applications of linear parameter...

78
APPLICATIONS OF LINEAR PARAMETER-VARYING CONTROL FOR AEROSPACE SYSTEMS By KRISTIN LEE FITZPATRICK A THESIS PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF MASTER OF SCIENCE UNIVERSITY OF FLORIDA 2003

Upload: doanhuong

Post on 10-May-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

APPLICATIONS OF LINEAR PARAMETER-VARYING CONTROL FORAEROSPACE SYSTEMS

By

KRISTIN LEE FITZPATRICK

A THESIS PRESENTED TO THE GRADUATE SCHOOLOF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OFMASTER OF SCIENCE

UNIVERSITY OF FLORIDA

2003

TABLE OF CONTENTSpage

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

CHAPTER

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

2 LINEAR PARAMETER-VARYING CONTROL THEORY . . . . . . . . . . 5

3 LINEAR PARAMETER-VARYING CONTROL FOR AN F/A-18 . . . . . . . 10

3.1 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103.2 Open-loop Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 123.3 Control Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143.4 Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153.5 Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

4 LINEAR PARAMETER-VARYING CONTROL FOR A HYPERSONICAIRCRAFT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

4.1 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224.2 Generic Hypersonic Vehicle . . . . . . . . . . . . . . . . . . . . . . . 254.3 Hypersonic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264.4 Linear Parameter-Varying System . . . . . . . . . . . . . . . . . . . . 274.5 Control Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304.6 Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

4.6.1 Open-Loop Simulation . . . . . . . . . . . . . . . . . . . . . . 334.6.2 Closed-Loop Simulation . . . . . . . . . . . . . . . . . . . . . 35

4.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

5 LINEAR PARAMETER-VARYING CONTROL FOR A DRIVEN CAVITY . 39

5.1 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395.2 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

ii

5.3 Driven Cavity Geometry . . . . . . . . . . . . . . . . . . . . . . . . . 415.4 Governing Equations of Motion . . . . . . . . . . . . . . . . . . . . . 425.5 Reduced-Order Linear Dynamics . . . . . . . . . . . . . . . . . . . . . 445.6 Creeping Flow in a Driven Cavity . . . . . . . . . . . . . . . . . . . . 475.7 Excitation Phase Differential . . . . . . . . . . . . . . . . . . . . . . . 495.8 Control Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

5.8.1 Control Objectives . . . . . . . . . . . . . . . . . . . . . . . . . 505.8.2 Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

5.9 Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 545.9.1 Open-Loop Simulation . . . . . . . . . . . . . . . . . . . . . . 545.9.2 Reduced-Order Closed-Loop Simulation . . . . . . . . . . . . . 575.9.3 Full-Order Closed-Loop Simulation . . . . . . . . . . . . . . . 60

5.10 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

6 CONCLUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

iii

LIST OF TABLESTable page

3–1 Original Design Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

3–2 Frequency and Damping Ratio of Design and Analysis Models . . . . . . . 14

3–3 Frequencies and Damping Ratios of the Target Model . . . . . . . . . . . . 16

3–4 Induced Norms of Closed-Loop System . . . . . . . . . . . . . . . . . . . 17

3–5 Numerical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

4–1 Model Dimensions and Flight Conditions . . . . . . . . . . . . . . . . . . 27

4–2 Modes of the Hypersonic Model . . . . . . . . . . . . . . . . . . . . . . . 30

4–3 Modes of the Target Model . . . . . . . . . . . . . . . . . . . . . . . . . . 32

4–4 Open-Loop Synthesis Norms . . . . . . . . . . . . . . . . . . . . . . . . . 33

4–5 Point Design Norms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

5–1 Induced Norms of Closed-Loop System . . . . . . . . . . . . . . . . . . . 53

iv

LIST OF FIGURESFigure page

2–1 H∞ Block Diagram (Gain-Scheduled) . . . . . . . . . . . . . . . . . . . . 7

3–1 F/A-18 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

3–2 Flight Envelope/Parameter Space . . . . . . . . . . . . . . . . . . . . . . . 12

3–3 Synthesis Block Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

3–4 Closed-loop Block Diagram . . . . . . . . . . . . . . . . . . . . . . . . . 18

3–5 Pitch Rate for Design Points . . . . . . . . . . . . . . . . . . . . . . . . . 18

3–6 Pitch Rate for Analysis Point . . . . . . . . . . . . . . . . . . . . . . . . . 19

3–7 Controller Elevator Deflection . . . . . . . . . . . . . . . . . . . . . . . . 19

4–1 Inner-Loop/Outer-Loop Design . . . . . . . . . . . . . . . . . . . . . . . . 25

4–2 Simplified Model of a Generic Hypersonic Vehicle . . . . . . . . . . . . . 25

4–3 Synthesis Block Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

4–4 Open-Loop Transfer Functions . . . . . . . . . . . . . . . . . . . . . . . . 34

4–5 Open-Loop Angle of Attack Result . . . . . . . . . . . . . . . . . . . . . 35

4–6 Input Elevon Deflection . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4–7 Closed-loop Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4–8 Closed-Loop Transfer Functions . . . . . . . . . . . . . . . . . . . . . . . 36

4–9 Closed-Loop Angle of Attack Result . . . . . . . . . . . . . . . . . . . . . 37

4–10 Elevon Deflection Command . . . . . . . . . . . . . . . . . . . . . . . . . 38

4–11 Elevon Deflection Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

5–1 Stokes Driven Cavity Flow Problem . . . . . . . . . . . . . . . . . . . . . 41

5–2 Controller Block Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . 51

5–3 Open-Loop Flow Velocities for Full-Order Model . . . . . . . . . . . . . . 55

v

5–4 Open-Loop Flow Velocities for Reduced-Order Model with 165o PhaseDifferential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

5–5 Open-Loop Flow Velocities for Reduced-Order Model with 210o PhaseDifferential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

5–6 Trajectory of Phase Differential . . . . . . . . . . . . . . . . . . . . . . . 56

5–7 Open-Loop Flow Velocities for Reduced-Order Model over a Trajectoryof Phase Differentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

5–8 Closed-loop System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

5–9 Closed-Loop Flow Velocities for Reduced-Order Model with 165o PhaseDifferential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

5–10 Closed-Loop Flow Velocities for Reduced-Order Model with 210o PhaseDifferential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

5–11 Closed-Loop Flow Velocities for Reduced-Order Model over a Trajec-tory of Phase Differentials . . . . . . . . . . . . . . . . . . . . . . . . . 60

5–12 Closed-Loop Flow Velocities for Full-Order Model . . . . . . . . . . . . . 61

5–13 Closed-Loop Flow Velocities for Full-Order Model with Controller As-sociated with 165o Phase Differential . . . . . . . . . . . . . . . . . . . 61

5–14 Closed-Loop Flow Velocities for Full-Order Model with Controller As-sociated with 210o Phase Differential . . . . . . . . . . . . . . . . . . . 62

5–15 Closed-Loop Flow Velocities for Full-Order Model over a Trajectory ofPhase Differentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

vi

Abstract of Thesis Presented to the Graduate Schoolof the University of Florida in Partial Fulfillment of the

Requirements for the Degree of Master of Science

APPLICATIONS OF LINEAR PARAMETER-VARYING CONTROL FORAEROSPACE SYSTEMS

By

Kristin Lee Fitzpatrick

December 2003

Chair: Richard C. Lind, Jr.Major Department: Mechanical and Aerospace Engineering

Gain-scheduling control has been an engineering practice for decades and can

be described as the linear regulation of a system whose parameters are changed as a

function of varying operating conditions. Several gain-scheduling techniques have been

researched for the control of systems that vary with time-varying parameters. These

techniques create controllers at various points within the parameter space of the system

and use an interpolation law to change controllers as the parameter changes with time.

The process of creating such an interpolation law can be very rigorous and time-

consuming and the resulting controller is not guarnateed to stablize the time-varying

system. The gain-scheduling technique known as linear parameter-varying control,

however, solves a linear matrix inequality convex problem to create a single controller

that has an automatic interpolation law and is guaranteed to stabilize the closed-loop

system. This paper demonstrates the use of this technique to create controllers for three

aerospace systems. The first system is the longitudinal dynamics of the F/A-18, the

second system is the structural dynamics of a hypersonic vehicle and the third system

is the flow dynamics within a driven cavity. Simulations are performed using the linear

vii

parameter-varying controller created for each system to show the usefulness of a linear

parameter-varying framework as a gain-schedule design technique.

viii

CHAPTER 1INTRODUCTION

1.1 Overview

The dynamics of aerospace systems that deal with flight and process control

are affected by variations in the parameters that make up their operating space (i.e.,

altitude, Mach number, temperature). Gain-scheduling techniques are used to create

a controlling scheme that will work throughout the system’s operating space. The

resulting controller will vary based on the same parameters as the system’s plant

model.

The traditional gain-scheduling technique can be broken down into three major

steps. The first step involves separating the operating range into subspaces and creating

a parameterized model of each subspace. In the second step, controllers are created

for each of these models. Finally, in the third step, a scheduling scheme is devised by

linearly interpolating between these regional controllers as the vehicle moves through

its operating range. This technique works well for some systems; however, it does not

guarantee stability and robustness of the closed-loop system. Another disadvantage

of this method is the possibility of a skipping behavior due to the switch between

controllers.

This thesis presents the technique of creating a single gain-scheduled controller

that can be treated as a single entity. This technique achieves gain-scheduling with a

parameter-dependent controller that will work throughout an operating range or flight

envelope. Benefits of this technique are that it removes the need of creating several

controllers for different parameters within an operating domain and removes the need

for the creation of a gain-scheduled control law.

1

2

This study applies this method to three specific aerospace systems. The first

application of this technique is creation of a controller for the F/A-18 longitudinal

axes over a specific flight envelope. The second application is control of the structural

dynamics of a hypersonic aircraft over a temperature range. The third application is

control of the velocity along the center of a driven cavity flow over a range of phase

differentials within the flow.

1.2 Background

Gain-scheduling has evolved hand in hand with the progress of mechanical

systems. Gain-scheduling techniques are currently used for control design for both

linear and nonlinear systems. This study focused on systems pertaining to aerospace

applications.

Gain-scheduling in aerospace applications came about during WWII as autopi-

loting became necessary with the birth of jet aircraft and guided missiles [1]. When

gain-scheduling was first conceived for the military, it was created through hardware

and was quite costly. Gain-scheduling was not adopted commercially until the cre-

ation of digital control, nearly a quarter-century after military use. The development

of gain-scheduling over past decades led to several design techniques and the use of

gain-scheduling for many different aerospace systems.

Several gain-scheduling methods have been developed for designing controllers for

linear systems. The three main classes of linear systems that apply to aerospace sys-

tems are linear time-invariant (LTI), linear time-varying (LTV), and linear parameter-

varying (LPV). Gain-scheduling is most often applied to linear parameter-varying

systems, which are affine functions of parameters that affect their operation.

Gain-scheduling can also be applied to the control of nonlinear systems. Several

linearization techniques can be used for nonlinear systems before a gain-scheduled

controlling scheme is developed. The most common approach is based on Jacobian

linearization of the nonlinear plant about a family of operating points (i.e., equilibrium

3

points) [2]. The system can also be linearized along a trajectory in the event the lin-

earized dynamics do not exhibit good performance or stability away from equilibrium

points [3]; however, the trajectory must be known in advance to perform the control

design. Once the control scheme is created, it can be applied to controlling the non-

linear system. Simulations of gain-scheduling controllers have also been applied to

nonlinear systems [4, 5, 6]; however, nonlinear systems do not necessarily have to be

linearized. Set-valued methods for LPV systems have also been applied to nonlinear

systems with quasi-LPV representations [7]. Linearization errors were accommodated

as linear state-dependent disturbances; constraints on systems’ states and controls were

specified; rates of transitions among operating regions were addressed, which allows

even the local-point designs to be nonlinear.

The classical gain-scheduling approach is to create a number of controllers within

the operating domain; and then, using a scheduling scheme, to switch between them as

the system parameters change. One method that uses this approach was demonstrated

for a missile autopilot that uses µ synthesis with D-K iteration to create controllers;

and an iteration scheme is designed over the operating domain [8]. Another method

for a missile autopilot creates controllers at distinct operating conditions using H∞

control synthesis; and then creates a schedule for the controllers by removing coupling

terms [9]. Another project involved creating H∞ point design controllers at specific

equilibrium points [10]. That project reduces the controllers to second order which

are then realized in a feedback path configuration for which a gain-scheduling law is

developed. A study also used a design algorithm for a state feedback law based on

gain-scheduling for an LPV multi-input multi-output system [11]. The state feedback

control law places the system’s poles in a neighborhood of desired locations and

stabilizes the closed-loop system. Though this classical approach has worked well for

many applications, there is no guarantee of robustness or stability of the closed-loop

systems.

4

A more recent approach, that appeared in the late 1990’s, involves creating an

LPV controller that uses an automatic interpolation law over the operating domain

(which has guaranteed closed-loop robustness and stability with the LPV system).

The method of D-K iteration with µ synthesis was used to create an LPV controller

for a missile whose operating parameters are angle of attack and Mach number [6].

As demonstrated in the creation of controllers for a tailless aircraft [12], an F-16

aircraft [13], and a hypersonic aircraft [14], an LPV controller can also be created

by letting the controller have the same linear fractional relationship with the varying

parameters as the system [15] while attempting to minimize the H∞ norm. This

technique is further expanded with the controlling of the longitudinal axes of an F-

16 aircraft in a project that breaks a parameter space into two smaller overlapping

parameter spaces, synthesizes an LPV controller for each space, and then uses blending

functions to form a single LPV controller [16]. An LPV controller was created [17]

for an LPV system, where parameter dependent feedback control laws are constructed

after transforming the original LPV system into canonical form. Separate longitudinal

and lateral-directional LPV controllers were designed for the F/A-18 [18]. The original

controllers were formed using H∞ synthesis and then robustness was increased to

meet military standards by using µ synthesis. Other recent efforts at using real-

time parameter information in control strategies included minimizing linear matrix

inequalities [19, 20].

This thesis presents one of the more recent gain-scheduling techniques for creating

an LPV controller using H∞ synthesis, which is designed to work for the LPV system’s

entire operating domain. The operating domain of an LPV system is also known as the

system’s parameter space. Linear parameter-varying control theory is discussed in more

detail in the next section.

CHAPTER 2LINEAR PARAMETER-VARYING CONTROL THEORY

Linear parameter-varying controller synthesis is a gain-scheduling technique for

designing one controller that will work over a range of parameters without having

to create a scheduling scheme. In order to use the LPV framework the plant model

must be created as a linear parameter-varying system. A linear parameter-varying

system depends affinely on a set of norm-bounded time-varying operating parameters.

It considers linear systems whose open-loop dynamics are affine functions of the

operating parameters. A method of identifying multivariable LPV state space systems

that are based on local parameterization and gradient search in the resulting parameter

space is presented in [21]. Two identification methods were purposed in [22] for a

class of multi-input multi-output discrete-time linear parameter-varying systems. Both

methods are based on the subspace state space method, which was suggested by [23]

in the early 1990s. LPV modeling of aircraft dynamics, known as the bounding box

approach and the small hull approach [24].

A general case of a linear parameter-varying plant, whose dynamical equations

depend on physical coefficients that vary during operation, has the form

P

θ x A

θ x B1

θ d B2

θ u

e C1θ x D11

θ d D12

θ u

y C2θ x D21

θ d D22

θ u (2.1)

where

θt

θ1t θn

t θi θi

t θi (2.2)

5

6

is a time-varying vector of physical parameters (i.e., velocity, angle of attack, stiffness);

A, B, C, D are affine functions of θt , x is the state vector, y is the measured outputs,

e is the regulated outputs or errors, d is the exogenous disturbances, and u is the

controlled input. When the coefficients undergo large variations it is often impossible

to achieve high performance over the entire operating range with a single robust

LTI controller. When parameter values are measured in real time controllers that

incorporate such measurements to adjust to current operating conditions would be

beneficial. These controllers are said to be scheduled by the parameter measurements.

This control theory typically achieves higher performance when considering large

variations in operating conditions. In the event that different parameters effect the

system differently weighting functions can be used to compensate for the differences.

If the parameter vector θt takes values within a geometric shape of Rn with

corners Πi Ni 1N 2n , the plant system matrix

Sθ : x

e

y

Aθt B

θt

Cθt D

θt

x

d

u

(2.3)

ranges in a matrix polytope with vertices SΠi . Given convex decomposition

θt α1Π1 αNΠN

αi 0

N

∑i 1

αi 1 (2.4)

of θ over the corners of the parameter region, the system matrix is given by

Sθ α1S

Π1 αNS

ΠN (2.5)

This suggests seeking parameter-dependent controllers with equations

K

θ ζ AKθ ζ BK

θ y

u CKθ ζ DK

θ y (2.6)

7

and with a vertex property where a given convex decomposition θt ∑n

i N αiΠi of

the current parameter value θt . The values of AK

θ ,BK

θ ,CK

θ ,DK

θ are derived

from the values AKΠi ,BK

Πi ,CK

Πi ,DK

Πi at the corners of the parameter region

by ! AKθ BK

θ

CKθ DK

θ "$#% N

∑i N

αi

! AKΠi BK

Πi

CKΠi DK

Πi

"$#% (2.7)

In other words, the controller state-space matrices at the operating point θt are

obtained by convex interpolation of the LTI vertex controllers

Ki : ! AKΠi BK

Πi

CKΠi DK

Πi

"$#% (2.8)

This yields a smooth scheduling of the controller matrices by the parameter measure-

ments θt .

As an example, consider the following H∞-like synthesis problem relative to the

interconnection in Figure 2–1. If there exists a continuous differentiable function Xθ

K

θ P

θ &d & e

u

&y'

Figure 2–1: H∞ Block Diagram (Gain-Scheduled)

defined on Rn where

Xθ )( 0

(2.9)

8

the worst-case closed-loop RMS gain from d to e does not exceed some level γ ( 0,

and !I 0

Aθ B

θ

0 I

Cθ D

θ

"$#######%* !

Xθ X

θ 0 0

Xθ 0 0 0

0 0 + γI 0

0 0 0 1γ I

"$#######% !

I 0

Aθ B

θ

0 I

Cθ D

θ

"$#######% , 0 (2.10)

hold for all θ - Rn, then the system is quadratically stable and the L2 norm from d

to e is smaller than γ. The quadratic stability of a system allows the parameter to

change with arbitrary speed without threatening stability of the system and is defined as

existing if there exists a real positive definite matrix P PT ( 0 such that

AT θ P Pθ , 0 . θ

t )- Rθ

(2.11)

The induced L2 norm of a quadratically stable LPV system G is defined as/G/ sup

θ 0 Rθ

sup1d1

2 2 0 3 d 0 L2

/e/

2/d/

2(2.12)

with Rθ being a set of feasible parameter trajectories.

There is more than one synthesis technique for designing an LPV controller once

the LPV model is formed. Currently there are three predominant synthesis techniques,

µ synthesis design [8], Linear Quadratic Gaussian (LQG) control design [25], and H∞

control design [26]. The µ synthesis technique attempts to minimize the µ value over

stabilizing the controller, K, and diagonal, D, while D-K iteration is used to reduce

the cost function. The LQG controller design method synthesizes a controller which is

optimal with respect to a specified quadratic performance index and takes into account

the Gaussian white noise disturbances acting on the system. The technique used

for the projects presented in this paper is the H∞ control synthesis technique which

uses the linear fractional form of the LPV system and creates the controller while

attempting to minimize the H∞ norm. By letting the controller have the same linear

9

fractional relationship with the varying parameters as the LPV system the H∞ control

problem is formulated using linear matrix inequalities (LMI). The appearance of LMIs

in the control synthesis shows how the control problem is a convex optimization

problem [27], as was described in the previous example. Another example of creating

a convex optimization problem with LMI expressions for the use of finding an LPV

controller for the attitude control of an X-33 is presented in [28].

The main benefit of using the LPV framework is that it allows gain-scheduled

controllers to be treated as a single entity, with the gain-scheduling being achieved with

the parameter-dependent controller and automatic interpolation law, which removes the

ad-hoc scheduling schemes that were necessary in the past.

CHAPTER 3LINEAR PARAMETER-VARYING CONTROL FOR AN F/A-18

3.1 Problem Statement

Several control designs have been applied to the control of F-18 aircraft. A

controller was designed using H∞ and µ synthesis techniques for a single flight

condition [29]. Though this technique works well for a single point in the flight

envelope a type of gain-scheduling is necessary for controlling the F/A-18 throughout

its operating domain. A longitudinal variable structure controller was created for an

F-18 model with parameter perturbations [30]. Though this technique can attain the

conventional goals of stability and tracking for uncertain nonlinear plants, a reference

trajectory for tracking control must be specified, which indicates that the controller

cannot operate over a large flight envelope. A lateral-directional controller was created

using µ synthesis with parametric uncertainty to account for gain differences between a

nominal model and trim models and multiplicative uncertainty to account for changes

between a nominal model and other trim models within the chosen flight envelope [31].

Because this technique uses constant blocks of uncertainty instead of gain-scheduling

the flight envelope used for the project had to be small, M 54 0 35055 6 and altitude74 20

28 6 k f t. Gain-scheduled approximations to H∞ controllers for the F/A-18

Active Aeroelastic Wing, located at NASA Langley Research Center, were developed

within another project [32]. Point design controllers were created within a small flight

envelope and then a scheduling scheme of the gains had to be formed. A multivariable

LPV controller was designed using H∞ synthesis for the F/A-18 System Research

Aircraft (SRA), located at NASA Langley Research Center, in [33]. Though this

technique is also chosen for control synthesis in the project presented in this chapter,

the flight envelope that the controller had to operate within is smaller in [33], with

10

11

M 74 0 35070 6 and altitude 84 15

32 6 k f t. A similar project involving an LPV

controller for the F/A-18 SRA, in [18], uses the same synthesis technique but uses an

even smaller flight envelope, M 94 0 45055 6 and altitude :4 20

25 6 k f t, than [33].

The project presented in this study discusses the formation and simulations of a

linear parameter-varying controller for the longitudinal dynamics of an F/A-18 over a

chosen flight envelope. The F/A-18, shown in Figure 3–1, has a ceiling of 50,000+ f t

and a speed of M=1.7+. As the aircraft’s altitude varies so does the air density which

affects the aircraft’s response to control surface deflections. Furthermore, the amount of

deflection necessary for a particular maneuver varies as the Mach number varies. These

aerodynamic changes that occur with the large range in altitude and Mach number

make it necessary to incorporate a gain scheduling technique for control. The flight

envelope for this project is limited to Mach numbers from 0.4 to 0.8, which includes

both incompressible and compressible subsonic flows, and an altitude range from

10,000 f t to 30,000 f t, which includes a density change of roughly 0.9 E ; 3 slug < f t3.

Figure 3–1: F/A-18

The flight envelope can also be considered the parameter space for which the

LPV controller will be designed. The parameter space is two dimensional with the first

parameter dimension being Mach number and the second parameter dimension being

altitude. Originally four points within this two-dimensional parameter space were to be

used to design the LPV controller and are listed in Table 3–1. However, the dynamic

12

pressure, q, for the model at Mach=0.40 at an altitude of 30 kft, P4, was too low to

control and therefore the model was discarded.

Table 3–1: Original Design Points

Design Point Mach Number Altitude ( f t)P1 0.4 10,000P2 0.8 10,000P3 0.8 30,000P4 0.4 30,000

The controller performance is tested with each of the remaining models and with

a model whose dynamics represent the aircraft at a Mach number of 0.6 and at an

altitude of 20,000 ft. A depiction of the flight envelope which represents the parameter

space and the placement of the models used for this project are shown in Figure 3–2.

0.3 0.4 0.5 0.6 0.7 0.8 0.95

10

15

20

25

30

35

P1

P2

P3

PA

Mach Number

Alti

tude

(kft

)

Design PointDesign PointDesign PointAnalysis Point

Figure 3–2: Flight Envelope/Parameter Space

3.2 Open-loop Dynamics

The F/A-18 models used for this project are longitudinal short-period approxima-

tions that were developed with two states, one input and one output. The states include

angle of attack (deg) and pitch rate (deg < sec). The input is the elevator deflection and

the output is pitch rate.

13

The model for the F/A-18 at Mach=0.40 at an altitude of 10 kft is given as P1

such that q P1 δ.

P1 + 07433 425

6200 + 0

5642+ 0

0022 + 0

4064 + 0

0662

0 573 0

(3.1)

The model for the F-18 at Mach=0.80 at an altitude of 10 kft is given as P2 such

that q P2 δ.

P2 + 18415 853

1909 + 2

0292+ 0

0192 + 0

9431 + 0

2568

0 573 0

(3.2)

The model for the F-18 for Mach=0.80 at an altitude of 30 kft is given as P3 such

that q P3 δ.

P3 + 08399 791

1313 + 0

9314+ 0

0075 + 0

4499 + 0

1190

0 573 0

(3.3)

The model of the analysis point with an altitude of 20,000 ft and Mach=0.6 is

given as PA such that q PA δ.

PA + 08280 617

0114 + 0

8269+ 0

0075 + 0

4499 + 0

0994

0 573 0

(3.4)

The frequency and damping ratio for the each of the models were determined and

are shown in Table 3–2. All of the damping ratios are greater than zero, which affirms

that the models are stable.

14

Table 3–2: Frequency and Damping Ratio of Design and Analysis Models

Model ω ζP1 1.113 0.5166P2 4.257 0.3271P3 2.512 0.2567PA 2.288 0.2764

The linear parameter-varying model for the parameter space is given as Pθ and

is given as q Pθ δ.

P = θ >@?BA P1 CED FGGGGHJI 1 K 0982 427 K 57 I 1 K 465I 0 K 017 I 0 K 5367 I 0 K 1906

0 0 0

LNMMMMO θ1 = t > D FGGGGH 1 K 0016 I 62 K 06 1 K 0978

0 K 0117 0 K 4932 0 K 1378

0 0 0

LNMMMMO θ2 = t >(3.5)

Where θ θ1 0

0 θ2

and where θ1 -P4 0 1 6 represents the systems dependence

on Mach number and θ2 -P4 0 1 6 represents the systems dependence on altitude.

The aircraft flying at a Mach number of 0.4 corresponds to a θ1 0 and at a Mach

number of 0.8 corresponds to a θ1 1. The aircraft flying at an altitude of 10,000 f t

corresponds to a θ2 0 and at an altitude of 30,000 f t corresponds to a θ2 1.

3.3 Control Objectives

The control objective for the F/A-18 longitudinal flight controller is to track a

given pitch rate command to within certain tolerances of a target response generated by

a target model that has desirable dynamics. The commanded pitch rate is a step input

which begins at zero magnitude and becomes 10deg < sec at the time of one second and

remains that magnitude until the simulation ends at ten seconds. The response of the

system with the linear parameter-varying controller to the commanded pitch rate must

have a rise time within Q 0.05 sec of the target rise time, an overshoot within Q 4%

of the target overshoot, and a settling time within Q 0.6 sec of the target settling time.

The controller should also have a level of robustness to account for errors in the signal.

15

3.4 Synthesis

The system shown in Figure 3–3 incorporates all the necessary elements needed

to create the controller which will accomplish the controller objectives. This system

contains the open-loop dynamics as described by P and a target model T used

for model following. The system also incorporates weighting functions used for loop

shaping, which are given as WpWnWkWu .

u & Wu& Wk& e2

& Pθ &- & Wp

& e1

qc&R

ST

Rn & Wn&+ & - & y

Figure 3–3: Synthesis Block Diagram

The system has 2 disturbances, 1 control, 2 errors and 1 measurement, which were

referred to as d u e y in the LPV Control section. The disturbances are random

noise n - R affecting the sensor measurement and the commanded pitch rate qc. The

control input is u - R, which affects the elevator deflection. The sensor measurement

of pitch rate, y, is used for feedback to the controller. The errors are, e1, the error

between the target pitch rate response and the LPV model response and the weighted

control effort, e2.

The target model T describes an F/A-18 model that has dynamics which outputs

a desirable pitch rate response. The target model is used for model following to aid in

obtaining the LPV controller. The target model is not a function of parameters in the

operation space. The frequency and damping ratio of the target model are presented in

16

Table 3–3.

T

+ 2 0 0 + 198

0 + 3 + 3 + 278

0 3 + 3 218+ 0

79 + 0

47 1

79 0

(3.6)

Table 3–3: Frequencies and Damping Ratios of the Target Model

Targetω ζ

2.0 1.04.2426 0.7071

The performance filter given as Wp serves to normalize the error in the model

following between the target model and the LPV model. The filter Wp is chosen to

meet design specifications in accordance to pitch rates of the aircraft in response to a

commanded pitch rates. The actual filter is realized as Wp 70.

The filter Wk is used to normalize the penalty placed on the amount of actuation

commanded by the controller. This filter reflects the capabilities of the actuation

system. The weighting is chosen as the inverse of the actuator’s magnitude of motion,

Wk 05.

The input matrix varies from model to model within the parameter space. This

variation necessitates an input filter within the synthesis. The filter shown as Wu is used

as the input filter and has a value of 1E5s T 1E5 .

Also, the filter Wn is included to account for noise that corrupts measurements by

the sensor. The inclusion of noise is needed to provide a minimal level of penalty on

the sensors, which will give robustness to the controller. The design did not want to

consider a large amount of noise so the filter was chosen as Wn 001.

Using the system in Figure 3–3, robust H∞ controllers were designed for the

models at each of the design points and a linear parameter-varying controller was

designed for the entire parameter space. The techniques of H∞ control are used to

17

reduce the induced norm from the input to the weighted errors. The software from the

µ Analysis and Synthesis Toolbox for Matlab is used for the actual computation

of the controller [34]. The same weightings are used to create the controllers in order

to achieve the same performance level for all of the points in the parameter space. The

resulting induced norms achieved by the individual controllers and the LPV controller

are shown in Table 3–4.

Table 3–4: Induced Norms of Closed-Loop System

Open-Loop Model H∞-normP1 0.891P2 0.775P3 0.775

P1 + P3 0.971

It is important to note that all of the closed-loop norms are less than unity. These

magnitudes indicate that the controllers are able to achieve the desired performance and

robustness objectives. The last entry in Table 3–4 is the norm associated with the LPV

controller. Allowing the altitude and Mach number to vary with time increases the

norm as expected. However, this norm did not raise much above the norm associated

with any of the point designed H∞ controllers and stayed below unity. This condition

indicates that the LPV controller is capable of accounting for the time-varying nature

of Mach number and altitude without excessive loss of performance.

3.5 Simulation

The closed-loop dynamics are simulated with a 10deg < s pitch rate step input to

demonstrate the performance of the controller for each of the design models and for the

analysis model. The diagram of the closed-loop system for the models can be seen in

Figure 3–4. The simulations use the same open-loop dynamics but include the linear

parameter-varying controller that was synthesized over the parameter space.

The response to the step input of the LPV controller with the point design models

and the response of the target model are shown in Figure 3–5. The point design

18

δ

&Kθ Pθ & qS

qc'-'

Figure 3–4: Closed-loop Block Diagram

responses only vary roughly 0.2% from the target model response. This characteristic

is due to the LPV controller being created with the models at those points. The

performance of the controller must also be tested with a model that lies away from the

vertices points of the parameter space that were used to create the controller.

0 2 4 6 8 100

2

4

6

8

10

12

Time (s)

Pitc

h R

ate

(deg

/sec

)

CommandTargetP

1P

2P

3

Figure 3–5: Pitch Rate for Design Points

The analysis point was chosen to be the farthest from the vertices of the parameter

space which results in a Mach number of 0.6 and an altitude of 20,000 f t. The

responses of the analysis model and the target model, using the same step command

that was used for the point design simulation, are shown in Figure 3–6. The results

appear to be quite close to the target response. Numerical results were pulled from the

plot to make a closer comparison and are shown in Table 3–5.

The same time response and delay time are apparent for both the analysis model

and target model responses. The settling time of the analysis model response lags the

target response by 0.5 seconds, which is within the control objectives. The maximum

19

0 2 4 6 8 100

2

4

6

8

10

12

Time (s)

Pitc

h R

ate

(deg

/sec

)

CommandTargetP

A

Figure 3–6: Pitch Rate for Analysis Point

Table 3–5: Numerical Results

Target Model Analysis ModelRise Time 0.21 sec 0.21 sec

Settling Time 1.63 sec 2.13 secPeak Overshoot 5 % 1.2 %

overshoot of the analysis model response was less than that of the target model

response and remains within the bounds of the controller objective.

The controller commanded elevator deflection from the simulations is shown in

Figure 3–7 and is used to determine if the actuation of the elevator is reasonable for

each of the tested models. All of the values are negative because a negative elevator

0 2 4 6 8 10−20

−15

−10

−5

0

Time (s)

Com

man

ded

Ele

vato

r Def

lect

ion

(deg

)

PA

P1

P2

P3

Figure 3–7: Controller Elevator Deflection

20

deflection causes a pitch up in the aircraft, which is the commanded input of the

simulations. The peaks appear at the time when the input pitch rate command is

initiated and the command response that follows is to maintain the pitch rate command.

The peak of the elevator command for the system with the model associated with a

point in the parameter space having a Mach number of 0.4 and an altitude of 10,000 f t,

P1, has the greatest value, -17o. The value is reasonable because at lower speeds a

greater angle is needed to get the same response as flying at faster speeds and there

is a less chance that the control surface will be damaged by the slower airflow. The

elevator command peak for the system using the model with a Mach number of 0.8

and an altitude of 10,000 f t, P2, has the smallest value, -4.5o. This lower value is all

that is needed for the desired pitch rate due to the speed of the airflow around the

control surface at the higher Mach number, which decreases the time the maneuver

requires. The peak value is -9.5o for the elevator command associated with the system

using the model that has a Mach number of 0.8 and an altitude of 30,000 f t, P3, which

is an acceptable magnitude. This value being higher than the value corresponding

to the same Mach number but with a lower altitude of 10,000 f t is expected because

at a higher altitude the density is less and so fewer air particles are present to be

affected by the deflection, therefore a larger angle is necessary. The peak commanded

elevator deflection for the system using the analysis model, PA, is -12o, which is also

acceptable. The magnitude is reasonable because it is less than the value commanded

for P1 due to the higher Mach number of the analysis model and is not too small that

the increase in altitude would have an adverse affect.

3.6 Conclusion

This project considered the control of the pitch rate of an F/A-18 aircraft with

a linear parameter-varying controller. This type of controller was chosen because the

change in dynamics of the aircraft could be modeled with a system whose state-space

matrix and input matrix were affine functions of the parameters, Mach number and

21

altitude. Once the controller was created, it was tested at certain points within the

parameter space using a step pitch rate input. The results allow for the conclusion

that the LPV controller performed the specified objectives and is therefore a sufficient

controller for the F/A-18 model presented in this project.

CHAPTER 4LINEAR PARAMETER-VARYING CONTROL FOR A HYPERSONIC AIRCRAFT

4.1 Problem Statement

All aircraft flown today fly within the subsonic, transonic and supersonic flight

regimes. The push toward faster and higher flying aircraft has moved the envelope

into the hypersonic regime. This push comes from both military and commercial

groups. The military wants a bomber that can fly at high altitude, over a long range

and at high speeds, so that the vehicle is nearly impossible to shoot down. Commercial

groups would like to have a more reliable way of sending satellites into low earth orbit.

The major problem with the use of rockets is that if something goes wrong during

ascension into orbit the cargo will most likely be destroyed along with the rocket. The

use of a hypersonic aircraft presents a more reliable transportation for the satellite

because if an error did occur during the flight there would be a chance that the aircraft

could maneuver to a landing area.

Though the concept of hypersonic flight has been discussed since the 1950s

the mass construction of hypersonic aircraft has been hindered by the necessity of

the technology and the price of materials that are able to withstand the elements

in which the vehicles must operate. This obstacle may have slowed the creation of

such vehicles but several control theories have still been created. The more popular

control theories include H∞ [35], µ synthesis [36] and linear parameter-varying

control [37]. The theories involving H∞ and µ synthesis, however, only considered

a single flight condition for the hypersonic vehicle. Also, the previous project that

used a linear-parameter varying controller for the hypersonic vehicle ignored the

mode shape of the vehicle and separated the rigid-body dynamics and the structural

dynamics of the hypersonic model. A scheduled longitudinal control scheme was

22

23

created which incorporated a set of parameter controllers, where the parameters were

Mach number and dynamic pressure, and was determined from linear designs using

analytic functions of the parameters [38]. That project focused on the control of

the rigid body dynamics and did not recognize the effect of structural modes on the

response of the hypersonic vehicle. Robust flight control systems are synthesized

for the longitudinal motion of a hypersonic vehicle using stochastic cost functions

and ten design parameters [39]. That project also focused on the control of the rigid

body dynamics of the hypersonic vehicle without addressing structural dynamics.

The control of the longitudinal motion of a hypersonic vehicle was also addressed,

where robust flight control systems with a nonlinear dynamic inversion structure were

synthesized [40]. Nonlinear control laws were designed so the control systems would

operate over a chosen flight envelope. Again, the rigid body dynamics were the focus

of control. A dual neural network structure was developed that served as feedback

control and optimized the vehicles trajectory to pre-specified burnout conditions in

velocity, flight path angle and altitude [41]. That project serves more as an aid in

the study of trajectory optimization than as a control theory for hypersonic vehicles.

Another project applied a hierarchical integrated control methodology to a hypersonic

vehicle to reduce stabilizing control power required for specific flight conditions [42].

That methodology decomposes the hypersonic model into decoupled subsystems,

creates a controller for each subsystem and a control law for each subsystem controller

is derived. The decoupling of a hypersonic system may not be feasible due to the

large degree of coupling between the physical structure and propulsion system of

the vehicle. Also, the creation of separate control laws is laborious compared to

the LPV method which forms an automatic interpolation law. The control of the

lateral dynamic stability characteristics of a hypersonic vehicle for a specified Mach

number and altitude trajectory has also been detailed in a project [43]. The controller

was designed using Multi-Model Eigenstructure, which designs a robust fixed-gain

24

controller that guarantees robust stability and desired flight qualities along a specified

reference trajectory. The controller would need to be altered if the vehicle deviated

from the preset trajectory or if a flight envelope was to be considered. The same

model of the longitudinal dynamics of a typical hypersonic vehicle were used, where

a unified approach to H2 and H∞ optimal control was used to design a controller for

a specific flight condition [44]. A unified approach alleviates difficulties with the

“over crowding” of a system’s roots inside the unit circle along with other numerical

difficulties. Using the technique in that project would require more controllers to be

created at other operating conditions along with a gain-scheduling law if the vehicle’s

operating range spanned more than a single condition.

Some of the challenges of hypersonic flight include the varying of the hypersonic

vehicle’s dynamic characteristics due to a wide range of operating conditions and

mass distributions for which a type of gain-scheduling technique appears to be

essential [45, 46]. Further discussion of a typical hypersonic vehicle’s dynamics

addresses how the combination of the propulsion system and aeroelastic effects

contribute to the overall dynamic character of the vehicle, which presents the need

of structural dynamic controller [47]. This necessity is the motivation for the project

presented in this chapter.

The controller designed for the hypersonic vehicle for this project was split into

an inner-loop controller and an outer-loop controller. The inner-loop controller is an

LPV controller which must actively damp the structural modes across a temperature

range. Unlike previous hypersonic controls, this controller will focus on the damping

of the mode shape that is associated with the structural dynamics of the vehicle, which

will operate throughout a range of a specific operating parameter and for which the

hypersonic model’s rigid-body and structural dynamics will not be separated. The

outer-loop controller of the aircraft will be a rigid-body controller which will work as

25

a traditional flight controller for rigid aircraft and will be designed in a future project.

The inner-loop structural damping controller is the focus of this project.

A diagram of the inner-loop/outer-loop control design is shown if Figure 4–1.

The Kθ controller is the linear parameter-varying inner-loop controller and the Kout

controller is the rigid-body outer-loop controller. The P is the hypersonic plant model.

Kout

Kθ P&&&input & out putR R ''

Figure 4–1: Inner-Loop/Outer-Loop Design

4.2 Generic Hypersonic Vehicle

The intended uses of hypersonic vehicles ranges from putting satellites into low

earth orbit to being the next stealth bomber. These missions require the vehicle to

travel through each flight regime: subsonic, transonic, supersonic, hypersonic and

orbital. This project will only consider the aircraft’s flight within the hypersonic

regime.

Several hypersonic designs have been created which attempt to maximize aerody-

namic and propulsive efficiency while still having enough controllability. Most of these

designs incorporate the elevator and aileron into one structure known as the elevon.

The form of hypersonic vehicle used for this project is similar to the NASP and X-30

vehicles. A generalized shape can be seen in Figure 4–2.

Figure 4–2: Simplified Model of a Generic Hypersonic Vehicle

26

This configuration of a hypersonic vehicle combines the fuselage with the

propulsion system. This combination greatly affects the flight dynamics of the vehicle.

The forebody of the vehicle acts as the compressor for the engine. The air flow

through this compressor creates a pitch up moment. The aftbody of the vehicle acts

as the exit nozzle for the engine. The airflow through the exit nozzle creates a pitch

down moment. Also, a change in angle of attack or sideslip affects the engine inlet

conditions which changes the propulsion performance. To create a controller for

this type of vehicle the angle of attack, pitch angle and pitch rate are measured for

feedback to the controller.

Another area of hypersonic flight that must be considered when creating a

controller is the speed, and consequently temperature, at which the vehicle flies. As the

vehicle enters the hypersonic regime, the strength of shock waves increase and lead to

higher temperatures in the region between the shock and the body. As Mach number

increases further, the shock layer temperature becomes large enough that chemical

reactions occur in the air. Also, an increase in temperature effects the structural

dynamics of the vehicle in that there is a reduction in the frequency of the structural

modes. Therefore, the controller created in this project will consider temperature as the

flight parameter.

4.3 Hypersonic Model

The hypersonic model [48] used for this project was limited to the longitudinal

motion and was developed with seven states, three inputs and six outputs. The states

include altitude, velocity, angle of attack, pitch angle, pitch rate, and two elastic states

for the fuselage bending mode. The inputs include elevon deflection, diffuser area

ratio and fuel flow ratio. The outputs include angle of attack, pitch rate at forebody,

pitch rate at aftbody, combustor inlet pressure, Mach and thrust which will be used

as feedback to the controller. Only the angle of attack and the two pitch rates are to

be used as feedback to the controller due to their strong dependence on the structural

27

dynamics. Aerodynamic, inertial, propulsive, and elastic forces were used to derive the

equations of motion for the hypersonic vehicle [37]. The model dimensions and flight

conditions are shown in Table 4–1.

Table 4–1: Model Dimensions and Flight Conditions

Length 150 f tMass 300,000 lb

Height 100,000 f tMach 8

Dynamic Pressure 1017 ps f

4.4 Linear Parameter-Varying System

The time-varying operating parameters, θ, are flight parameters which affect the

aircraft during flight. These parameters are measured by sensors on the aircraft and

are sent to the controller. This project takes into account only one flight parameter,

temperature, due to the large affect that temperature has on a hypersonic vehicle’s

structural dynamics. This parameter will have a range from (0oF to 5000oF) to match

the temperature ranges noted for the hypersonic flight of the X-30 and the HyperX

vehicles [49]. The parameter dependence of the model is shown in the matrices below,

θ 0 for the coldest temperature and θ 1 for the hottest temperature within the

range. As the flight parameter, temperature, changes during flight so does the amount

it affects changes in the aircraft. This problem can be compensated with the use of

weighting functions which will be discussed in the next section.

Aθ VU A W θ U Aθ W (4.1)

28

A XYZZZZZZZZZZZZZZZZ[

0 0 ; 7 \ 9248E3 7 \ 9248E3 0 0 0

1 \ 5026E ; 4 ; 3 \ 2374E ; 3 ; 5 \ 2818E1 ; 3 \ 2200E1 2 \ 3762E ; 2 5 \ 7314E ; 1 7 \ 5583E ; 3

1 \ 1744E ; 7 ; 3 \ 1848E ; 7 ; 3 \ 3921E ; 2 0 1 1 \ 4681E ; 4 2 \ 8801E ; 6

0 0 0 0 1 0 0; 5 \ 7586E ; 6 9 \ 6079E ; 6 1 \ 5833E0 0 ; 5 \ 1609E ; 2 9 \ 2411E ; 2 ; 1 \ 8285E ; 4

0 0 0 0 0 0 1; 7 \ 4858E ; 1 1 \ 0158E ; 1 2 \ 4280E3 0 ; 7 \ 4847E0 ; 3 \ 1086E2 ; 9 \ 4975E ; 1

] ^^^^^^^^^^^^^^^^_ (4.2)

0 0 0 0 0 0 0

0 0 0 0 0 0 0

0 0 0 0 0 0 0

0 0 0 0 0 0 0

0 0 0 0 0 0 0

0 0 0 0 0 0 0

0 0 0 0 0 + 25E2 + 0

2

(4.3)

B

0 0 0+ 6435E1 + 7

462E1 1

261E3+ 1

448E + 2 + 1

596E + 2 + 2

253E + 2

0 0 0+ 2455E0 8

111E + 1 5

190E0

0 0 0

6740E2 + 2

925E1 2

209E2

(4.4)

Cθ `U C W θ U Cθ W (4.5)

29

C abccccccccccccccd

0 0 1 0 0 1 e 7453E f 2 0

0 0 0 0 1 0 1 e 7453E f 2

0 0 0 0 1 0 f 1 e 7453E f 2f 4 e 6971E f 5 2 e 0641E f 4 6 e 2428E0 0 f 1 e 0921E f 2 1 e 0896E f 1 0f 5 e 3709E f 6 1 e 0095E f 3 0 0 0 0 0f 3 e 5754E f 1 6 e 0213E f 1 1 e 8399E4 0 f 3 e 2185E1 3 e 2112E2 0

gihhhhhhhhhhhhhhj(4.6)

0 0 0 0 0 C16 lk 1

05 0

0 0 0 0 0 0 C17 mk 0

05

0 0 0 0 0 0 C227 lk 0

05

0 0 0 0 0 C236 lk 0

05 0

0 0 0 0 0 0 0

0 0 0 0 0 C256 lk 0

05 0

(4.7)

D

0 0 0

0 0 0

0 0 0

0 + 7229 0

0 0 0

0 + 3158E4 5

995E5

(4.8)

As seen in the linear parameter-varying matrices above, both the state matrix 4 A 6and the observation matrix 4C 6 change with temperature. It is common for the state

matrix to change as operating parameters change, but it is not common, in traditional

aircraft, for the observation matrix to change. This change in the observation matrix

accounts for the mode shape changes of the hypersonic vehicle.

The modes of the hypersonic model are shown for different temperatures in

Table 4–2 . The table shows the frequency of each of the modes and the damping

30

corresponding to the frequency. The four modes of the open-loop dynamics are (i)

a height mode, (ii) an unstable phugoid-like mode, (iii) an unstable pitch mode and

(iv) the structural mode. As can be seen in the table, the structural mode for the

model at the cold temperature has a higher frequency than the structural mode at the

hot temperature. Minimizing the affect that the temperature has on this mode is the

objective of the inner-loop LPV controller.

Table 4–2: Modes of the Hypersonic Model

Cold HotMode ω

rad < sec ζ ω

rad < sec ζ

i 0.0024 1.00 0.0024 1.00ii 0.1666 1.00 0.1790 1.00

0.1677 -1.00 0.1804 -1.00iii 1.462 -1.00 1.518 -1.00

1.554 1.00 1.608 1.00iv 17.65 0.0268 15.84 0.0062

4.5 Control Design

The control objective of the linear parameter-varying controller is to damp out

the structural mode in order to minimize the affect that temperature has on the model.

The controller should also contain a level of robustness to account for errors in signals.

A system was created that incorporated the necessary elements to accomplish these

objectives. The first step in the finding the LPV controller was to create a synthesis

model shown in Figure 4–3.

The system has 2 disturbance inputs, 1 control input, 2 error outputs and 1

measurement output. The disturbance vector n - R3 is random noise which affects

sensor measurements. The incorporation of noise creates a small level of robustness

within the controller. The disturbance δ - R is a commanded elevon deflection. The

control input u - R is the excitation from the controller affecting the control actuators.

The error ep - R is the weighted measurements of the angle of attack by the sensors.

The error ek - R is the error of the control actuation. The measurements in the vector

31

u & Wk& ek

R+δ & & Pθ SSSWn&n &+ S&+ S&+ S

y

& T &-R & Wp& ep

X 'S

Figure 4–3: Synthesis Block Diagram

y - R3 are the sensor measurements of angle of attack, pitch rate at the forebody and

pitch rate at the aftbody which will be used for feedback to the controller.

The open-loop dynamics of the LPV system is described by Pθ . Where,

Pθ A

θ B

Cθ D

(4.9)

A target model, T , is created to describe a hypersonic model with desirable

structural damping and therefore incorporates the controller objective. The target model

was used for model following to aid in obtaining the LPV controller. The target model

modes and corresponding damping are shown in Table 4–3. The target model has

a large magnitude of damping corresponding to its structural mode compared to the

damping found in the hot and cold temperature models. It is this amount of damping

that the controller must impose upon the hypersonic model throughout the temperature

range.

The performance filter, Wp, would normally be used to define the design specifi-

cations in the frequency domain. For this synthesis Wp was made equal to 1.5 which

32

Table 4–3: Modes of the Target Model

TargetMode ω

rad < sec ζ

i 0.0024 1.00ii 0.1728 1.00

0.1735 -1.00iii 1.478 -1.00

1.590 1.00iv 16.75 0 .2381

allows measurements through all frequencies to pass through with only a slight de-

crease in gain. This passage throughout all frequencies was allowed because of the

simple controller X which was incorporated into the system to stabilize the vehicle.

A simple H∞ controller, X , is created in order to stabilize the rigid-body dynamics

of the hypersonic vehicle without an affect on the structural mode. This small con-

troller was implemented so that the structural dynamics controller would not try to alter

the rigid-body dynamics. Stabilizing the rigid-body of the model allows the creation of

the LPV controller for the structural dynamics.

The filter, Wn, passes an allowed amount of noise to the sensors. Wn 001

because only a small amount of noise was needed to pass into the system to ensure that

the controller would be robust. The filter, Wk, is used to normalize the restriction on

the amount of actuation the controller commands. Wk was chosen so that the weighting

is the inverse of the actuators’ magnitudes of motion, Wk s T 180s T 1000 .

The results of the open-loop synthesis were then used to create the LPV controller,

Kθ , using the LMI ControlToolbox [50]. To determine how well the controller would

work the H∞ norm was found for the system throughout the temperature range, along

with the H∞ norm for the system at the cold temperature and at the hot temperature.

The frequencies at which the H∞ norm occurred for the model at the hot and cold

temperatures were also found. The results of this test are shown in Table 4–4. The

magnitude of the H∞ norms of the model at the hot and cold temperatures mainly

33

draws from the connection of the first input, q, to the first output, the ep, meaning that

the largest error comes from the performance of the angle of attack meeting the elevon

deflection command.

Table 4–4: Open-Loop Synthesis Norms

H∞ norm ωrad < sec

System 0.9386Cold 0.9159 17.90Hot 0.91196 19.81

H∞ controllers were made specifically for the model at the cold temperature and

the model at the hot temperature. The H∞ norms of these point designs were found

and used to compare to those found for the full LPV system. The results are shown in

Table 4–5.

Table 4–5: Point Design Norms

H∞ normCold 0.1476Hot 0.1679

Compared to the norms of the system with the LPV controller at the hot and cold

temperatures and the norms of H∞ controllers at the point designs, the norm of the

LPV system is relatively high. This difference results from the time-varying nature

of the parameters of the system. Despite this increase in magnitude the LPV system’s

H∞ norm is still less than one, showing that the LPV controller that was created is

capable of controlling the system.

4.6 Simulation

4.6.1 Open-Loop Simulation

The frequency response of the open-loop transfer function between the elevon

defection and the angle of attack for the target model, cold model and hot model is

shown in Figure 4–4. The plot of the response in the frequency domain demonstrates

the need for the control of the structural dynamics. The peak in the response that

34

is located in the high frequency region is the structural mode. The target model’s

structural mode peak shows a damped response, which is desirable. The hot and cold

hypersonic models’ structural mode peaks, however, are very sharp which implies that

there is very little damping. These peaks in magnitude correspond to a bending of the

aircraft at the frequencies at which the peaks occur, which brings the desire for the

controller to be able to damp structural mode.

10−2 100 10210−4

10−3

10−2

10−1

100

101

102

Frequency (rad/sec)

Ope

n L

oop

Tra

nsfe

r Fun

ctio

n

targetcoldhot

Figure 4–4: Open-Loop Transfer Functions

The simulation of an angle of attack response to an elevon deflection input for

the open-loop hypersonic models at both the hot and cold temperatures and the target

model is shown in Figure 4–5. The elevon deflection used for the following simulation

of is shown in Figure 4–6. The angle magnitude is small due to the speed at which the

vehicle flies, a large angle would be harmful at high speeds.

Unlike the target model response the open-loop model response at both the

hot and cold temperatures show an oscillation for approximately five seconds. This

oscillation is due to the lack of structural damping and should be removed by the

controller during the closed-loop simulation.

35

0 5 10 15−4

−3

−2

−1

0

1

2

3

Time (s)

Ang

le o

f Atta

ck (d

eg)

targetcoldhot

Figure 4–5: Open-Loop Angle of Attack Result

0 5 10 150

1

2

3

4

5

6

Time(s)

Inpu

t Ele

von

Def

lect

ion

(deg

)

Input Elevon Deflection

Figure 4–6: Input Elevon Deflection

4.6.2 Closed-Loop Simulation

The closed-loop dynamics are simulated to demonstrate the performance of

the controller for the hypersonic models at both the hot and cold temperature. The

closed-loop system for both models can be seen in Figure 4–7.

The system shown in Figure 4–7 has one input signal and six output signals. The

input signal δ remains the elevon deflection. The outputs include angle of attack (α),

pitch rate at forebody (q f ), pitch rate at aftbody (qa), combustor inlet pressure (Pi),

Mach (M) and thrust (T h). The X is the same simple controller used to stabilize the

36

Rδ & & Pθ &

T h&

M& Pi

& qa& q f& α

'''''XS

Figure 4–7: Closed-loop Design

vehicle for the open-loop synthesis. The Kθ is the LPV controller that was created

using the synthesis model.

The frequency response of the transfer function between the elevon deflection and

the angle of attack for the closed-loop system for both the cold and hot temperature

hypersonic models is shown in Figure 4–8.

10−2 100 10210−4

10−3

10−2

10−1

100

101

102

Frequency (rad/sec)

Tra

nsfe

r Fun

ctio

n

targetcoldhot

Figure 4–8: Closed-Loop Transfer Functions

Again, the peaks in the high frequency region correspond to the structural mode of

the target model and the hot and cold temperature hypersonic models. The structural

mode is clearly damped by the controller for the hypersonic models. It is important to

note that these responses correspond to the end points of the temperature range, which

implies that if the model were tested at a temperature that falls within the temperature

37

range that a similar damped peak would result. So the control objective of damping the

structural mode was fulfilled by the LPV controller.

The closed-loop simulation of the angle of attack response to the same elevon

deflection used in the open-loop simulation is shown in Figure 4–9. The results are

again presented for the system at both the hot and cold temperatures and for the target

model.

0 5 10 15−4

−3

−2

−1

0

1

2

3

Time (s)

Ang

le o

f Atta

ck (d

eg)

targetcoldhot

Figure 4–9: Closed-Loop Angle of Attack Result

As can be seen, the oscillations that were apparent in the open-loop simulation

have been removed by the controller. This response is due to the damping which the

controller imposed on the system. The hypersonic models’ responses also follow the

target model response more closely throughout the simulation.

The controller commanded elevon deflection in Figure 4–10 is plotted for the

closed-loop simulation in order to verify that the motion commanded did not violate

the limited motion due to the high Mach number. Because the command never exceeds

a magnitude of 5o the command does not violate the constraint associated with the

elevon actuator. The corresponding deflection rate in Figure 4–11 is plotted to verify

that the command does not violate the motion tolerances of the elevon actuator. The

magnitude of the deflection rate is within the limits associated with the actuator.

38

0 5 10 15−4

−3

−2

−1

0

1

2

3

4

5

Time(s)

Ele

von

Def

lect

ion

(deg

)

cold commandhot command

Figure 4–10: Elevon Deflection Command

0 5 10 15

−60

−40

−20

0

20

40

60

80

Time(s)

Ele

von

Def

lect

ion

Rat

e (d

eg/s

)

coldhot

Figure 4–11: Elevon Deflection Rate

4.7 Conclusion

This project considered the control of the structural dynamics of a hypersonic

vehicle with a linear parameter-varying controller. This type of controller was chosen

because the change in the dynamics of the hypersonic vehicle could be modeled

with a system whose state-space matrix and observation matrix were affine functions

of the parameter, temperature. Once this controller was created, it was tested over

a temperature range with an elevon deflection input. The results allowed for the

conclusion that the LPV controller performed the specified objective and is therefore a

sufficient controller for the hypersonic model presented in this project.

CHAPTER 5LINEAR PARAMETER-VARYING CONTROL FOR A DRIVEN CAVITY

5.1 Problem Statement

Research into flow control techniques has been continually evolving as related

technologies mature. These technologies include hardware development, such as

sensors and actuators [51], and software development, such as models and simula-

tions [52], associated with fluid dynamics. In each case, the technologies are being

developed with careful consideration of the requirements for control design and

implementation [53].

A particular challenge for flow control has been the development of open-loop

models for which controllers can be designed. The equations of motion for such

dynamics are well known and detailed computational simulations are routinely

performed. Unfortunately, the equations of motion are highly nonlinear and no methods

are currently practical that can directly utilize them for feedback control synthesis.

A recent study has shown that models can indeed be generated that are amenable

to control a specific type of flow [54]. The system in that study is restricted to creep-

ing flow in a driven cavity. Specifically, the left and right sides of the cavity have zero

flow velocity whereas the top and bottom boundaries are driven by exogenous flow

with fixed velocity and frequency. Models are generated by considering the linearized

dynamics associated with modes obtained via proper orthogonal decomposition [55].

These modes were used to derive controllers for disturbance rejection. The derived

controllers were able to keep the flow nearly stationary at various points throughout the

cavity for varying flow regime despite the exogenous input[56, 57].

This project extends the work of Feng [54] to consider different flow conditions

for the driven cavity. Specifically, the open-loop models are generated by considering

39

40

the phase differential between the disturbances at the top and bottom of the cavity.

Feng’s study only used models of the flow resulting from upper and lower flow

boundaries that were 180o different in phase. This project will consider using models

of the flow resulting from upper and lower flow boundary conditions that are 210o,

195o, 180o, 165o and 150o different in phase. A range of phase differentials was used

to derive controllers that accommodate the unpredictable phase difference that would be

associated with a real driven cavity.

A linear parameter-varying controller is designed for the models corresponding

to the range of various phase differentials. The group of models can be considered as

lying within a parameter space with the parameter being the phase differential between

the excitation and disturbance. The open-loop simulations show a rise in the horizontal

velocity along the centerline as the phase differential is increased. The closed-loop

simulations show a distinct reduction in the centerline velocity and therefore prove the

effectiveness of the LPV controller.

5.2 Background

The development of control algorithms has made much progress in the last few

decades for aerospace applications. The control of fluid flow, however, is one of the

most difficult applications and remains mostly unresolved. The primary difficulty

lies with the inherent nonlinearity of the Navier-Stokes equations which are the

governing equations of the fluid flow. Even approximations of this equation can have

large dimensionality. Within the last decade studies of flow control have addressed

specific flow regimes. These regimes include the control of a driven cavity using LQR

control [56], using piezoceramic actuators to control a channel flow [58] and the use of

synthetic jets for flow control [59].

In more recent years, much research has been done to derive methods to generate

reduced-order flow models. Two specific techniques that employ a reduced basis are

proper orthogonal decomposition (POD) [60] and fluid mode methods [53]. Proper

41

orthogonal decomposition is a model reduction technique in which the most energetic

modes are systematically extracted from numerical simulations. This method of

reduction was used to create the models used in this project. The fluid mode method

uses basis functions which are closely related to the physics of the problem being

solved.

Another area of interest for this project is what is known as Stokes or creeping

flow. The limitations of using Stokes flow are that the flow must be incompressible and

have a Reynolds number less than one. One side effect of lowering a flow’s Reynolds

number is that the acceleration term within the Navier-Stokes governing equation

becomes small compared to the viscous force term. This change allows the equation to

be simplified into the linear Stokes equation [61, 62].

5.3 Driven Cavity Geometry

This project will investigate flow control for the cavity shown in Figure 5–1,

where h0t is the velocity along the top of the cavity, β

t is the velocity along the

bottom of the cavity and Γ ΓL n ΓR n ΓT n ΓB is the boundary of the domain. This

cavity is enclosed by rigid walls with no-slip boundary conditions on the right and left

sides. The top and bottom, however, have nonzero boundary conditions in general.

Figure 5–1: Stokes Driven Cavity Flow Problem

The flow at the top and bottom boundaries have uniform spatial distribution. This

restriction implies that the flow at any point along the upper boundary is identical

42

to the flow at any other point along the upper boundary. Similarly, the flow at any

point along the bottom boundary is identical to the flow at any other point along the

bottom boundary. Such a perfect distribution is not possible because of the singularity

at the points on the corners where the flow is moving on the horizontal boundary but

stationary on the vertical boundary. Such a situation is obviously an approximation, but

this example does serve as an initial problem to demonstrate the methodology.

The approximation within the 2-D cavity is based on a grid with an index of

21x21 points. It is assumed that the measurements of the flow velocity are taken at

19 points along the horizontal centerline of the cavity, with the outer points lying one

grid point away from the closest boundary wall. These measurements only provide the

horizontal velocity of the flow. Also, the sensors generating these measurements are

assumed to exist within the cavity without altering the flow. Again, such a situation is

obviously an approximation, but the example serves to demonstrate the methodology.

5.4 Governing Equations of Motion

Consider first the unsteady Navier-Stokes equations

ρ∂ oV∂t ρ oV p ∇ oV :+ ∇p µ∆ oV (5.1)

subject to boundary conditions described in the past section. The parameter oV is the

velocity field, p is the pressure, ρ is the density and µ is the viscosity of the fluid. The

constants that will be used to nondimensionalize the problem include a characteristic

dimension L, characteristic velocity Vs, and characteristic

43

frequency f . Non-dimensional variables are defined as

x q) x < L (5.2)

y q y < L (5.3)oV q 9oV < Vs (5.4)

p q p < ρLVs (5.5)

t q t f (5.6)

The resulting non-dimensional Navier-Stokes equation can be written as

Re p St∂ oV q∂t q Re r oV q p ∇ qsoV qut :+ Re p St∇ q p q ∆ qvoV q (5.7)

where St is Strouhal number defined as

St L fVs

(5.8)

Define

λ Re p St ρL2 fµ

(5.9)

This project will consider flow such that λ w O1 and the Reynolds number is

Re=0.1. In this case, the terms on the right hand side will balance the first term on the

left hand side and the second term will be neglected.

Define the constant

θ 1λ 1

Re p St(5.10)

Finally, the governing equations for the approximation of Stokes flow is achieved

∂ oV q∂t q + θ∆ oV q ∇p 0 (5.11)

For convenience, Equation 5.11 is rewritten as

∂ oV∂t+ θ∆ oV ∇p 0 (5.12)

44

while keeping in mind that these are nondimensional variables.

5.5 Reduced-Order Linear Dynamics

For the geometry depicted in Figure 5–1, define the flow domain as Ω. The

boundary of the fluid flow domain is given by

∂Ω ΓT n ΓB n ΓL n ΓR (5.13)

It is assumed that the input velocity imparted by the moving wall can be repre-

sented in term of Nc functions ogii 1

nc. Each of the functions ogi are defined on

the entire domain Ω, but are assumed to exhibit specific properties on the boundary. It

is required that ogi x ∂Ωx y 1 f or x - ΓB

0 f or x - ∂Ω z ΓB

(5.14)

Additionally, it is required that ∂Ωogi p ndS 0 (5.15)

for i 1

Nc. This last condition is required to guarantee compatibility of the flow

field with the continuity equation.

The dynamics for incompressible, two-dimensional creeping flows can be repre-

sented in the strong form of Stokes equation by rewriting the Equation 5.12.

∂ oV∂t+ θ∆ oV ∇p of (5.16)

Likewise, the strong form of the continuity equation for incompressible flow is

∇ p oV 0 (5.17)

In these equations, oV is the flow velocity, p is the pressure and of is the body force. It

is assumed that spatially-varying functions oφi , for i 1

Ns have been determined

from the POD procedure which will be discussed in detail later. The functions oφi

constitute the reduced basis used to represent the Ns states in the control model,

45

which include phase differential information. There is a different set of oφi functions

associated with each parameter within the phase differential subspace δ=[150,210],

which is used to create reduced-order models at specific phase differentails. Similarly,

we assume that spatially-varying functions ogi, for i 1

Nc, have likewise been

derived from a simulation or experiment. The functions ogi comprise the influence

functions that determine the controls acting on the fluid flow. It is assumed that these

functions satisfy the following conditions which are conventional in many reduced

basis formulations :

∇ p oφi 0 for i 1

Ns (5.18)

∇ p ogi 0 for i 1

Nc (5.19)oφi x ∂Ω 0 for i 1

Ns (5.20)ogi 0 for x <- ΓB (5.21)

In terms of these reduced basis functions, the velocity appearing in the Stokes equa-

tions is assumed to take the formoV x y t oVMt | NC

∑i 1

ogixy βi

t NS

∑i 1oφixy αi

t (5.22)

To derive a reduced-order model appropriate for control synthesis, it is necessary to

convert the strong form of the governing equations to weak form. The inner product,

bilinear form a p p~ are defined, respectively, as ou ov 4 L2 Ω 6 2 ∆ 2

∑i 1

Ω

uividx (5.23)

a ou ov ∆ θ

2

∑i 1

2

∑j 1

Ω

∂ui

∂xk

∂vi

∂xkdx (5.24)

By substituting the velocity into the strong form of the governing equations, taking

the inner product of the resulting expression with an arbitrary basis function oφi , and

46

integrating over the domain Ω, we obtainNC

∑i 1

ogixy βi

t NS

∑i 1oφixy αi

t oφ j 4 L2 Ω 6 2 a

oVM NC

∑i 1

ogixy βi

t NC

∑i 1oφixy αi

t oφ j r of oφ j t 4 L2 Ω 6 2

(5.25)

The terms are gathered and put the equations in a canonical form appropriate for

control synthesis.

NC

∑i 1

ogi oφ j 4 L2 Ω 6 2βi

t NS

∑i 1

oφi oφ j 4 L2 Ω 6 2αi

t a

oVM oφ j NC

∑i 1

a ogi oφ j βi

t NS

∑i 1

a oφi oφ j αi

t r of oφ j t 4 L2 Ω 6 2

(5.26)

A new set of states that include both the original reduced state variables and the

amplitude of the controls is defined as

Xt α

t

βt B (5.27)

The new set of controls is defined to be the time derivative of the original controls

ut β

t (5.28)

With these definitions of the state Xt and controls u

t , it is possible to write the

weak form of the governing equations asM X

t

A Xt |

B ut

Ct (5.29)

It is important to note that the matrix 4 M 6 will be diagonal if the reduced basis vectors

are derived from a proper orthogonal decomposition. The final form of the reduced

47

state equations can be obtained by premultiplying by the inverse of the matrix 4 M 6Xt

M 1 A Xt

M s 1 QXt

M 1 B ut (5.30)

which can be rewritten in the final form as

Xt 94 A θ 6 X t |P4 B 6 u t |4 E θ t 6 (5.31)

where 4 A θ 6 is a linear function of θ, 4 B 6 is a constant matrix, and 4 E θ t 6 is a

function of h0t .

5.6 Creeping Flow in a Driven Cavity

Since the dynamical model is linear, we can construct oVMt asoVM

t oV0 p h0

t (5.32)

where oV0 is the velocity field of creeping flow when constant horizontal velocity is

imposed on the top of the cavity. It is required that

oV0 x ∂Ωx y 1 f or x - ΓT

0 f or x - ∂Ω z ΓT

(5.33)

Also, we choose Nc 1 and og as the velocity field that corresponds to a constant

horizontal velocity imposed on the bottom wall of the cavity. It is required that

oV0 x ∂Ωx y 1 f or x - ΓB

0 f or x - ∂Ω z ΓB

(5.34)

The velocity can subsequently be decomposed asoV x y t oV0h0t | og x y β t NS

∑i 1oφixy αi

t (5.35)

48

Thus, the reduced-order model in Equation 5.31 can be rewritten explicitly as α

β 4 A1θ 64 A2

θ 6

0 0

α

β 4 E1θ 64 E2 6

0 0

h0

h0 4 B1 61

β (5.36)

Consider just the partition associated with the time-domain equation for α

α A1θ α t | A2

θ β t | B1β

t E1

θ h0

t E2h0

t (5.37)

Formulate the equivalent frequency-domain expression

sαs A1

θ α s A2

θ β s | sB1β

s E1

θ h s | sE2h

s (5.38)

49

A transfer function representation of the inputs to outputs can now be solved.

αs V sE2 E1

θ

s + A1θ h0

s sB1 A2

θ

s + A1θ β

s

A1θ E1

θ | A1

θ E2

I E2

hs

A1 A2 A1B1

I B1

βs

A1θ E1

θ A2

θ

I E2 B1

hs

βs P

s h

s

βs (5.39)

The reduced-order model described by P is a fairly accurate representation of the

open-loop dynamics for Stokes flow in the driven cavity with θ w O1 . All of the

models will be based on a θ 1 in this project. There is a decrease in accuracy as θ

changes from unity because of unmodeled nonlinearities. The linear parameter varying

controller is designed and tested for models over the range of phase differentials

δ 4 150210 6 .

5.7 Excitation Phase Differential

Open-loop models of the flow dynamics are generated by analyzing simulated

responses of the driven cavity. This simulation uses the Stokes flow as the fluid

dynamics. Separate models are generated for each set of flow conditions that corre-

sponded to different relationships between the upper and lower boundaries.

Each of these models have physical limitations associated with them. The

Reynolds number for each model is Re=0.1. The combination of the low Reynolds

number and the approximation of the Navier-Stokes equation creates a flow which is

dominated by viscous effects.

50

The fluid at the upper and lower boundaries was constrained to move at a

sinusoidal frequency with constant amplitude. The frequency of this sinusoidal is

essentially considered a non-dimensional unity because the time scales in the dynamics

are all normalized. This excitation acts as a disturbance that affects the entire flow

within the cavity.

The variations between the flow conditions used for model development were

phase relationships between the upper and lower boundaries. Specifically, the sinu-

soidal flows differed in phase by 210o, 195o, 180o, 165o, and 150o between the upper

and lower boundaries. These phase differences induced different modal structures

within the flow so the various models were generated to represent a basis for this range

of flow conditions.

The reduced-order models associated with each phase differential, which contain

three states, can actually be considered as subspaces of the full-order flow, which

contains 623 states. The exogenous disturbances will, in general, not have a constant

phase differential so the flow field will contain modes associated with many phase dif-

ferentials throughout any time evolution. As such, each reduced-order model represents

a subspace of the modes observed in that full-order flow field. Thus, synthesizing con-

trollers for these models with phase differentials is essentially designing compensators

that are optimal with respect to each subspace.

5.8 Control Design

5.8.1 Control Objectives

The objective of flow control in this project is to reject the effects of the ex-

ogenous disturbance at the top of the driven cavity. Physically, the control seeks to

minimize the horizontal component of the fluid velocity at a set of sensor locations.

These sensors are located at 19 sites evenly distributed along the horizontal centerline

of the cavity. The end sensors are located 1 grid point away from the boundary wall.

51

The controller is designed for disturbance rejection using the model shown in Fig-

ure 5–2. This model contains the open-loop dynamics as described by VmΦΨP .

The weighting functions used for loop shaping are given as WpWnWkWyWhWuWd .

The mathematical objective of the control

design is to choose a K such that the closed-loop transfer function from disturbances to

errors has an induced norm less than unity for all plants within the parameter space.

u

R& Wk

RWa

R∆a

'+ && e2

Wuβ &h & Wh&

Pδ & Ψ

R& Φα &+ & Wp& e1

& Vm S RWy

R+Wn&n & & y

Figure 5–2: Controller Block Diagram

The system has 3 input vectors and 3 output vectors. The inputs are the

random noise n - R 19 affecting the sensor measurements, the exogenous disturbance

h - R affecting the upper boundary of the cavity, and the control input u - R affecting

the lower boundary of the cavity. The outputs are the (frequency domain) weighted

horizontal velocity measurements e1 - R 19, the (frequency domain) weighted control

effort e2 - R , and the sensor measurements y - R 19 used for feedback to the controller.

The open-loop dynamics consist of the cavity’s velocity at the top, a control basis

and flow shape basis. The part of the open-loop dynamics known as Vm is the mean

velocity along the top of the cavity and has the equation Vm hot oVo where ho is the

initial top disturbance and Vo is a static Stokes flow along the top. The control basis

for the plant takes the form Ψβt , which was depicted as og x y β t in the equations

presented in the previous sections, and coincides with the controlled velocity along the

52

bottom of the cavity. The shape of the fluid flow within the cavity is determined by

the flow shape basis Φ. The flow shape can take several forms; for example, the flow

could be one large vortex circling the entire cavity or two vortices of equal size with

opposite rotation that meet along the horizontal center line of the cavity.

The filter given as Wp serves to normalize the measurement of flow velocity

collected by the sensors. This filter achieves loop-shaping that defines the design

specifications in the frequency domain. The filter Wp is chosen to reflect the inverse of

acceptable velocities in response to unity disturbances. The actual filter is realized as

Wp 100 which implies the flow velocities should be less than 0.01 in the closed-loop

system.

The filter Wk is used to normalize the penalty placed on the amount of actuation

commanded by the controller. This filter reflects the capabilities of the actuation

system. The magnitude of motion for the control actuator is allowed to be 02 in the

non-dimensional system. The weighting is chosen as the inverse so that Wk 5.

The creation of a linear parameter-varying controller requires that an output filter

be used because a certain part of the observation matrix and feedthrough matrix vary

with the different plants within the parameter space. The filter Wy is used as the output

filter for this synthesis. The filter was chosen as Wy 1E5s T 1E5 .

Also, the input matrix varies among the different plant models within the param-

eter space. This variation necessitates input filters within the synthesis. The filters

shown as Wu and Wh are used as input filters and have the value of 1E5s T 1E5 .

Multiplicative uncertainty was introduced into the control actuation in order

to increase the system’s robustness with respect to the unmodeled dynamics and

disturbances. The uncertainty is represented in the synthesis as ∆a such that/

∆a/

∞ 1. A weighting was used to limit the amount of uncertainty allowed into the control

actuation and was chosen to be Wa 02, which allows Q 20% uncertainty in the input

u.

53

Also, the filter Wn is included to account for noise that corrupts measurements by

the sensors. The inclusion of noise is needed to provide a minimal level penalty on the

sensors. The design did not want to consider a large amount of noise so the filter was

chosen as Wn 0001.

5.8.2 Synthesis

Robust controllers and a linear parameter-varying controller have been designed

for the system in Figure 5–2. The techniques of H∞ control are used to reduce the

induced norm from exogenous inputs to weighted errors [63]. The software from the µ

Analysis and Synthesis Toolbox for Matlab is used for the actual computation of

the controller [34].

Separate controllers are synthesized for each of the open-loop models based on

excitation phase differentials. For each model, the same weightings are used to reflect

the desire of achieving the same performance level for each phase differential. The

resulting induced norms achieved by the controllers are shown in Table 5–1.

Table 5–1: Induced Norms of Closed-Loop System

Phase Differential H∞ norm150o 7.475165o 8.177180o 7.647195o 10.143210o 10.829

150o + 210o 12.785

The closed-loop norms are all greater than unity. Intuitively, these magnitudes

imply the controller is not able to achieve the desired performance and robustness

objectives. Realistically, it must be kept in mind that there are twenty inputs and

twenty outputs creating a large number of transfer functions. This fact suggests that

the magnitude of the norms is not unreasonable. The resulting closed-loop properties

are studied in more detail shortly. It is shown that the large norms are caused by

excessive control actuation. Essentially, the controller is not able to achieve the

54

desired disturbance attenuation without exceeding the actuation limits. Fortunately, this

violation is at low frequencies and is not expected to have a dramatic impact on the

closed-loop simulations.

Also, the values in Table 5–1 are interesting in the sense that the induced norms

increase as the level of phase differential increases. Such behavior indicates that the

excitation phase differential does indeed have a large impact on the fluid dynamics.

The increasingly poor performance of the controllers demonstrates that the flow modes

for a phase differential of 210o have properties that are more difficult to control than

those for a phase differential of 150o, for example.

The last entry in Table 5–1 is the norm associated with the LPV controller.

Allowing the phase differential to be time-varying increases the norm as expected.

What is important to note is that this norm did not raise much above the norm

associated with the H∞ controller for the 210o phase differential model. This condition

indicates that the LPV controller is able account for the time-varying nature of the

phase differential without excessive loss of performance.

5.9 Simulation

5.9.1 Open-Loop Simulation

A series of open-loop simulations are performed to demonstrate the fluid qual-

itative response resulting from the disturbance for both full-order and reduced-order

models. These simulations are similar in the sense that the same magnitude of dis-

turbance is used for the boundary conditions on the top of the domain. Conversely,

the simulations involving the reduced-order models differ in that the flow on the bot-

tom boundary has different values of phase lag with respect to the flow on the top

boundary.

A series of plots will be shown to visualize the flow conditions. In each, the value

of horizontal velocity will be shown as a function of time. The plots are 3-dimensional

because the velocity measured at each of the 19 sensors is shown as a function of time.

55

Again, it is important to note that all measurements are non-dimensional. This

characteristic applies to both the time and velocity component so no units are noted for

the simulations.

The open-loop flow for the full-order model is used as a comparison for the

reduced-order model simulations (Figure 5–3). This plot clearly shows the sinusoidal

nature of the flow that results from the top exogenous disturbance changing with the

sine function, ho sin2πt . The flow near the center of the cavity, near point 11,

shows the largest velocity with a magnitude near -0.2 at t 03 to +0.2 at t 0

7.

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint Index

Vel

ocity

Figure 5–3: Open-Loop Flow Velocities for Full-Order Model

The flow for the reduced-order model with a phase differential of 165o is shown

in Figure 5–4. This plot also demonstrates a sinusoidal nature, but has a smaller

open-loop magnitude compared to the full-order flow with the highest velocity being

0.07.

The flow for the reduced-order model with an phase differential of 210o is shown

in Figure 5–5. The flow again demonstrates a sinusoidal nature and the velocities are

slightly larger than those of the full-order model.

A sinusoidal trajectory of phase differentials shown in Figure 5–6 is used in a

simulation which shows the open-loop characteristics of the reduced-order flow as

phase differential changes.

56

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint IndexV

eloc

ity

Figure 5–4: Open-Loop Flow Velocities for Reduced-Order Model with 165o PhaseDifferential

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint Index

Vel

ocity

Figure 5–5: Open-Loop Flow Velocities for Reduced-Order Model with 210o PhaseDifferential

0 0.2 0.4 0.6 0.8 1150

160

170

180

190

200

210

Time

Phas

e D

iffe

rent

ial

Figure 5–6: Trajectory of Phase Differential

57

The flow velocities for the reduced-order model throughout the time-varying phase

trajectory are shown in Figure 5–7. The sinusoidal nature that is apparent in all of the

other open-loop flows is slightly different for this open-loop flow. This difference is

due to the changing of the parameter through the trajectory. The full-order flow does

not have a dependence on phase differential, therefore, the velocities for the full-order

model’s flow over the phase differential trajectory are the same as those plotted in

Figure 5–3.

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint Index

Vel

ocity

Figure 5–7: Open-Loop Flow Velocities for Reduced-Order Model over a Trajectory ofPhase Differentials

An interesting feature to note is that the flow for each reduced-order model

with phase differential has a similar shape but significantly different magnitude. The

maximum velocity measured at the center of the cavity is smaller in magnitude for the

models with phase differentials located at the beginning of the range than the models

with phase differentials near the end of the range. This feature indicates the flow is

indeed strongly dependent on phase differential and should be considered for control

design.

5.9.2 Reduced-Order Closed-Loop Simulation

The closed-loop dynamics are also simulated to demonstrate the performance of

the controller for the reduced-order models, in this section, and the full-order model,

in the next section. The diagram of the closed-loop system for both the reduced-order

58

models and the full-order model can be seen in Figure 5–8. These simulations use the

same open-loop dynamics but include the linear parameter-varying controller that was

synthesized over the range of phase differentials, which contains 24 states. In each

simulation, the flow on the upper boundary is the same, but now the flow on the lower

boundary results only from the commands issued by the controller. In this section,

the controller was tested with reduced-order models for two specific cases of phase

differential and over a time-varying trajectory of phase differentials.

K '&h &

β Pδ & Ψ

R& Φα && Vm S

Figure 5–8: Closed-loop System

The measured velocities for the reduced-order model with a phase differential of

165o in response to the LPV controller with a phase differential of 165o is shown in

Figure 5–9. The comparison of these velocities with the open-loop measurements in

Figure 5–4 demonstrate a reduction of velocity along the center of the cavity, where the

velocity is greatest, of roughly 70%.

The measured velocities for the reduced-order model with a phase differential

of 210o in response to the LPV controller with a phase differential of 210o is shown

in Figure 5–10. The reduction in velocities is apparent by comparing the closed-loop

velocities in Figure 5–10 with the open-loop velocities in Figure 5–5, which shows a

reduction along the center of the cavity of roughly 90%.

The closed-loop simulation of the reduced-order models over the phase differential

trajectory, whcih also effects the controller, is shown in Figure 5–11. The velocity

magnitude shows a clear reduction in magnitude compared to the open-loop simulation

59

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint IndexV

eloc

ity

Figure 5–9: Closed-Loop Flow Velocities for Reduced-Order Model with 165o PhaseDifferential

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint Index

Vel

ocity

Figure 5–10: Closed-Loop Flow Velocities for Reduced-Order Model with 210o PhaseDifferential

of flow over the phase differential trajectory, which was shown in Figure 5–7. The

reduction along the center is roughly 80%.

The disturbance rejection is significant for the LPV controller with the reduced-

order models. These reductions confirm that the LPV controller will work not only

for reduced-order models at specific phase differentials but also over a time-varying

trajectory of phase differentials. The simulations did show some differences between

each of the reduced-order models. In particular, the amount of attenuation was slightly

less for the reduced-order model with a phase differential of 165o but much higher for

the reduced-order model with a phase differential of 210o. This decrease in attenuation

60

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint IndexV

eloc

ity

Figure 5–11: Closed-Loop Flow Velocities for Reduced-Order Model over a Trajectoryof Phase Differentials

seems almost contradictory considering that the open-loop simulations showed a

decrease in flow velocities for the same models.

5.9.3 Full-Order Closed-Loop Simulation

The simulations that were performed for the reduced-order models were repeated

using the full-order model. The reduced-order models are subspaces of this full-order

model so the performance of the controllers on the full-order model is actually of

predominant interest.

The measured velocities in response to an H∞ controller, created specifically for

the full-order model, are shown in Figure 5–12. Clearly, the magnitude of the velocity

is dramatically decreased below the open-loop level. The velocities in Figure 5–12

are several orders of magnitude less than the corresponding open-loop velocities in

Figure 5–3. This response will be used as a comparison for the responses from the

full-order model controlled by the LPV controller.

The velocities for the full-order model in response to the LPV controller with a

phase differential of 165o is shown in Figure 5–13. Though the velocities were not

reduced to the extent of the full-order simulation in Figure 5–12, they were reduced

by an amount comparable to the response shown by the reduced-order model at a 165o

61

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint IndexV

eloc

ity

Figure 5–12: Closed-Loop Flow Velocities for Full-Order Model

phase differential. The reduction in the velocity magnitude is evident along the center

of the cavity and is roughly 80%.

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint Index

Vel

ocity

Figure 5–13: Closed-Loop Flow Velocities for Full-Order Model with Controller Asso-ciated with 165o Phase Differential

The velocity magnitudes for the full-order model in response to the LPV controller

for a phase differential of 210o is shown in Figure 5–14. The reduction in velocity

compared to the open-loop flow of the full-order flow in Figure 5–3 is very clear.

The velocities along the centerline of the cavity were reduced by 60%. Though the

velocities were not as reduced as much as those in the simulation in Figure 5–12, the

velocities were reduced by an amount comparable to the reduced-order model at a 210o

phase differential.

62

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint IndexV

eloc

ity

Figure 5–14: Closed-Loop Flow Velocities for Full-Order Model with Controller Asso-ciated with 210o Phase Differential

The closed-loop simulation of the full-order model controlled by the LPV

controller over the phase differential trajectory is shown in Figure 5–15. The velocities

show a clear reduction in magnitude compared to the open-loop full-order flow in

Figure 5–3. The flow along the center of the cavity is reduced by 66%.

0

0.5

1

510

1520

−0.2

0

0.2

TimePoint Index

Vel

ocity

Figure 5–15: Closed-Loop Flow Velocities for Full-Order Model over a Trajectory ofPhase Differentials

The disturbance rejection is significant for both the reduced-order models and

the full-order model. These reductions confirm that the LPV controller, created for a

phase differential parameter, will work not only for the reduced-order models, which

are dependent on phase differential, but also for the full-order model.

63

5.10 Conclusion

Flow control is an exceedingly difficult challenge because of the nonlinearities

and time variations inherent to flow fields. These inherent difficulties can be avoided

when restricting the flow to creeping Stokes flow within a driven cavity. This project

has introduced a control methodology suitable for such a system. In particular, the

controllers are designed by considering subspaces of the flow field that describe modes

associated with phase differential between exogenous disturbances. The models of

these subspaces are realized as state-space systems and a controller can be designed

using the linear parameter-varying framework. The resulting controller is shown to

significantly decrease the flow velocities within the cavity for both the reduced-order

subspaces and also the full-order flow.

CHAPTER 6CONCLUSION

Practically all mechanical systems that involve motion need to be controlled with a

gain-scheduling technique. Aerospace systems in particular have the possibility to have

very extensive operating domains. Three specific aerospace systems were discussed in

this paper, the longitudinal dynamics of an

F/A-18, the structural dynamics of a hypersonic vehicle and the flow dynamics of a

driven cavity. The parameters that depicted the operating domain of the F/A-18 prob-

lem were altitude and Mach number. The parameter that depicted the operating domain

of the structure of the hypersonic vehicle was temperature and the operating domain

of the driven cavity was depicted by the phase differential within the fluid. This paper

has introduced a gain-scheduled control methodology, which uses H∞ synthesis to

create a linear parameter-varying controller, that is suitable for such systems. The LPV

controller created for the F/A-18 longitudingal dynamics proved to induce a pitch rate

for the aircraft that was similar to a designated target pitch rate. The LPV controller

created for the structural dynamics of a hypersonic aircraft successfully damped out

the vibrations induced by a temperature change. The LPV controller for the fluid

dynamics within a driven cavity significantly decreased the horizontal component of the

flow velocities along the centerline of the cavity for both the reduced-order subspaces

and the full order flow. The results of the control methodology to create proficient

controllers for three very different aerospace applications leads to the conclusion that

this methodology could be useful for other aerospace applications.

64

REFERENCES

[1] W. Rugh and J. Shamma, “Research on Gain Scheduling,” Automatica Vol. 36,Issue 10, John Hopkins University, Baltimore and UCLA, Los Angeles, October2000.

[2] D.J. Leith and W.E. Leithead, “Survey of Gain-Scheduling Analysis and Design,”International Journal of Control, Vol. 73, 2000, pp. 1001-1025.

[3] B. Bamieh and L. Giarre, “LPV Models: Identification for Gain SchedulingControl,” European Control Conference, 2001.

[4] L. Lee and M. Spillman, “Control of Slowly Varying LPV Systems: An Appli-cation to Flight Control,” AIAA Guidance, Navigation and Control Conference,AIAA-96-3805, 1996.

[5] J. Biannic, P. Apkarian and W. Garrard, “Parameter Varying Control of a HighPerformance Aircraft,” AIAA Guidance, Navigation and Control Conference,AIAA-96-3807, 1996.

[6] S. Bennani, D.M.C. Willemsen and C.W. Scherer, “Robust LPV Control WithBounded Parameter Rates,” Journal of Guidance, Control and Dynmaics, Vol. 21,No. 6, 1998, pp. 916-922.

[7] K. Tu and J.S. Shamma, “Nonlinear Gain-scheduled Control Design UsingSet-valued Methods,” American Control Conference, Vol. 2, 1998, pp. 1195-1199.

[8] J.S. Shamma and J.R. Cloutier, “Gain Scheduled Missile Autopilot Design UsingLinear Parameter Varying Transformations,” Journal of Guidance, Control andDynamics, Vol. 16, No. 2, 1993, pp. 256-263.

[9] R.A. Nichols, R.T. Reichert and W.J. Rugh, “Gain Scheduling for H-InfinityControllers: A Flight Control Example,” IEEE Transactions on Control SystemsTechnology, Vol. 1, No. 2, 1993, pp. 69-79.

[10] Z. Lin and M. Kammash, “Robust Gain-scheduled Aircraft Lontitudinal ControllerDesign using an H∞ Approach,” American Control Conference, 2001, pp. 2724-2729.

[11] S.M. Shahruz and S. Behtash, “Design of Controllers for Linear Parameter-Varying Systems by the Gain Scheduling Technique,” IEEE Conference onDecision and Control, Vol. 4, 1990, pp. 2490-2491.

65

66

[12] A. Sparks, “Linear Parameter Varying Control for a Tailless Aircraft,” AIAAGuidance, Navigation and Control Conference, AIAA-97-3636, 1997.

[13] L. Lee and M. Spillman, “Robust, Reduced-Order, Linear Parameter-VaryingFlight Control for an F-16,” AIAA Guidance, Navigation and Control Conference,AIAA-97-3637, 1997.

[14] R. Lind, J. Buffington, J. and A. Sparks, “Multi-loop Aeroservoelastic Controlof a Hypersonic Vehicle,” AIAA Guidance, Navigation and Control Conference,AIAA-99-4123, 1999.

[15] P. Gahinet and P. Apkarian, “A Linear Matrix Inequality Approach to H∞Control,” International Journal of Robust and Nonlinear Control, Vol. 4, 1995,pp. 421-448.

[16] J. Shin, “Optimal Blending Functions in Linear Parameter Varying ControlSynthesis for F-16 Aircraft,” American Control Conference, 2002, pp. 41-46.

[17] D. Malloy and B.C. Chang, “Stablilizing Controller Design for Linear ParameterVarying Systems Using Parameter Feedback,” AIAA Meeting Paper, AIAA-96-3808, July 1996.

[18] J. Mueller and G. Balas, “Impelmentattion and Testing of LPV Controllers forthe NASA F/A-18 Systems Research Aircraft,” AIAA Guidance, Navigation, andControl Conference, AIAA-2000-4446, 2000.

[19] P. Apkarian and P. Gahient, “Self-Scheduled H∞ Control of Missile via LinearMatrix Inequalities,” Journal of Guidance, Control and Dynamics, Vol. 18, No. 3,1995.

[20] G. Becker and A. Packard, “Robust Performance of Linear ParametericallyVarying Systems Using Parametrically-Dependent Linear Feedback,”Systems andControl Letters, Vol. 23, 1994, pp. 205-215.

[21] V. Verdult and M. Verhaegen, “Identification of Multivariable LPV State SpaceSystems By Local Gradiant Search,” European Control Conference, 2001,pp. 3675-3680.

[22] A. Nakajima and K. Tsumura, “Identification Methods for LPV MIMO Systems,”SICE Annual Conference, Vol. 2, 2002, pp. 1241-1245.

[23] M. Verhaegen and P. Dewilde, “Subspace Model Identification PartI: The Output-error State Space Model Identification Class of Algorithms,” International Journalof Control, Vol. 56, 1992, pp. 1187-1210.

[24] A. Kumar and M. Anderson, “A Comparison of LPV Modeling Techniques forAircraft Control,” AIAA Guidance Navigation and Control Conference, AIAA-2000-4458, 2000.

67

[25] F. Wu and A. Packard, “LQG Control Desgin for LPV Systems,” AmericanControl Conference, Vol. 6, June 1995, pp. 4440-4444.

[26] P. Apkarian, P. Gahinet and G. Becker, “Self-Scheduled H∞ Control of LinearParameter Varying Systems,” IEEE Conference on Decision and Control, Vol. 3,1994, pp. 2026-2031.

[27] P. Apkarian and P. Gahinet, “A Convex Characterization of Gain-Schedulede H∞Controllers,” IEEE Transactions on Automatic Control, Vol. 40, No. 5, 1995,pp. 853-864.

[28] R. Smith and A. Ahmed, “Robust Parmeterically Varying Attitude Controller De-sign for the X-33 Vehicle,” AIAA Guidance, Navigation and Control Conference,AIAA-2000-4158, 2000.

[29] R. Lind, “µ-Synthesis of an F/A-18 Controller,” AIAA Guidance, Navigation andControl Conference, AIAA-2000-4445, 2000.

[30] E. Jafarov and R. Tasaltin, “Design of Longitudinal Variable Structure FlightControl System for the F-18 Aircraft Model with Parameter Perturbations,” IEEEInternational Symposium on Computer Aided Control System Design, 1999,pp. 607-612.

[31] K.Tu, A. Sideris, K.D. Mease, J. Nathan and J. Carter, “Robust Lateral-DirectionalControl Design for the F/A-18,” AIAA Guidance, Navigation and ControlConference, AIAA-99-4204, Vol. 2, 1999, pp. 1213-1219.

[32] R. Lind, “Gain-Scheduled Approximations to H∞ Controllers for the F/A-18Active Aeroelastic Wing,” AIAA Guidance, Navigation and Control Conference,AIAA-99-4205, Vol. 2, 1999, pp. 1220-1230.

[33] G.J. Balas, J.B. Mueller and J. Barker, “Application of Gain-Scheduled Mul-tivariable Control Techniques to the F/A-18 System Research AIrcraft,” AIAAGuidance, Navigation and Control Conference, AIAA-99-4206, 1999.

[34] G.J. Balas, J.C. Doyle, K. Glover, A. Packard, and R. Smith, µ-Analysis andSynthesis Toolbox, The MathWorks, 2001.

[35] H. Buschek and A.J. Calise, “Fixed Order Robust Control Design for HypersonicVehicles,” AIAA Guidance, Navigation and Control Conference, AIAA-94-3662,1994.

[36] H. Buschek and A.J. Calise, “Robust Control of Hypersonic Vehicles ConsideringPropulsive and Aeroelastic Effects,” AIAA Guidance, Navigation and ControlConference, AIAA-93-3762, 1993.

[37] R. Lind, “Linear Parameter-Varying Modeling and Contol of Structural Dynamicswith Aerothermoelastic Effects,” Journal of Guidance, Control and Dynamics,Vol. 25, No.4, 2001, pp. 733-739.

68

[38] P. Vu and D.J. Biezad, “Longitudinal Control of Hypersonic Aircraft: An AlphaFollow-Up Scheme,” IEEE Regional Aerospace Control Systems Conference,1999, pp. 440-444.

[39] C.I. Marrison and R.F. Stengel, “Design of Robust Control Systems for aHypersonic Aircraft,” AIAA Guidance, Control of Dynamics Journal, Vol. 21,No. 1, 1998, pp. 58-63.

[40] Q. Wang and R.F. Stengel, “Robust Nonlinear Control of a Hypersonic Aircraft,”AIAA Guidance, Control and Dynamcis Journal, Vol. 23, No. 4, 2000, pp. 577-585.

[41] S.N. Balakrishnan, J. Shen and J.R. Grohs, “Hypersonic Vehicle TrajectoryOptimization and Control,” AIAA Guidance, Navigation and Control Conference,AIAA-97-3531, 1997.

[42] M. Sinai, “Hierarchical Design of Integrated Control for Hypersonic Vehicles,”IEEE Intelligent Control Symposium, Vol. 2, 1990, pp. 904-908.

[43] M. Heller, F. Hozapfel and G. Sachs, “Robust Lateral Control of HypersonicVehicles,” AIAA Guidance, Navigation and Control Conference, AIAA-2000-4248,2000.

[44] D.S. Naidu, S.S. Banda and J.L. Buffington, “Unified Approach to H2 and H∞Optimal Control of a Hypersonic Vehicle,” American Control Conference, Vol. 4,1999, pp. 2737-2741.

[45] F.R. Chavez and D.K. Schmidt, “An Integrated Analytical Aeropropul-sive/Analytical Model for the Dynamic Analysis of Hypersonic Vehicles,” AIAAAtmospheric Flight Mechanics Conference, AIAA-92-4567, 1992, pp. 551-563.

[46] D.K. Schmidt, H. Mamich and F.R. Chavez, “Dynamics and Control of Hyper-sonic Vehicles - The Integration Challenge for the 1990’s,” AIAA InternationalAerospace Planes Conference, AIAA-91-5057, 1991.

[47] F.R. Chavez and D.K Schmidt, “Analytical Aeropropulsive/AeroelasticHypersonic-Vehicle Model with Dynamic Analysis,” AIAA Guidance, Controland Dynamics Journal, Vol. 17, No. 6, 1994, pp. 1308-1319.

[48] F.R. Chavez and D.K. Schmidt, “Flight Dynamics and Control of Elastic Hyerp-sonic Vehicles: Modeling Uncertainties,” AIAA Guidance, Navigation and ControlConference, AIAA Paper 94-3629, 1994.

[49] J. Heeg, M.G. Gilbert and A.S. Pototzky, “Active Control of AerothermoelasticEffects for a Conceptual Hypersonic Aircraft,” Journal of Aircraft, Vol. 30, No. 4,July-August 1993, pp. 453-458.

[50] P. Gahinet, A. Nemirovski, A. Laub, M. Chilali, LMI Control Toolbox - UsersGuide, The MathWorks, Inc. Natick, MA, 1995.

69

[51] H. Chih-Ming, “Control of Fluid Flows by Micro Transducers,” IEEE Interna-tional Symposium on Micro Machine and Human Science, 1996, pp. 29-33.

[52] J. Ortiz and A. Barhorst, “Modeling Fluid-Structure Interaction,” Journal ofGuidance, Control and Dynamics, Vol. 20, No. 6, 1997, pp. 1221-1228.

[53] K. Ito and S.S. Ravindran, “A Reduced Order Method for Control of FluidFlows,” IEEE Conference on Decision and Control, Vol. 4, 1996, pp. 3705-3710.

[54] Y. Feng, A.J. Kurdila, R. Lind and D.W. Mikolaitis, “Linear Parameter-VaryingFlow Control for a Driven Cavity using Reduced-Order Models,” AIAA Guidance,Navigation and Control Conference, AIAA-2003-5351, 2003.

[55] J. Lumley, “The Structure of Inhomogeneous Turbulent Flows,” AtmopshericTurbulence and Radio Wave Propagation, edited by A.M. Yaglom and V.I.Tararsky, Nauka, Moscow, 1967, pp. 166-178.

[56] J.A. Burns and O. Yuh-Roung, “Feedback Control of the Driven Cavity ProblemUsing LQR Designs,” IEEE Conference on Decision and Control, Vol. 1, 1994,pp. 289-294.

[57] L. Baramov, O. Tutty and E. Rogers, “Robust Control of Linearized PoiseuilleFlow,” Journal of Guidance, Control and Dynamics, Vol. 25, No. 1, 2002,pp. 145-151.

[58] H.T. Banks, K. Ito, “Structural Actuator Control of Fluid/Structure Interactions,”IEEE Conference on Decision and Control, Vol. 1, 1994, pp. 283-288.

[59] O.K. Rediniotis, J. Ko, X. Yue, A.J. Kurdila, “Synthetic Jets, their Reduced OrderModeling and Applications to Flow Control,” AIAA Aerospace Sciences Meetingand Exhibit, AIAA-99-1000, 1999.

[60] S.S. Ravindran, “Reduced-Order Adaptive Controllers for MHD Flows UsingProper Orthogonal Decomposition,” IEEE Conference on Decision and Control,Vol. 3, 2001, pp. 2454-2459.

[61] L. Ji and J. Zhou, “The Boundary Element Method For Boundary Control of TheLinear Stokes Flow,” IEEE Conference on Decision and Control, Vol. 3, 1990,pp. 1192-1194.

[62] C.F.M. Coimbra and R.H. Rangel, “Spherical Particle Motion in Unsteady ViscousFlows,” AIAA Aerospace Science Meeting and Exhibit, 1999, AIAA-99-1031.

[63] K. Zhou, J.C. Doyle and K. Glover, “Robust and Optimal Control,” Prentice Hall,1st Edition, 1997.

BIOGRAPHICAL SKETCH

Kristin Fitzpatrick was born in Blue Hill, Maine on March 26, 1980. Her

family moved to Florida in 1988 after the death of her mother. She received her

high school diploma from the Center for Advanced Technologies, a magnet program

in St. Petersburg, Florida. She then attended the University of Florida and received a

degree in Aerospace Engineering with Honors in December 2002. She has worked with

the aerospace dynamics and control research group under the direction of Dr. Rick

Lind and Dr. Andy Kurdila and is projected to receive her Master of Science degree in

aerospace engineering in December 2003. She will stay at the University of Florida to

pursue a doctorate in aerospace engineering with the focus in dynamics and control.

70