

-200

-300

-400

-500

-600
-1 0 1 2 3
10 10 10 10 10

Frequency, rad/sec

Figure 32: Frequency of Pure Time-Delay and Padé Approximations

The time and frequency domain responses show that the Padé approximation is able

to approximate the pure-time delay with increasing fidelity by increasing the order of the

Padé approximation. However, increased order can lead to numerical instability, and

107
therefore in practice the designer must balance accuracy of the approximation against

numerical robustness requirements. In the case of the 1st order Padé approximation, the

corresponding transfer function is:

 s  200
(5.8)
s  200

Therefore in this example, the dynamic uncertainty in terms of (5.7) would be:

400
1  (5.9)
 I11 s  200
 s 

Returning to the problem formulation, the control goal remains the same: Design the

control input u such that the system controlled (i.e. regulated) output

y  Cp xp  m
(5.10)

tracks any given bounded time-varying external command while keeping all the signals in

the closed-loop system bounded, in the presence of the system uncertainties

 A ,  ,  ,  .
p d u

Define the system tracking error:

ey  t   y  t   r  t  (5.11)

Augmenting the system dynamics (5.1) with the integrated tracking error

eyI  ey  y  r (5.12)

yields extended open-loop dynamics:

 
x  A x  B   I mm    s   u  d  x p   Bc r (5.13)

where x   eTyI xTp   n


is the extended system state vector. The system matrices are:

108
 0mm Cp   0mm    I mm 
A  , B   , Bc  
  0n m 
(5.14)
 0 n m Ap 
 p  Bp   p 

with system controlled output

y   0 pm Cp  x  C x (5.15)

While constant matrix A is assumed to be unknown, we assume that the well-known

model matching conditions [7] are met for the system of interest. That is, given a known

nn
constant matrix Ak  and the unknown positive-definite diagonal constant matrix

 , there must exist a possibly unknown constant gain matrix K d such that:

A  Ak  B  KdT (5.16)

Using assumption (5.16), system dynamics (5.13) may be written as:

 
  x  
  
x  Ak x  B   u   K dT Td Tu    d  x p     Bc r (5.17)
  


T
 u  s  u  
  

As in previous chapters, a baseline control design is performed for the system

dynamics (5.17) in the absence of uncertainties, which corresponds to the situation when

  I , Td  d  x p   0, KdT  0 and the time delay   0    s   0 . Using the LQR

servomechanism design technique [34], an optimal linear state feedback control solution

in PI form is derived as:

ubl   K xT x (5.18)

109
We note that any properly designed state feedback control solution of the form (5.18)

would be acceptable for baseline design. We choose the LQR optimal architecture

because of the robustness properties inherent in the design. Since the adaptive control

signal will be designed as an augmentation, it is desirable to have maximum robustness in

the baseline design to guarantee adequate system performance in the absence of

uncertainties. Substituting baseline control (5.18) into the system dynamics (5.17) in the

case without uncertainties leads to the closed-loop dynamics:

x   Ak  B  K xT  x  Bc r (5.19)
Aref

The closed-loop system (5.19) represents the desired closed-loop dynamics in the

absence of uncertainties, and will be used as the reference model for the model reference

adaptive control augmentation scheme.

Chapter 5.2: Adaptive Augmentation via Classical MRAC Architecture

The baseline controller was designed for the system under nominal conditions. If

uncertainties are present, the baseline controller may no longer provide adequate

performance or stability. Therefore, we introduce classical MRAC adaptive augmentation

to cope with the system uncertainties. We use the reference model defined in (5.19) and

synthesize the total control input as the sum of the baseline LQRPI component and

adaptive augmentation (to be constructed):

110
u  ubl  uad   K xT x  Kˆ xT x  dˆ  x p   ˆ  s  u
ubl uad (5.20)
ˆT 
 ubl  

In (5.20), Kˆ x  nm
is the incremental adaptive feedback gain designed to counteract the

effects of system uncertainties  and K d , ˆ  s   m


is the on-line approximator of the

system input time-delay uncertainty approximation, and dˆ  x p   m


is the on-line

approximator of the matched system uncertainty d  x p  . On-line approximator ˆ  s 

contains regressor matrix u  s   Nu m


whose components are strictly proper transfer

ˆ 
functions with stable poles and incremental adaptive weight  Nu  m
:
u

ˆ  s   
ˆ T  s
u u (5.21)

With the Padé approximations, the regressor matrix u  s  in (5.21) would contain

Nu  m Padé transfer functions which may be a combinations of various orders and time-

delay values selected based on the knowledge of the system characteristics. The on-line

approximator dˆ  x p  contains the regressor vector   x p   N m


with N radial basis

ˆ 
functions (RBFs) and the adaptive weight matrix  N m
:
d

dˆ  x p   
ˆ T  x 
d d p (5.22)

Substituting (5.20) into the original system dynamics (5.17) gives:

 
 
x  Aref x  B   uad   I    K x x      Bc r
1 T T
(5.23)
 
 T  

111
or equivalently,

x  Aref x  B   uad  T    Bc r (5.24)

with the defined regressor vector:


   x  d  x p  u  s  u 
T
(5.25)

and the matrix of unknown / ideal parameters:

T
 
 
   K dT   I   1  K xT  (5.26)
 
 K xT 

Substituting adaptive component uad from (5.20) into dynamics (5.24) yields

 
T
ˆ   B r
x  Aref x  B   (5.27)
c



where  is the matrix of parameter estimation errors. We now introduce the state

tracking error

e  x  xref (5.28)

and calculate the tracking error dynamics:

e  Aref e  B  T  (5.29)

Using Lyapunov design approach, bounded output tracking is achieved through on-

line parameter adaptation laws [34]:

ˆ    eT P B
 (5.30)

Or, in terms of the system parameters:

112
Kˆ x   x x eT P B
ˆ     x  eT P B
 (5.31)
d d d p

ˆ     s  u eT P B
 u u u


In (5.31), symmetric positive-definite matrices  x , d , u  represent the adaptation

rates and P is the unique symmetric positive-definite solution of the Lyapunov algebraic

equation:

P Aref  Aref
T
P  Q (5.32)

for a symmetric positive-definite matrix Q. Extending the design to MIMO systems with

non-parametric uncertainties is straightforward and can be accomplished using well-

known in adaptive control robustness methods. For this control scheme in

implementation, the dead-zone and the Projection Operator robustifications are critical.

We note that design approach similar to that presented here resulting in adaptive laws

(5.31) is well established and has been applied to many aerospace applications in the past

decade [1]-[3], [5], [19], [31], [32], [34]. A notable difference presented here is the

inclusion of term (5.3). The motivation for this term is to address time-delays, but the

analysis can be used for any stable proper dynamic uncertainty meeting the conditions of

(5.3).

Relations (5.20) and (5.30) solve the tracking problem with globally asymptotically

stable closed-loop dynamics for any symmetric positive definite rates of adaptation   .

However, it is well known that if this matrix has large singular values, the system will

113
often contain undesirable transient oscillations. We will show such behavior in the

forthcoming aerospace simulations.

Chapter 5.3: Alternative Adaptive Augmentation via B-SPM Model

In order to explore the performance of adaptive design (5.31) under time-delay, we

use an alternative design architecture developed in [7], [28]. Begin by assuming the ideal

case, that is, when the uncertainties present in the system are exactly known. Substituting

total control signal (5.20) into the system dynamics (5.17) yields:

 
x  Ak x  B   K xT x  Kˆ xT x  dˆ  x p   ˆ  s  u  T   Bc r (5.33)

We make the following assumptions:

1. Matrix  is invertible

2. Matrix B is full rank

Note: assumption 1 is satisfied by the definition of  as a positive definite diagonal

matrix. In order to achieve the desired system response indicated by the reference model

(5.19), setting:

Kˆ xT  K dT   I   1  K xT
dˆ  x p   
ˆ T   x   T   x 
d d p d d p (5.34)
ˆ  s   
ˆ T   s   T   s 
u u u u

will recover the reference dynamics in the presence of uncertainties. This is due to the

augmentation explicitly cancelling the system uncertainties. The fact that relations (5.34)

may be derived to solve the tracking problem proves the existence of a control solution,

indicating that the problem must be solvable.

114
The assumption that the system uncertainties are exactly known is unrealistic – if it

were true such terms would have been included in the baseline system dynamics and not

treated as uncertainties. We therefore attempt to approximate parameters Kˆ x , ˆ  s  ,

dˆ  x p  using methods from [7], [28] and bilinear static parametric model (B-SPM).

Given (5.17), we wish to express the system in the form:

Y    u  T   (5.35)

by collecting unknown terms on one side and then filtering both sides. First, define the

matrix of unknown parameters:

T   KdT Td Tu  (5.36)

and the known regressor matrix:

T    xT   su 
T
Td (5.37)
 u

The system dynamics (5.17) may now be expressed as:

x  Ak x  B   u  T    Bc r (5.38)

Dynamics (5.38) make up the model that will be used for estimation. Collecting known

terms on one side, and again assuming that B is full rank yields:

B   x  Ak x  Bc r     u  T   (5.39)

where B  is the Moore-Penrose inverse. We do not assume availability of the state

derivative as a measurement, and therefore filter each side of (5.39) with a strictly proper

stable transfer function G  s  . An additional benefit of the filtering technique is the

ability to shape the measured quantities and to avoid known frequency regions associated

115
with uncertain dynamics, noise, and/or disturbances. Denote filtered quantities with a

subscript, i.e.

xf  G s x (5.40)

Filtering each side of (5.39) results in:

B   x f  Ak x f  Bc rf     u f  T  f  (5.41)
Y

We can now compute x f due to relation (5.40), and see that (5.41) is in the desired form

of the B-SPM (5.35). The unknown parameters are on the right-hand side and are

replaced with their respective estimates to form the estimated output:

Yˆ   
ˆ u 
f
ˆT 
f  (5.42)

Using a gradient-based adaptive law [7], [28] we obtain:

ˆ    T
  f
(5.43)
ˆ  u 
  f 
ˆ T  T
f 
where the normalized estimation error is:

Y  Yˆ
 (5.44)
ms2

and the normalizing signal ms is designed to bound  f , u f from above. An example of

ms with this property is [7], [28]:

ms2  1  Tf  f  uTf u f (5.45)

The adaptive gains are free design parameters that satisfy   T  0 and   T  0 .

For simplicity we may allow   to take the form:

116
 0 0 
 d 
   0  d 0  (5.46)
 
 0 0  u 

Using (5.46) we may express the parameter update equations (5.43) as:

Kˆ d   d x f  T
ˆ    T
 d d df
(5.47)
ˆ    su  T
 u u u f

f  d f
ˆ T  x  T
ˆ   u  Kˆ T x  
  d df p 
Extending the design to MIMO systems with non-parametric uncertainties is

straightforward and can be accomplished using well-known in adaptive control

robustness methods. For this control scheme in implementation, the dead-zone and the

Projection Operator robustifications are critical. The total control signal (5.20) is derived

in terms of adaptive parameters (5.47):

 
u  ubl  uad   K xT x  Kˆ dT  I     x
ˆ T x   
ˆ 1 K T x  
d d p
ˆ T  su
u u (5.48)
ubl
uad

This is a dynamic controller and its output, the control signal u , is realizable (i.e.

computable) since u  s  is a strictly proper transfer function. Following analysis

presented in [7], [28] adaptive law (5.47) provides closed-loop stability with all signals in

the closed loop system bounded. In addition, it provides that the regulation error will be

of the order of the modeling error in mean square sense. In this case, the modeling error

does not exist since we have assumed that the dynamic uncertainty may be exactly

matched via (5.3). This implies that e  0 as t   , which indicates that the system

117
state trajectory will track the reference model. Consequently, the closed-loop system is

stable, all signals in the closed-loop system are bounded, and the model reference

tracking error asymptotically converges to zero forward in time.

Chapter 5.4: Design Example: AOA Tracking

We demonstrate the efficacy of the adaptive schemes via simulation on a simplified

model which is representative of flight dynamics and which has been used extensively in

published adaptive control literature. The simulation will also be used to show the

improved performance characteristics of the proposed alternative adaptive design

including normalization. As in previous chapters, the simulation model is again chosen to

represent longitudinal dynamics of an F-16 aircraft. Similar to previously derived

longitudinal dynamics (1.15), by neglecting the effects of gravity and thrust the short

period aircraft dynamics may be written in matrix form:

Z Zq   Z e 
    1      
  V V     V e (5.49)
 q  M M q     M e 
q
 

where  is aircraft AOA, q is aircraft pitch rate,  e is elevator deflection (control


input), V is the trimmed (constant) airspeed, and Z , Z q , Ze  
and M  , M q , M e  are

partial derivatives of the aerodynamic vertical force Z and pitching moment M with

respect to  , q,  e  respectively. The same numerical values for the vehicle aerodynamic

derivatives were employed as in Chapter 3, with resulting open-loop system matrices:

118
 1.0189 0.9051   0.0022 
A , B  (5.50)
 0.8223 1.0774   0.1756 

where  is in radians, q is in radians/second, and  e is in degrees.

Baseline flight control is designed for the baseline system without uncertainties.

Similar to previous chapters, we define the system controlled output to be the vehicle

AOA:

 
y    1 0    (5.51)
C
q

Augmenting the system dynamics with the integrated tracking error following the

procedure shown in (5.11)-(5.15) and designing baseline control (5.18) via LQR methods

with weighting matrices:

Q  diag 100 0 0 , R 1 (5.52)

yields baseline LQR feedback gains:

K xT  10 10.8786 6.0589  (5.53)

Baseline feedback control gains (5.53) lead to the reference model (5.19) that will be

used in adaptive augmentation designs. The eigenvalues of the reference dynamics along

with their corresponding natural frequencies are shown below:

1  0.613  0.67 j 1  0.675, 1  0.907


 
2  0.613  0.67 j   2  0.675, 2  0.907 (5.54)
  1.96   1,   1.96
 3  3 3

The baseline system response to a series of angle-of-attack command doublets is shown

in Figure 33. The response shows that in the absence of uncertainties, the baseline system

119
tracks the AOA doublet commands while maintaining elevator deflection and rate that are

well within acceptable limits.

10
cmd
, deg.

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

1
e, deg

0.5

-0.5

-1
0 10 20 30 40 50 60 70 80

2
, dps

1
dot

0
e

-1
0 10 20 30 40 50 60 70 80

Time, sec

Figure 33: Baseline Closed-Loop Response to AOA Doublet without Uncertainties

Including system uncertainties yields dynamics in the form of (5.17):

120
 
 0 1 0   0    0 
 eyI        eyI     1 (5.55)
    Z e     
  
 
Z Z q   Z e 

    0 1 


 V 
 0  K xTpert  
    V   e    d     0   cmd
 
 q   V V     0
 q  
  M q   M  e   M    
u
K dT
 0 M   x  e 
x
  Bc
 Ak  B

Similar to Chapter 3, three types of matched uncertainties are added to the system: 1)

linear-in-state uncertainty K Tpert x , 2) control effectiveness constant uncertainty   0 ,

and 3) nonlinear-in-state uncertainty in the form of (5.2). Numerical values for the

uncertainties were chosen as:

 c 2

K xTpert   4.6839 9.8197  ,   0.5, d  x p   d    0.5 e 2 2
(5.56)

where the center of the Gaussian was set to  c  2  /180 and its width was   0.0233 .

This particular selection of numerical values for K xTpert and  is equivalent to 50%

increase in the static instability M  , 80% decrease in the pitch damping M q , and 50%

decrease in the control input effectiveness. These changes imply that the vehicle became

50% more statically unstable, lost 80% of its pitch damping ability, and the aircraft

controllability decreased by 50%. In fact, this combination causes the open-loop system

to become unstable (eigenvalues enter the right-half plane), and in addition there is the

simultaneous injection of a significant nonlinear-in-state uncertainty. Such drastic

changes were motivated by intent to demonstrate the effectiveness of the proposed design

methodology. This particular example was also selected to be similar to previous work

presented in [31], [32] as well as in Chapter 3 so that relevant performance comparisons

may be available.

121
With only the baseline controller in operation and with the uncertainties included, the

closed-loop system tracking performance degradation can be clearly observed from the

data that are shown in Figure 34. Although the tracking performance is poor, both the

control input and its rate remain small.

10
cmd
, deg.

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

4
e, deg

-2

-4
0 10 20 30 40 50 60 70 80

2
, dps

0
dot

-1
e

-2
0 10 20 30 40 50 60 70 80

Time, sec

Figure 34: Baseline Closed-Loop Response in Presence of Matched Uncertainties

122
To counteract the effects of the uncertainties, MRAC adaptive laws (5.31) were

constructed by solving algebraic Lyapunov equation (5.32) with reference matrix Aref as

designed via (5.53) and with

Qref  diag 0.1 1 800 (5.57)

The regressor vector  d   consisted of 11  -dependent and evenly spaced Gaussians

(RBFs). RBF centers were placed at [-10:2:10] degrees of AOA, and all RBF widths were

set to   0.0233 . The regressor matrix u  s  consisted of 11 1st-order Padé

approximations of input time-delays with magnitudes of [0:0.02:0.2] seconds. The rates

of adaptation were chosen to be:

 x  diag 1 200 200 , d  20, u  0.5 (5.58)

and the total control (elevator deflection) was formed as shown in (5.20). In order to

improve robustness, projection and dead-zone modifications are included in the adaptive

laws. With the baseline + MRAC adaptive augmentation control active, closed-loop

system performance was recovered with acceptable elevator deflection and rates, as

shown in Figure 35.

123
10
cmd
, deg. 5

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

4
e, deg

-2
0 10 20 30 40 50 60 70 80

5
, dps

0
dot

-5
e

-10
0 10 20 30 40 50 60 70 80

Time, sec

Figure 35: Baseline+MRAC Closed-Loop Response in Presence of Uncertainties

The norms of the MRAC adaptive parameters are shown in Figure 36. The effect of

the dead-zone is seen in the adaptive parameter response, where once the tracking error

becomes small the adaptive parameters stop evolving and remain constant.

124
5

4
Kx hat 3

0
0 10 20 30 40 50 60 70 80

2.5

1.5
d hat

0.5

0
0 10 20 30 40 50 60 70 80

0.2

0.15
u hat

0.1

0.05

0
0 10 20 30 40 50 60 70 80
Time, sec

Figure 36: MRAC Adaptive Parameter Norms vs. Time

As previously discussed, if the adaptive gains are too high in the classical MRAC

design, poor performance characteristics may occur. Increasing the gains to:

d  diag 1 200 200 , d  40, u  5,   8 (5.59)

125
that is, setting gains  d and  u to twice and ten times as high as in the original MRAC

design, yields the response shown in Figure 37. While the tracking response appears

acceptable, inspection of the control input positions and rates show unacceptable

characteristics and indicate the stability of the system may be tenuous.

10
cmd
, deg.

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

10
e, deg

-5
0 10 20 30 40 50 60 70 80

500
, dps

0
dot

-500
e

-1000
0 10 20 30 40 50 60 70 80

Time, sec

Figure 37: Baseline+MRAC Closed-Loop Response with High Adaptive Gains

126
In order to improve the performance of the baseline + adaptive augmentation system,

we next implement the alternative design featuring normalization developed in (5.35)-

(5.48). The rates of adaptation are selected to be those defined in (5.59), where the

selection of  d is chosen to be similar to  x in the MRAC case for comparison, while

gains  d and  u are twice and ten times as high as the original MRAC case for the

alternative design. We note that this gain set showed poor performance characteristics in

the classical MRAC case. Increasing rates of adaptation leads to oscillations and

instability in general, yet we will show that the normalized design is able to facilitate

larger rates of adaptation while still improving performance by reducing oscillations. The

filter in (5.40) is selected to be a single pole low-pass filter:

20
G s  (5.60)
s  20

We note that while the Padé approximations may not accurately represent the phase loss

of the pure time-delay (as illustrated in Figure 32), as previously stated the design of filter

(5.60) allows the ability to shape the measured quantities to avoid known frequency

regions associated with uncertain dynamics, noise, and/or disturbances. More rigorous

investigation of the filter design and its relationship to the effects of time-delay remains a

future research area. With the baseline + normalized adaptive augmentation control

active, closed-loop system performance was again recovered with acceptable elevator

deflection and rates, as shown in Figure 38.

127
10
cmd
, deg. 5

0
ref

-5

-10
0 10 20 30 40 50 60 70 80

4
e, deg

-2
0 10 20 30 40 50 60 70 80

4
, dps

0
dot

-2
e

-4
0 10 20 30 40 50 60 70 80

Time, sec

Figure 38: Baseline+B-SPM Normalized Response in Presence of Uncertainties

The norms of the B-SPM normalized adaptive parameters are shown in Figure 39.

Once again, the effect of the dead-zone is seen in the adaptive parameter response, where

once the tracking error becomes small the adaptive parameters stop evolving and remain

128
constant. The adaptive norms show smoother evolution than the data from the MRAC

design.

2
Kd hat

0
0 10 20 30 40 50 60 70 80

2
d hat

0
0 10 20 30 40 50 60 70 80

0.4
u hat

0.2

0
0 10 20 30 40 50 60 70 80

1.5
 hat

0.5
0 10 20 30 40 50 60 70 80
Time, sec

Figure 39: Baseline+B-SPM Adaptive Parameter Norms vs. Time

129
To further demonstrate the advantage of the alternative adaptive design, we compare

the responses of the two designs. As shown in Figure 40, the alternative design shows

smoother elevator transients and does not contain unwanted high frequency oscillations.

10
cmd
, deg.

0
 MRAC
 Alt
-5

-10
0 10 20 30 40 50 60 70 80

4
e, deg

-2
0 10 20 30 40 50 60 70 80

5
, dps

0
dot

-5
e

-10
0 10 20 30 40 50 60 70 80

Time, sec

Figure 40: Comparison of Adaptive Closed-Loop Responses: with Uncertainties

130
Another way of quantifying control input activity is via the frequency domain by

plotting the power spectral densities (PSDs) of the elevator inputs. As shown in Figure

41, the alternative adaptive design displays less control input and does not contain the

high frequency control input components present in the MRAC response.

20
MRAC
Alternative

-20

-40
Elev. Pos.(deg2/Hz dB)

-60

-80

-100

-120

-140
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Frequency, Hz

Figure 41: Comparison of Control Input Power Spectral Densities

131
In order to more formally quantify the performance improvements provided by the

alternative adaptive design scheme, we investigate the effects of time delays on the

control input signal using each of the adaptive schemes. The input time delay margin of

the baseline LQR system can be obtained through linear analysis. Many attempts have

been made to quantify time delay margins in adaptive systems [35]-[37], but only

recently have theoretical results been established that can determine the time delay

margin of such nonlinear systems [38]. These recent results are only applicable for single

input state feedback MRAC systems, and the vector (MIMO) case remains an open

problem. Therefore in this work the time delay margins of the adaptive controllers will be

found numerically through simulation. We include a realistic 2nd order actuator model

with parameters:

act  60 rad/sec, act  0.7 (5.61)

Using LTI analysis, the phase margin of the baseline LQR system defined by (5.53)

and actuator (5.61) without uncertainties at the plant input is 66.56 degrees at a crossover

frequency at 1.24 rad/sec corresponding to a time-delay margin of approximately 933ms.

This result is verified as shown in Figure 42, which shows baseline system response

when time-delay is set to 935ms in simulation, with the response just becoming unstable.

132
Time-Delay = 0.935
10
cmd
, deg. 5

0 ref

-5 

-10
0 10 20 30 40 50 60 70 80

2
e, deg

-1

-2
0 10 20 30 40 50 60 70 80

4
, dps

0
dot

-2
e

-4
0 10 20 30 40 50 60 70 80

Time, sec

Figure 42: Baseline Response: Input Time-Delay of 935ms

It is expected that the adaptive controllers will decrease the time-delay margin while

simultaneously adding improved performance under uncertainty. To determine the time-

delay margins of the adaptive designs, the same AOA doublets are used for tracking and

133
the uncertainties are removed from the simulation and the same gains, projection bounds,

and adaptive laws are used. The input time-delay is increased until instability occurs. The

baseline + MRAC design is able to remain stable until a time-delay of 30ms, or only

roughly 3% of the baseline system. This is a significant reduction in system performance

due to the addition of classical MRAC augmentation.

The baseline + B-SPM normalized adaptive augmentation system response with an

input time-delay of 935ms is shown in Figure 43. The response shows the B-SPM

normalized design able to retain acceptable AOA excursions with stability. In effect, the

B-SPM system is able to match the time-delay characteristics of the baseline LQR design,

but with the significantly improved performance in the presence of uncertainties shown in

previously.

134
Time-Delay = 0.935
10
cmd
, deg. 5

0 ref

-5 

-10
0 10 20 30 40 50 60 70 80

2
e, deg

-1

-2
0 10 20 30 40 50 60 70 80

4
, dps

0
dot

-2
e

-4
0 10 20 30 40 50 60 70 80

Time, sec

Figure 43: Baseline + B-SPM Normalized Response: Input Time-Delay of 935ms

135
The time-delay results are summarized below, with the phase margin calculated by

using the baseline closed-loop crossover frequency:

Corresponding
Time-Delay Margin
Control Scheme Phase Margin
(msec)
(deg)
Baseline LQR 933 66.29
LQR + MRAC 30 2.13
LQR + B-SPM 933 66.29

The same simulation analyses were performed for AOA maneuvers with magnitude of

one degree and generated similar results, which indicates that the adaptive design

approach developed here is scalable.

136
Chapter 6: Concluding Remarks and Suggestions for Future Work

In this dissertation, we focus on the development of robust baseline and adaptive

control augmentation designs for general classes of dynamic systems, with results that

address known challenges in flight dynamics and aerospace applications. Our design

philosophy combines well-known modern control techniques to provide robust baseline

control design with augmentation-based adaptive control and state limiting to retain

system performance in off-nominal and degraded conditions. Throughout, we present

rigorous analysis to theoretically justify the approach while using application to relevant

real world aerospace examples to illustrate the benefit of our developments. We believe

the contributions of this work are the following:

 Use of modern control techniques (dynamic inversion, LQR) for baseline design

provides robust system performance under nominal conditions, with well known

LTI analysis tools to quantify robustness and system performance.

 Theoretically justified state limiting augmentation combined with robust dynamic

inversion design satisfies performance specifications while keeping the system

within a prescribed state space. In practice such a design allows an aircraft to

meet robustness and performance specifications while staying within its flight

envelope. Application to a flight test program illustrated the efficacy of the

design, while also receiving positive pilot feedback.

 Adaptive augmentation combined with robust dynamic inversion provides a

method to retain desired system performance in the presence of parametric

uncertainties. Additional robustness modifications are employed to address non-

137
parametric uncertainties, resulting in a system that provides performance both

under nominal and degraded conditions. Application to representative flight

dynamics illustrated the design is able to retain desired closed-loop flight

dynamics even in the presence of severe uncertainty and system degradation.

 Combining state limiting and adaptive augmentation with robust dynamic

inversion effectively combines the benefits of the three design components.

Application to representative flight dynamics illustrated the design is able to

satisfy robustness and performance specifications under nominal conditions, keep

the aircraft within its prescribed flight envelope, and retain desired closed-loop

dynamics in the presence of severe uncertainty and system degradation.

 Use of LQR optimal baseline design provides attractive guaranteed stability and

robustness properties. Again, use of classical model reference adaptive control

augmentation provides a method to retain nominal closed-loop performance in the

presence of uncertainty. However, it is known that increased adaptive gains can

lead to oscillation and instability. In this work, we presented an alternative design

approach using bilinear static parameter model adaptive design with modification

to specifically address input time delay. The resulting design is able to retain

desired closed-loop dynamics in the presence of severe uncertainty and system

degradation, with the additional benefit of significantly improved time-delay

margin as illustrated via representative flight dynamics.

138
There are a number of worthwhile potential extensions to the research presented in

this thesis. In Chapter 2 we designed state limiting augmentation for application to 2 nd

order cascaded system dynamics often found in flight dynamics. A future extension of

this work could generalize the state limiting augmentation to a more generic class of

dynamic systems, such as that employed in Chapter 5. Some work has been accomplished

in this regard [39], but more investigation is warranted.

In the B-SPM design presented in Chapter 5, a stable strictly proper filter is used to

filter the system dynamics as part of the adaptive design. Future work should focus on

selection of this filter and the corresponding impact on system performance and time-

delay margin. Some recent work on employing closed-loop reference models [40], [34]

approaches this problem from another direction and shows promising results.

The adaptive augmentation schemes presented are able to recover desired closed-loop

performance, and in the B-SPM design show significantly improved performance under a

class of dynamic input uncertainties. However, in all cases the time-delay margin is

numerically derived without a complete analytic solution. Recent work [38] has provided

analytic derivation of time-delay margin, but the results are only applicable for single

input state feedback MRAC systems. Future work should extend such results to the

MIMO case to facilitate more practical applications.

In a larger sense, a more thorough robustness analysis needs to be developed for the

adaptive augmentation schemes presented in this thesis. Both the classical MRAC design

derived in Chapter 3 and the B-SPM design derived in Chapter 5 contain modifications

designed to improve robustness to non-parametric uncertainties. Chapter 3 includes

139
analysis of the estimation error and corresponding UUB properties of the closed-loop

system, including both static parametric and bounded non-parametric uncertainties, as

seen in the uncertainty definition:

D  x    D*   D  x    D  x 
T
(6.1)

In this case, we derived the ultimate bound and found that it depended on the size of the

estimation error, but did not do additional robustness analyses. The B-SPM design in

Chapter 5 assumes both the static and dynamic parametric uncertainties may be perfectly

matched, without considering non-parametric uncertainties such as estimation error:


x p  Ap x p  Bp   I mm  Tu u  s   u  Td  d  x p   (6.2)

Addition of such estimation errors complicates the stability and robustness results,

and while some generalized results exist [7], these results need to be extended and

applied to the aerospace examples presented in this work. Future work should expand the

analysis to first include static and dynamics non-parametric uncertainties:


x p  Ap x p  Bp   I mm    s   u  Td  d  x p   (6.3)

where   s  and Td  d  x p  may not be perfectly estimated. Beyond this case, future

work should include unmatched dynamic uncertainties. For example, consider the

dynamics

 
x p  Ap x p  Bp   I mm    s   u  Td  d  x p     t  (6.4)

where   t  is a bounded function of time which it is assumed cannot destroy the

controllability of the system. The derivation of robustness properties similar to those

140
known for existing LTI designs (gain and phase margin, Nyquist margins, etc.) will be

crucial for increased certification and implementation of robust adaptive control schemes

in aerospace applications.

141
Bibliography

[1] Wise, K. A., Lavretsky, E., Hovakimyan, N., “Adaptive Control of Flight: Theory,
Applications, and Open Problems,” Proceedings of American Control Conference,
Minneapolis, MN, 2006. doi: 10.1109/ACC.2006.1657677

[2] Young, A., Cao, C., Hovakimyan, N., and Lavretsky, E., “An Adaptive Approach to
Nonaffine Control Design for Aircraft Applications,” Guidance, Navigation, and
Control Conference, AIAA Paper 2006-6343, Keystone, CO, 2006.

[3] Lavretsky, E., and Wise, K. A., “Adaptive Flight Control for Manned / Unmanned
Military Aircraft,” Proceedings of the American Control Conference, Portland, OR,
2005.

[4] Stevens, B., Lewis, F., Aircraft Control and Simulation, Wiley, New York, 2003,
pp. 206, 210, 339, 402.

[5] Dydek, Z. T., Annaswamy, A. M., and Lavretsky, E., “Adaptive Control and the
NASA X-15-3 Flight Revisited,” IEEE Control Systems Magazine, vol. 1066, no.
033X/10, 2010.

[6] Narendra, K. S. and Annaswamy, A. M., Stable Adaptive Systems, Dover, New
York, 2005.

[7] Ioannou, P. A., and Sun, J., Robust Adaptive Control, Prentice Hall, Upper Saddle
River, NJ, 1996.

[8] Khalil, H. K., Nonlinear Systems, Prentice Hall, Upper Saddle River, NJ, 2002.

[9] Karason, S. P., and Annaswamy, A. M., “Adaptive Control in the Presence of Input
Constraints,” IEEE Transactions on Automatic Control, vol. 39, no. 11, p. 2325-
2330, November 1994.

[10] Dydek, Z. T., Jain, H., Jang, J., Annaswamy, A. M., and Lavretsky, E.,
“Theoretically Verifiable Stability Margins for an Adaptive Controller,” Guidance,
Navigation, and Control Conference, AIAA Paper 2006-6416, Keystone, CO, 2006.

[11] Jang, J., Annaswamy, A. M., and Lavretsky, E., “Towards Verifiable Adaptive
Flight Control in the Presence of Actuator Anomalies,” Proc. Conference on
Decision and Control, San Diego, CA, December 2006, p 3300-3305.

[12] Tao, G., Adaptive Control Design and Analysis, John Wiley & Sons, Hoboken, NJ,
2003.

142
[13] Krstic, M, Kokotovic, P., and Kanellakopoulos, I., Nonlinear and Adaptive Control
Design, John Wiley & Sons, New York, 1995.

[14] Ioannou, P. A., and Tsakalis, K., “A Robust Direct Adaptive Controller,” IEEE
Transactions on Automatic Control, vol. 31, no. 11, p. 1033-1043, November 1986.

[15] Sastry, S., and Bodson, M., Adaptive Control: Stability, Convergence, and
Robustness, Prentice Hall, Upper Saddle River, NJ, 1989.

[16] Egardt, B., and Whitacre, D., Stability of Adaptive Controllers, Springer-Verlag,
NJ, 1979.

[17] Åström, K., and Wittenmark, B., Adaptive Control, Addison-Wesley, Boston, 1995.

[18] Brinker, J., and Wise, K., “Flight Testing of Reconfigurable Control Law on the X-
36 Tailless Aircraft,” Journal of Guidance, Control, and Dynamics, vol. 24, no. 5,
p. 903-909, 2001.

[19] Sharma, M., Lavretsky, E., and Wise, K., “Application and Flight Testing of an
Adaptive Autopilot on Precision Guided Munitions,” Guidance, Navigation, and
Control Conference, Keystone, CO, 2006.

[20] Lavretsky, E., and Gadient, R., “Robust Adaptive Design for Aerial Vehicles with
State Limiting Constraints,” Journal of Guidance, Control, and Dynamics, Vol. 33,
No. 6, November-December, 2010, pp. 1743-1752. doi: 10.2514/1.50101

[21] Sanner, R., and Slotine, J.-J. E., “Gaussian Networks for Direct Adaptive Control,”
IEEE Transactions on Neural Networks, Vol. 3, No. 6, 1992, pp. 837-864. doi:
10.1109/72.165588

[22] Lim, Y. S., and Kazda, L. F., “A Study of Second Order Nonlinear Systems,”
Journal of Mathematical Analysis and Applications, Vol. 8, No. 3, pp. 423-444,
June 1964.

[23] Regan, C. D., “In-Flight Stability Analysis of the X-48B Aircraft,” Atmospheric
Flight Mechanics Conference, AIAA Paper 2008-6571, Honolulu, HI, 2008.

[24] Risch, T., Cosentino, G., Regan, C. D., Kisska, M., and Princen, N., “X-48B Flight-
Test Progress Overview,” Aerospace Sciences Meeting, AIAA Paper 2009-934,
Orlando, FL, 2009.

[25] Goldthorpe, S. H., Rossitto, K. F., Hyde, D. C., and Krothapalli, K. R., “X-48B
Blended Wing Body Flight Test Performance of Maximum Sideslip and High to

143
Post Stall Angle-of-Attack Command Tracking,” Atmospheric Flight Mechanics
Conference, AIAA Paper 2010-7514, Toronto, ON, 2010.

[26] Hyde, D. C., Gadient, R., and Lavretsky, E., “Flight Testing the X-48B Angle-of-
Attack and Sideslip Limiting System,” Guidance, Navigation, and Control
Conference, Portland, OR, 2011.

[27] Anon., “U.S. Naval Test Pilot School Flight Test Manual, Fixed Wing Stability and
Control Theory and Flight Test Techniques,” U.S. Naval Test Pilot School,
USNTPS-FTM-103, Patuxent River, MD, Jan 1997.

[28] Ioannou, P. A., and Fidan, B., Adaptive Control Tutorial, Society for Industrial and
Applied Mathematics, Philadelphia, 2006, pp. 31-32.

[29] Park, J. and Sandberg, I., “Universal Approximation Using Radial-Basis-Function


Networks,” Neural Computation, Vol. 3, No. 2, 1991, pp. 246-257.

[30] Pomet, J. B., and Praly, L., “Adaptive Nonlinear Regulation: Estimation from the
Lyapunov Equation,” IEEE Transactions on Automatic Control, Vol. 37, No. 6,
1992, pp. 729-740.

[31] Lavretsky, E., Gadient, R., and Gregory, I. “Predictor-Based Model Reference
Adaptive Control,” Journal of Guidance, Control, and Dynamics, Vol. 33, No. 4,
July-August, 2010, pp. 1195-1201. doi: 10.2514/1.46849

[32] Lavretsky, E., “Combined/Composite Model Reference Adaptive Control,” IEEE


Transactions on Automatic Control, Vol. 54, No. 11, 2009, pp. 2692-2697. doi:
10.1109/TAC.2009.2031580

[33] Dorf, R. C., Bishop, R. H., Modern Control Systems, Prentice Hall, Upper Saddle
River, NJ, 2011, pp. 671-673, 692.

[34] Lavretsky, E., Wise, K. A., Robust and Adaptive Control: With Aerospace
Applications, Springer, 2012.

[35] Lavretsky, E., Annaswamy, A. M., Dydek, Z. T., Vega-Brown, W., “On the
Computation of Stability Margins for Adaptive Controllers using Linear System
Tools,” AIAA Guidance, Navigation, and Control Conference, Chicago, IL, 2009.

[36] Nguyen, N. T., Ishihara, A. K., Krishnakumar, K. S., Baktiari-Nejad, M., “Bounded
Linear Stability Analysis – A Time Delay Margin Estimation Approach for
Adaptive Control,” AIAA Guidance, Navigation, and Control Conference, Chicago,
IL, 2009.

144
[37] Annaswamy, A. M., Jang, J., Lavretsky, E., “Stability Margins for Adaptive
Controllers in the Presence of Time-Delay,” AIAA Guidance, Navigation, and
Control Conference, Honolulu, HI, 2008.

[38] Annaswamy, A. M., Lavretsky, E., Dydek, Z. T., Gibson, T. E., Matsutani, M.,
“Recent Results in Robust Adaptive Flight Control Systems,” International Journal
of Adaptive Control and Signal Processing, Special Issue: Robust Adaptive
Control: Legacies and Horizons, Vol. 27, Issue 1-2, January-February 2013, pp. 4-
21.

[39] Muse, J. A., “A Method for Enforcing State Constraints in Adaptive Control,” AIAA
Guidance, Navigation, and Control Conference, Portland, OR, 2011.

[40] Lavretsky, E., “Adaptive Output Feedback Design Using Asymptotic Properties of
LQG/LTR Controllers,” IEEE Transactions on Automatic Control, Vol. 57, No. 6,
2012, pp. 1587-1591.

145

You might also like