09.03.2014 Views

Mathematical Models for Control of Aircraft and Satellites - NTNU

Mathematical Models for Control of Aircraft and Satellites - NTNU

Mathematical Models for Control of Aircraft and Satellites - NTNU

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

MATHEMATICAL MODELS FOR CONTROL OF<br />

AIRCRAFT AND SATELLITES<br />

THOR I. FOSSEN<br />

Pr<strong>of</strong>essor <strong>of</strong> Guidance, Navigation <strong>and</strong> <strong>Control</strong><br />

Centre <strong>for</strong> Autonomous Marine Operations <strong>and</strong> Systems<br />

Department <strong>of</strong> Engineering Cybernetics<br />

Norwegian University <strong>of</strong> Science <strong>and</strong> Technology<br />

April 2013<br />

3rd edition<br />

Copyright c○1998-2013 Department <strong>of</strong> Engineering Cybernetics, <strong>NTNU</strong>.<br />

1st edition published February 1998.<br />

2nd edition published January 2011.<br />

3rd edition published April 2013.


Contents<br />

Figures<br />

iii<br />

1 Introduction 1<br />

2 <strong>Aircraft</strong> Modeling 3<br />

2.1 Definition <strong>of</strong> <strong>Aircraft</strong> State-Space Vectors . . . . . . . . . . . . . . . . . . . . 3<br />

2.2 Body-Fixed Coordinate Systems <strong>for</strong> <strong>Aircraft</strong> . . . . . . . . . . . . . . . . . . 4<br />

2.2.1 Rotation matrices <strong>for</strong> wind <strong>and</strong> stability axes . . . . . . . . . . . . . 5<br />

2.3 <strong>Aircraft</strong> Equations <strong>of</strong> Motion . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.3.1 Kinematic equations <strong>for</strong> translation . . . . . . . . . . . . . . . . . . . 6<br />

2.3.2 Kinematic equations <strong>for</strong> attitude . . . . . . . . . . . . . . . . . . . . 7<br />

2.3.3 Rigid-body kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

2.3.4 Sensors <strong>and</strong> measurement systems . . . . . . . . . . . . . . . . . . . . 8<br />

2.4 Perturbation Theory (Linear Theory) . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.4.1 Definition <strong>of</strong> nominal <strong>and</strong> perturbation values . . . . . . . . . . . . . 9<br />

2.4.2 Linearization <strong>of</strong> the rigid-body kinetics . . . . . . . . . . . . . . . . . 9<br />

2.4.3 Linear state-space model based using wind <strong>and</strong> stability axes . . . . . 11<br />

2.5 Decoupling in Longitudinal <strong>and</strong> Lateral Modes . . . . . . . . . . . . . . . . . 12<br />

2.5.1 Longitudinal equations . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

2.5.2 Lateral equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

2.6 Aerodynamic Forces <strong>and</strong> Moments . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.6.1 Longitudinal aerodynamic <strong>for</strong>ces <strong>and</strong> moments . . . . . . . . . . . . . 16<br />

2.6.2 Lateral aerodynamic <strong>for</strong>ces <strong>and</strong> moments . . . . . . . . . . . . . . . . 16<br />

2.7 Propeller Thrust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.8 Aerodynamic Coeffi cients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

2.8.1 Aerodynamic <strong>for</strong>ces <strong>and</strong> moments . . . . . . . . . . . . . . . . . . . . 19<br />

2.8.2 Drag coeffi cient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.8.3 Lift coeffi cient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.8.4 Side<strong>for</strong>ce coeffi cient . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.8.5 Rolling moment coeffi cient . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.8.6 Pitching moment coeffi cient . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2.8.7 Yawing moment coeffi cient . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.8.8 Equations <strong>of</strong> motion including aerodynamic coeffi cients . . . . . . . . 23<br />

2.9 St<strong>and</strong>ard <strong>Aircraft</strong> Maneuvers . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

i


ii<br />

CONTENTS<br />

2.9.1 Dynamic equation <strong>for</strong> coordinated turn (bank-to-turn) . . . . . . . . 24<br />

2.9.2 Dynamic equation <strong>for</strong> altitude control . . . . . . . . . . . . . . . . . . 25<br />

2.10 <strong>Aircraft</strong> Stability Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.10.1 Longitudinal stability analysis . . . . . . . . . . . . . . . . . . . . . . 27<br />

2.10.2 Lateral stability analysis . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.11 Design <strong>of</strong> flight control systems . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3 Satellite Modeling 31<br />

3.1 Attitude Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.1.1 Euler’s 2nd Axiom applied to satellites . . . . . . . . . . . . . . . . . 31<br />

3.1.2 Skew-symmetric representation <strong>of</strong> the satellite model . . . . . . . . . 32<br />

3.2 Actuator Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

3.3 Design <strong>of</strong> Satellite Attitude <strong>Control</strong> Systems . . . . . . . . . . . . . . . . . . 34<br />

3.3.1 Nonlinear quaternion set-point regulator . . . . . . . . . . . . . . . . 34<br />

3.3.2 PD-controller <strong>of</strong> Wen <strong>and</strong> Kreutz-Delago (1991) . . . . . . . . . . . . 35<br />

4 Matlab Simulation <strong>Models</strong> 37<br />

4.1 Boeing-767 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

4.1.1 Longitudinal model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

4.1.2 Lateral model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

4.2 F-16 Fighter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

4.2.1 Longitudinal model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

4.3 F2B Bristol Fighter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

4.3.1 Lateral model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

4.4 <strong>NTNU</strong> student cube-satellite . . . . . . . . . . . . . . . . . . . . . . . . . . . 41


List <strong>of</strong> Figures<br />

1.1 Sketch showing a modern fighter aircraft (Stevens <strong>and</strong> Lewis 1992). . . . . . 1<br />

2.1 Definition <strong>of</strong> aircraft body axes, generalized velocities, <strong>for</strong>ces, moments <strong>and</strong><br />

Euler angles (McLean 1990). . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

2.2 Definition <strong>of</strong> stability <strong>and</strong> wind axes <strong>for</strong> an aircraft (Stevens <strong>and</strong> Lewis 1992). 5<br />

2.3 <strong>Control</strong> inputs <strong>for</strong> conventional aircraft. Notice that the two ailerons can be<br />

controlled by using one control input: δ A = 1/2(δ AL + δ AR ). . . . . . . . . . 14<br />

2.4 Lift <strong>and</strong> drag as a function <strong>of</strong> angle <strong>of</strong> attack. . . . . . . . . . . . . . . . . . 22<br />

2.5 <strong>Aircraft</strong> longitudinal eigenvalue configuration plotted in the complex plane. . 28<br />

3.1 The NUTS CubeSat project at <strong>NTNU</strong>- Courtesy to http://nuts.cubesat.no/ 31<br />

4.1 Schematic drawing <strong>of</strong> the Bristol F.2B Fighter (McRuer et al 1973). . . . . . 40<br />

iii


iv<br />

LIST OF FIGURES


Chapter 1<br />

Introduction<br />

This booklet is a collection <strong>of</strong> lecture notes used in the course TTK4109 Guidance <strong>and</strong> <strong>Control</strong><br />

<strong>of</strong> Vehicles, which is given at the Department <strong>of</strong> Engineering Cybernetics, <strong>NTNU</strong>. The<br />

kinematic <strong>and</strong> kinetic equations <strong>of</strong> a marine craft (ships, high-speed craft <strong>and</strong> underwater<br />

vehicles) can be modified to describe aircraft <strong>and</strong> satellites by minor adjustments <strong>of</strong> notation<br />

<strong>and</strong> assumptions. Consequently, the vectorial notation introduced in the two Wiley<br />

textbooks:<br />

Fossen, T. I. (1994). Guidance <strong>and</strong> <strong>Control</strong> <strong>of</strong> Ocean Vehicles<br />

Fossen, T. I. (2011). H<strong>and</strong>book <strong>of</strong> Marine Craft Hydrodynamics <strong>and</strong> Motion <strong>Control</strong><br />

is used to describe the aircraft <strong>and</strong> satellite equations <strong>of</strong> motion.<br />

Figure 1.1: Sketch showing a modern fighter aircraft (Stevens <strong>and</strong> Lewis 1992).<br />

The booklet is organized according to:<br />

• Chapter 2: <strong>Aircraft</strong> Modeling<br />

• Chapter 3: Satellite Modeling<br />

• Chapter 4: Matlab Simulation <strong>Models</strong><br />

1


2 CHAPTER 1. INTRODUCTION<br />

Other useful references on flight control are:<br />

Blakelock, J. H. (1991). <strong>Aircraft</strong> <strong>and</strong> Missiles (John Wiley & Sons Ltd.)<br />

Etkin, B. <strong>and</strong> L. D. Reid (1996). Dynamics <strong>of</strong> Flight: Stability <strong>and</strong> <strong>Control</strong><br />

Wiley & Sons Ltd.)<br />

(John<br />

Fortescue, P., G. Swinerd <strong>and</strong> J. Stark (2011). Spacecraft Systems Engineering (John<br />

Wiley & Sons Ltd.)<br />

Hughes, P. C. (1986). Spacecraft Attitude Dynamics (John Wiley & Sons Ltd.)<br />

McLean, D. (1990). Automatic Flight <strong>Control</strong> Systems (Prentice Hall Inc.)<br />

McRuer, D., D. Ashkenas <strong>and</strong> A. I. Graham (1973). <strong>Aircraft</strong> Dynamics <strong>and</strong> Automatic<br />

<strong>Control</strong> (Princeton University Press)<br />

Nelson, R. C. (1998). Flight Stability <strong>and</strong> Automatic <strong>Control</strong> (McGraw-Hill Int.)<br />

Roskam, J. (1999). Airplane Flight Dynamics <strong>and</strong> Automatic Flight <strong>Control</strong>s (Darcorporation)<br />

Schmidt, D. K. (2012). Modern Flight Dynamics (McGraw-Hill Int.)<br />

Stevens, B. L. <strong>and</strong> F. L. Lewis (2003). <strong>Aircraft</strong> <strong>Control</strong> <strong>and</strong> Simulation (John Wiley<br />

& Sons Ltd.)<br />

In<strong>for</strong>mation about the graduate coursesTTK4109 Guidance <strong>and</strong> <strong>Control</strong> <strong>of</strong> Vehicles <strong>and</strong><br />

TK8109 Advanced Topics in Guidance <strong>and</strong> Navigation are found on the Wiki-pages<br />

URL: http://www.itk.ntnu.no/emner/ttk4190<br />

URL: http://www.itk.ntnu.no/emner/tk8109<br />

Thor I. Fossen<br />

Trondheim —3 April 2013


Chapter 2<br />

<strong>Aircraft</strong> Modeling<br />

This chapter gives an introduction to aircraft modeling. The equations <strong>of</strong> motion are linearized<br />

using perturbation theory <strong>and</strong> the final results are state-space models <strong>for</strong> the longitudinal<br />

<strong>and</strong> lateral motions. The models can be used <strong>for</strong> aircraft simulation <strong>and</strong> design <strong>of</strong><br />

flight control systems.<br />

2.1 Definition <strong>of</strong> <strong>Aircraft</strong> State-Space Vectors<br />

The aircraft generalized velocity vector is defined according to (see Figure 2.1):<br />

2<br />

ν :=<br />

6<br />

4<br />

2<br />

η :=<br />

6<br />

4<br />

U<br />

V<br />

W<br />

P<br />

Q<br />

R<br />

3<br />

X E<br />

Y E<br />

Z E , h<br />

Φ<br />

Θ<br />

Ψ<br />

2<br />

=<br />

7 6<br />

5 4<br />

3<br />

=<br />

7 6<br />

5 4<br />

longitudinal (<strong>for</strong>ward) velocity<br />

lateral (transverse) velocity<br />

vertical velocity<br />

roll rate<br />

pitch rate<br />

yaw rate<br />

2<br />

Earth-fixed x-position<br />

Earth-fixed y-position<br />

Earth-fixed z-position (axis downwards), altitude<br />

roll angle<br />

pitch angle<br />

yaw angle<br />

3<br />

7<br />

5<br />

3<br />

7<br />

5<br />

(2.1)<br />

(2.2)<br />

Forces <strong>and</strong> moments are defined in a similar manner:<br />

2 3 2<br />

X longitudinal <strong>for</strong>ce<br />

Y<br />

transverse <strong>for</strong>ce<br />

Z<br />

6 L<br />

:=<br />

vertical <strong>for</strong>ce<br />

7 6 roll moment<br />

4 M 5 4 pitch moment<br />

N yaw moment<br />

3<br />

7<br />

5<br />

(2.3)<br />

Comment 1: Notice that the capital letters L, M, N <strong>for</strong> the moments are different from<br />

3


4 CHAPTER 2. AIRCRAFT MODELING<br />

Figure 2.1: Definition <strong>of</strong> aircraft body axes, generalized velocities, <strong>for</strong>ces, moments <strong>and</strong> Euler<br />

angles (McLean 1990).<br />

those used <strong>for</strong> marine craft—that is, K, M, N. The reason <strong>for</strong> this is that L is reserved as<br />

length parameter <strong>for</strong> ships <strong>and</strong> underwater vehicles.<br />

Comment 2: For aircraft it is common to use capital letters <strong>for</strong> the states U, V, W, etc.<br />

while it is common to use small letters <strong>for</strong> marine craft.<br />

2.2 Body-Fixed Coordinate Systems <strong>for</strong> <strong>Aircraft</strong><br />

For aircraft it is common to use the following body-fixed coordinate systems:<br />

• BODY axes, superscript b<br />

• STABILITY axes, superscript s<br />

• WIND axes, superscript w<br />

The axis systems are shown in Figure 2.2 where the angle <strong>of</strong> attack α <strong>and</strong> sideslip angle β<br />

are defined as:<br />

tan(α) := W U<br />

(2.4)<br />

sin(β) := V V T<br />

(2.5)<br />

where<br />

V T = √ U 2 + V 2 + W 2 (2.6)


2.2. BODY-FIXED COORDINATE SYSTEMS FOR AIRCRAFT 5<br />

Figure 2.2: Definition <strong>of</strong> stability <strong>and</strong> wind axes <strong>for</strong> an aircraft (Stevens <strong>and</strong> Lewis 1992).<br />

is the total speed <strong>of</strong> the aircraft. Aerodynamic effects are classified according to the Mach<br />

number:<br />

M := V T<br />

a<br />

(2.7)<br />

where a = 340 m/s = 1224 km/h is the speed <strong>of</strong> sound in air at a temperature <strong>of</strong> 20 o C on<br />

the ocean surface. The following terminology is used <strong>for</strong> varying speed:<br />

Subsonic speed M < 1.0<br />

Transonic speed 0.8 ≤ M ≤ 1.2<br />

Supersonic speed 1.0 ≤ M ≤ 5.0<br />

Hypersonic speed 5.0 ≤ M<br />

An aircraft will break the sound barrier at M = 1.0 <strong>and</strong> this is clearly heard as a sharp<br />

crack. If you fly at low altitude <strong>and</strong> break the sound barrier, windows in building will break<br />

due to pressure-induced waves.<br />

2.2.1 Rotation matrices <strong>for</strong> wind <strong>and</strong> stability axes<br />

The relationship between vectors expressed in different coordinate systems can be derived<br />

using rotation matrices. The BODY axes are first rotated a negative sideslip angle −β about<br />

the z-axis. The new coordinate system is then rotated a positive angle <strong>of</strong> attack α about<br />

the new y-axis such that the resulting x-axis points in the direction <strong>of</strong> the total speed V T .<br />

The first rotation defines the WIND axes while the second rotation defines the STABILITY


6 CHAPTER 2. AIRCRAFT MODELING<br />

axes. This can be mathematically expressed as:<br />

2<br />

cos(β) sin(β)<br />

3<br />

0<br />

p w = R z,−β p s = 4 − sin(β) cos(β) 0 5 p s (2.8)<br />

2<br />

0 0 1<br />

3<br />

cos(α) 0 sin(α)<br />

p s = R y,α p b = 4 0 1 0 5 p b (2.9)<br />

− sin(α) 0 cos(α)<br />

The rotation matrix becomes:<br />

Hence,<br />

R w b = R z,−β R y,α (2.10)<br />

p w = R w b p b (2.11)<br />

⇕<br />

p w =<br />

⇕<br />

p w =<br />

2<br />

4<br />

2<br />

4<br />

cos(β) sin(β) 0<br />

− sin(β) cos(β) 0<br />

0 0 1<br />

3 2<br />

5 4<br />

cos(α) 0 sin(α)<br />

0 1 0<br />

− sin(α) 0 cos(α)<br />

cos(α) cos(β) sin(β) sin(α) cos(β)<br />

− cos(α) sin(β) cos(β) − sin(α) sin(β)<br />

− sin(α) 0 cos(α)<br />

3<br />

3<br />

5 p b (2.12)<br />

5 p b (2.13)<br />

This gives the following relationship between the velocities in body <strong>and</strong> wind axes:<br />

2<br />

v b = 4<br />

Consequently,<br />

U<br />

V<br />

W<br />

3<br />

2<br />

5 = (R w b ) ⊤ v w = R ⊤ y,αR ⊤ 4<br />

z,−β<br />

V T<br />

0<br />

0<br />

3<br />

2<br />

5 = 4<br />

U = V T cos(α) cos(β)<br />

V = V T sin(β)<br />

W = V T sin(α) cos(β)<br />

V T cos(α) cos(β)<br />

V T sin(β)<br />

V T sin(α) cos(β)<br />

3<br />

5 (2.14)<br />

(2.15)<br />

2.3 <strong>Aircraft</strong> Equations <strong>of</strong> Motion<br />

2.3.1 Kinematic equations <strong>for</strong> translation<br />

The kinematic equations <strong>for</strong> translation <strong>and</strong> rotation <strong>of</strong> a body-fixed coordinate system<br />

BODY with respect to a local geographic coordinate system NED (North-East-Down) can<br />

be expressed in terms or the Euler angles:<br />

2<br />

4<br />

Ẏ E<br />

ŻE<br />

5 = R n b<br />

Ẋ E<br />

3<br />

2<br />

4<br />

U<br />

V<br />

W<br />

3<br />

2<br />

5 = R z,Ψ R y,Θ R x,Φ<br />

4<br />

U<br />

V<br />

W<br />

3<br />

5 (2.16)


2.3. AIRCRAFT EQUATIONS OF MOTION 7<br />

Exp<strong>and</strong>ing this expression gives:<br />

2 3 2<br />

Ẋ E cΨ −sΨ 0<br />

4 Ẏ E<br />

5 = 4 sΨ cΨ 0<br />

ŻE 0 0 1<br />

2<br />

4<br />

3 2<br />

5 4<br />

cΘ 0 sΘ<br />

0 1 0<br />

−sΘ 0 cΘ<br />

3 2<br />

5 4<br />

1 0 0<br />

0 cΦ −sΦ<br />

0 sΦ cΦ<br />

3 2<br />

5 4<br />

⇕ (2.17)<br />

3 2<br />

3 2 3<br />

Ẋ E cΨcΘ −sΨcΦ + cΨsΘsΦ sΨsΦ + cΨcΦsΘ U<br />

Ẏ E<br />

5 = 4 sΨcΘ cΨcΦ + sΦsΘsΨ −cΨsΦ + sΘsΨcΦ 5 4 V 5<br />

ŻE −sΘ cΘsΦ cΘcΦ<br />

W<br />

2.3.2 Kinematic equations <strong>for</strong> attitude<br />

U<br />

V<br />

W<br />

3<br />

5<br />

The attitude is given by:<br />

2<br />

P<br />

3 2<br />

4 Q 5 = 4<br />

R<br />

˙Φ<br />

0<br />

0<br />

3 2<br />

5 + R ⊤ 4<br />

x,Φ<br />

0<br />

˙Θ<br />

0<br />

3<br />

2<br />

5 + R ⊤ x,ΦR ⊤ 4<br />

y,Θ<br />

0<br />

0<br />

˙Ψ<br />

3<br />

5 (2.18)<br />

which gives:<br />

2<br />

4<br />

˙Φ<br />

˙Θ<br />

˙Ψ<br />

3<br />

2<br />

5 = 4<br />

1 sΦtΘ cΦtΘ<br />

0 cΦ −sΦ<br />

0 sΦ/cΘ cΦ/cΘ<br />

3 2<br />

5 4<br />

P<br />

Q<br />

R<br />

3<br />

5 , cΘ ≠ 0 (2.19)<br />

2.3.3 Rigid-body kinetics<br />

The aircraft rigid-body kinetics can be expressed as (Fossen 1994, 2011):<br />

m( ˙ν 1 + ν 2 × ν 1 ) = τ 1 (2.20)<br />

I CG ˙ν 2 + ν 2 × (I CG ν 2 ) = τ 2 (2.21)<br />

where ν 1 := [U, V, W ] T , ν 2 := [P, Q, R] T , τ 1 := [X, Y, Z] T <strong>and</strong> τ 2 := [L, M, N] T . It is assumed<br />

that the coordinate system is located in the aircraft center <strong>of</strong> gravity (CG). The<br />

resulting model is written:<br />

M RB ˙ν + C RB (ν)ν = τ RB (2.22)<br />

where<br />

M RB =<br />

<br />

<br />

mI3×3 O 3×3<br />

mS(ν2 ) O<br />

, C<br />

O 3×3 I RB (ν) =<br />

3×3<br />

CG O 3×3 −S(I CG<br />

ν 2 )<br />

<br />

(2.23)<br />

3<br />

The inertia tensor is defined as (assume that I xy = I yz = 0 which corresponds to xz−plane<br />

symmetry):<br />

2<br />

I x 0 −I xz<br />

−I xz 0 I z<br />

I CG := 4 0 I y 0 5 (2.24)<br />

The <strong>for</strong>ces <strong>and</strong> moments acting on the aircraft can be expressed as:<br />

τ RB = −g(η) + τ (2.25)


8 CHAPTER 2. AIRCRAFT MODELING<br />

where τ is a generalized vector that includes aerodynamic <strong>and</strong> control <strong>for</strong>ces. The gravitational<br />

<strong>for</strong>ce f G = [0 0 mg] T acts in the CG (origin <strong>of</strong> the body-fixed coordinate system) <strong>and</strong><br />

this gives the following vector expressed in NED:<br />

2<br />

3<br />

mg sin(Θ)<br />

<br />

−mg cos(Θ) sin(Φ)<br />

g(η) = −(R n fG<br />

b )⊤ =<br />

−mg cos(Θ) cos(Φ)<br />

O 3×1 6 0<br />

(2.26)<br />

7<br />

4 0 5<br />

0<br />

Hence, the aircraft model can be written in matrix <strong>for</strong>m as:<br />

or in component <strong>for</strong>m:<br />

M RB ˙ν + C RB (ν)ν + g(η) = τ (2.27)<br />

m( ˙U + QW − RV + g sin(Θ)) = X<br />

m( ˙V + UR − W P − g cos(Θ) sin(Φ)) = Y<br />

m(Ẇ + V P − QU − g cos(Θ) cos(Φ)) = Z<br />

I x P˙<br />

− I xz (Ṙ + P Q) + (I z − I y )QR = L (2.28)<br />

I y ˙Q + I xz (P 2 − R 2 ) + (I x − I z )P R = M<br />

I z Ṙ − I xz P ˙ + (I y − I x )P Q + I xz QR = N<br />

2.3.4 Sensors <strong>and</strong> measurement systems<br />

It is common that aircraft sensor systems are equipped with three accelerometers. If the<br />

accelerometers are located in the CG, the measurement equations take the following <strong>for</strong>m:<br />

a xCG = X m = ˙U + QW − RV + g sin(Θ) (2.29)<br />

a yCG = Y m = ˙V + UR − W P − g cos(Θ) sin(Φ) (2.30)<br />

a zCG = Z m<br />

= Ẇ + V P − QU − g cos(Θ) cos(Φ) (2.31)<br />

In addition to these sensors, an aircraft is equipped with gyros, magnetometers <strong>and</strong> a sensor<br />

<strong>for</strong> altitude h <strong>and</strong> wind speed V T . These sensors are used in inertial navigation systems (INS)<br />

which again use a Kalman filter to compute estimates <strong>of</strong> U, V, W, P, Q <strong>and</strong> R as well as the<br />

Euler angles Φ,Θ <strong>and</strong> Ψ. Other measurement systems that are used onboard aircraft are<br />

global navigation satellite systems (GNSS), radar <strong>and</strong> sensors <strong>for</strong> angle <strong>of</strong> attack.<br />

2.4 Perturbation Theory (Linear Theory)<br />

The nonlinear equations <strong>of</strong> motion can be linearized by using perturbation theory. This is<br />

illustrated below.


2.4. PERTURBATION THEORY (LINEAR THEORY) 9<br />

2.4.1 Definition <strong>of</strong> nominal <strong>and</strong> perturbation values<br />

According to linear theory it is possible to write the states as the sum <strong>of</strong> a nominal value<br />

(usually constant) <strong>and</strong> a perturbation (deviation from the nominal value). Moreover,<br />

Total state = Nominal value + Perturbation<br />

The following definitions are made:<br />

2 3 2 3<br />

2<br />

X 0 δX<br />

Y 0<br />

δY<br />

τ := τ 0 + δτ =<br />

Z 0<br />

6 L 0<br />

+<br />

δZ<br />

7 6 δL<br />

, ν := ν 0 + δν =<br />

7<br />

6<br />

4 M 0<br />

5 4 δM 5<br />

4<br />

N 0 δN<br />

3<br />

U 0<br />

V 0<br />

W 0<br />

P 0<br />

7<br />

Q 0<br />

5<br />

R 0<br />

2<br />

+<br />

6<br />

4<br />

u<br />

v<br />

w<br />

p<br />

q<br />

r<br />

3<br />

7<br />

5<br />

(2.32)<br />

Similar, the angles are defined according to:<br />

2 3 2<br />

Θ<br />

4 Φ 5 := 4<br />

Ψ<br />

3 2<br />

Θ 0<br />

Φ 0<br />

5 + 4<br />

Ψ 0<br />

θ<br />

φ<br />

ψ<br />

3<br />

5 (2.33)<br />

Consequently, a linearized state-space model will consist <strong>of</strong> the following states u, v, w, p, q, r, θ, φ<br />

<strong>and</strong> ψ.<br />

2.4.2 Linearization <strong>of</strong> the rigid-body kinetics<br />

The rigid-body kinetics can be linearized by using perturbation theory.<br />

Equilibrium condition<br />

If the aerodynamic <strong>for</strong>ces <strong>and</strong> moments, velocities, angles <strong>and</strong> control inputs are expressed<br />

as nominal values <strong>and</strong> perturbations τ = τ 0 + δτ , ν = ν 0 + δν <strong>and</strong> η = η 0 + δη, the aircraft<br />

equilibrium point will satisfy (it is assumed that ˙ν 0 = 0):<br />

This can be exp<strong>and</strong>ed according to:<br />

C RB (ν 0 )ν 0<br />

+ g(η 0<br />

) = τ 0 (2.34)<br />

m(Q 0 W 0 − R 0 V 0 + g sin(Θ 0 )) = X 0<br />

m(U 0 R 0 − P 0 W 0 − g cos(Θ 0 ) sin(Φ 0 )) = Y 0<br />

m(P 0 V 0 − Q 0 U 0 − g cos(Θ 0 ) cos(Φ 0 )) = Z 0<br />

(I z − I y )Q 0 R 0 − P 0 Q 0 I xz = L 0 (2.35)<br />

(P 2 0 − R 2 0)I xz + (I x − I z )P 0 R 0 = M 0<br />

(I y − I x )P 0 Q 0 + Q 0 R 0 I xz = N 0


10 CHAPTER 2. AIRCRAFT MODELING<br />

Perturbed equations<br />

The perturbed equations—that is, the linearized equations <strong>of</strong> motion are usually derived by<br />

a 1st-order Taylor series expansion about the nominal values. Alternatively, it is possible<br />

to substitute (2.32) <strong>and</strong> (2.35) into (2.27) <strong>and</strong> neglect higher-order terms <strong>of</strong> the perturbed<br />

states. This is illustrated <strong>for</strong> the first degree <strong>of</strong> freedom (DOF):<br />

Example 1 (Linearization <strong>of</strong> surge using perturbation theory)<br />

m[ ˙U + QW − RV + g sin(Θ)] = X<br />

⇓ (2.36)<br />

m[ ˙U 0 + ˙u + (Q 0 + q)(W 0 + w) − (R 0 + r)(V 0 + v) + g sin(Θ 0 + θ)] = X 0 + δX<br />

This can be written:<br />

sin(Θ 0 + θ) = sin(Θ 0 ) cos(θ) + cos(Θ 0 ) sin(θ) θ small<br />

≈ sin(Θ 0 ) + cos(Θ 0 )θ (2.37)<br />

Since ˙U 0 = 0 <strong>and</strong><br />

m(Q 0 W 0 − R 0 V 0 + g sin(Θ 0 )) = X 0 (2.38)<br />

Equation (2.36) is reduced to:<br />

m[ ˙u + Q 0 w + W 0 q + wq − R 0 v − V 0 r − vr + g cos(Θ 0 )θ] = δX (2.39)<br />

If it is assumed that the 2nd-order terms wq <strong>and</strong> vr are negligible, the linearized model<br />

becomes:<br />

□<br />

m[ ˙u + Q 0 w + W 0 q − R 0 v − V 0 r + g cos(Θ 0 )θ] = δX (2.40)<br />

Linear state-space model <strong>for</strong> aircraft<br />

If all DOFs are linearized, the following state-space model is obtained:<br />

m[ ˙u + Q 0 w + W 0 q − R 0 v − V 0 r + g cos(Θ 0 )θ] = δX<br />

m[ ˙v + U 0 r + R 0 u − W 0 p − P 0 w − g cos(Θ 0 ) cos(Φ 0 )φ + g sin(Θ 0 ) sin(Φ 0 )θ] = δY<br />

m[ẇ + V 0 p + P 0 v − U 0 q − Q 0 u + g cos(Θ 0 ) sin(Φ 0 )φ + g sin(Θ 0 ) cos(Φ 0 )θ] = δZ<br />

I x ṗ − I xz ṙ + (I z − I y )(Q 0 r + R 0 q) − I xz (P 0 q + Q 0 p) = δL (2.41)<br />

I y ˙q + (I x − I z )(P 0 r + R 0 p) − 2I xz (R 0 r + P 0 p) = δM<br />

I z ṙ − I xz ṗ + (I y − I x )(P 0 q + Q 0 p) + I xz (Q 0 r + R 0 q) = δN<br />

This can be expressed in matrix <strong>for</strong>m as:<br />

M RB ˙ν + N RB ν + Gη = τ (2.42)


2.4. PERTURBATION THEORY (LINEAR THEORY) 11<br />

where<br />

M RB =<br />

N RB =<br />

G =<br />

2<br />

6<br />

4<br />

2<br />

6<br />

4<br />

2<br />

6<br />

4<br />

3<br />

m 0 0<br />

m 0 0 3×3<br />

0 0 m<br />

I x 0 −I xz<br />

7<br />

0 3×3 0 I y 0 5<br />

−I xz 0 I z<br />

3<br />

0 −mR 0 mQ 0 0 mW 0 −mV 0<br />

mR 0 0 −P 0 −W 0 0 U 0<br />

−mQ 0 mP 0 0 mV 0 −mU 0 0<br />

−I xz Q 0 (I z − I y )R 0 − I xz P 0 (I z − I y )Q 0<br />

7<br />

0 3×3 (I x − I z )R 0 −2I xz P 0 (I x − I z )P 0 − 2I xz R 0<br />

5<br />

(I y − I x )Q 0 (I y − I x )P 0 + I xz R 0 I xz Q 0<br />

3<br />

0 mg cos(Θ 0 ) 0<br />

0 3×3 −mg cos(Θ 0 ) cos(Φ 0 ) mg sin(Θ 0 ) sin(Φ 0 ) 0<br />

mg cos(Θ 0 ) sin(Φ 0 ) mg sin(Θ 0 ) cos(Φ 0 ) 0<br />

7<br />

0 3×3 0 3×3<br />

5<br />

In addition to this, the kinematic equations must be linearized.<br />

2.4.3 Linear state-space model based using wind <strong>and</strong> stability axes<br />

An alternative state-space model is obtained by using α <strong>and</strong> β as states. If it is assumed<br />

that α <strong>and</strong> β are small such that cos(α) ≈ 1 <strong>and</strong> sin(β) ≈ β, Equation (2.15) can be written<br />

as:<br />

U = V T U = V T<br />

V = V T β<br />

W = V T α<br />

Furthermore, the state-space vector:<br />

2 3 2<br />

u<br />

β<br />

x =<br />

α<br />

6 p<br />

=<br />

7 6<br />

4 q 5 4<br />

r<br />

⇒<br />

β = V V T<br />

α = W V T<br />

(2.43)<br />

surge velocity<br />

sideslip angle<br />

angle <strong>of</strong> attack<br />

roll rate<br />

pitch rate<br />

yaw rate<br />

3<br />

7<br />

5<br />

(2.44)<br />

is chosen to describe motions in 6 DOF. The relationship between the body-fixed velocity<br />

vector:<br />

ν = [u, v, w, p, q, r] T (2.45)<br />

<strong>and</strong> the new state-space vector x can be written as:<br />

ν = Tx = diag{1, V T , V T , 1, 1, 1, 1}x (2.46)


12 CHAPTER 2. AIRCRAFT MODELING<br />

where V T > 0. If the total speed is V T = U 0 = constant (linear theory), it is seen that:<br />

˙α = 1<br />

V T<br />

ẇ (2.47)<br />

˙β = 1<br />

V T<br />

˙v (2.48)<br />

˙V T = 0 (2.49)<br />

If nonlinear theory is applied, the following differential equations are obtained:<br />

˙α =<br />

UẆ − W ˙U<br />

U 2 + W 2 (2.50)<br />

˙β = ˙V V T − V ˙V T<br />

V 2<br />

T cos β (2.51)<br />

˙V T = U ˙U + V ˙V + W Ẇ<br />

V T<br />

(2.52)<br />

In the linear case it is possible to trans<strong>for</strong>m the body-fixed state-space model:<br />

to<br />

where<br />

˙ν = Fν + Gu (2.53)<br />

ẋ = Ax + Bu (2.54)<br />

A = T −1 FT, B = T −1 G (2.55)<br />

For ˙V T ≠ 0 this trans<strong>for</strong>mation is much more complicated. The linear state-space trans<strong>for</strong>mation<br />

is commonly used by aircraft manufactures. An example is the Boeing B-767 model<br />

(see Chapter 4).<br />

2.5 Decoupling in Longitudinal <strong>and</strong> Lateral Modes<br />

For an aircraft it is common to assume that the longitudinal modes (DOFs 1, 3 <strong>and</strong> 5)<br />

are decoupled from the lateral modes (DOFs 2, 4 <strong>and</strong> 6). The key assumption is that the<br />

fuselage is slender—that is, the length is much larger than the width <strong>and</strong> the height <strong>of</strong> the<br />

aircraft. It is also assumed that the longitudinal velocity is much larger than the vertical<br />

<strong>and</strong> transversal velocities.<br />

In order to decouple the rigid-body kinetics (2.41) in longitudinal <strong>and</strong> lateral modes it<br />

will be assumed that the states v, p, r <strong>and</strong> φ are negligible in the longitudinal channel while<br />

u, w, q <strong>and</strong> θ are negligible when considering the lateral channel. This gives two sub-systems:


2.5. DECOUPLING IN LONGITUDINAL AND LATERAL MODES 13<br />

2.5.1 Longitudinal equations<br />

Kinetics:<br />

2<br />

4<br />

m[ ˙u + Q 0 w + W 0 q + g cos(Θ 0 )θ] = δX<br />

m[ẇ − U 0 q − Q 0 u + g sin(Θ 0 ) cos(Φ 0 )θ] = δZ (2.56)<br />

3 2<br />

m 0 0<br />

0 m 0 5 4<br />

0 0 I y<br />

˙u<br />

ẇ<br />

˙q<br />

3<br />

2<br />

5 + 4<br />

0 mQ 0 mW 0<br />

−mQ 0 0 −mU 0<br />

0 0 0<br />

2<br />

4<br />

I y ˙q = δM<br />

mg cos(Θ 0 )<br />

mg sin(Θ 0 ) cos(Φ 0 )<br />

0<br />

3 2<br />

5 4<br />

3<br />

u<br />

w<br />

q<br />

3<br />

5 +<br />

2<br />

5 θ = 4<br />

δX<br />

δZ<br />

δM<br />

3<br />

5<br />

(2.57)<br />

Kinematics:<br />

˙θ = q (2.58)<br />

2.5.2 Lateral equations<br />

Kinetics:<br />

2<br />

4<br />

m 0 0<br />

3 2<br />

0 I x −I xz<br />

5 4<br />

m[ ˙v + U 0 r − W 0 p − g cos(Θ 0 ) cos(Φ 0 )φ] = δY<br />

I x ṗ − I xz ṙ + (I z − I y )Q 0 r − I xz Q 0 p = δL (2.59)<br />

I z ṙ − I xz ṗ + (I y − I x )Q 0 p + I xz Q 0 r = δN<br />

˙v<br />

ṗ<br />

ṙ<br />

3<br />

2<br />

5 + 4<br />

0 −mW 0 mU 0<br />

3<br />

0 −I xz Q 0 (I z − I y )Q 0<br />

5<br />

2<br />

4<br />

2<br />

4<br />

−mg cos(Θ 0 ) cos(Φ 0 )<br />

0<br />

0<br />

v<br />

p<br />

r<br />

3<br />

3<br />

5 +<br />

2<br />

5 φ = 4<br />

δY<br />

δL<br />

δN<br />

3<br />

5<br />

(2.60)<br />

Kinematics:<br />

˙φ<br />

˙ψ<br />

<br />

=<br />

1 tan(Θ0 )<br />

0 1/ cos(Θ 0 )<br />

p<br />

r<br />

<br />

(2.61)


14 CHAPTER 2. AIRCRAFT MODELING<br />

Figure 2.3: <strong>Control</strong> inputs <strong>for</strong> conventional aircraft. Notice that the two ailerons can be<br />

controlled by using one control input: δ A = 1/2(δ AL + δ AR ).<br />

2.6 Aerodynamic Forces <strong>and</strong> Moments<br />

In the <strong>for</strong>thcoming sections, the following abbreviations <strong>and</strong> notation will be used to describe<br />

the aerodynamic coeffi cients:<br />

X index =<br />

Y index =<br />

Z index =<br />

∂X L<br />

∂ index index = ∂L<br />

∂ index<br />

∂Y<br />

∂ index<br />

M index = ∂M<br />

∂ index<br />

∂Z<br />

∂ index<br />

N index = ∂N<br />

∂ index<br />

In order to illustrate how control surfaces influence the aircraft, an aircraft equipped with<br />

the following control inputs will be considered (see Figure 2.3):


2.6. AERODYNAMIC FORCES AND MOMENTS 15<br />

δ T Thrust Jet/propeller<br />

δ E Elevator<br />

<strong>Control</strong> surfaces on the rear <strong>of</strong> the aircraft used <strong>for</strong> pitch <strong>and</strong><br />

altitude control<br />

δ A Aileron<br />

Hinged control surfaces attached to the trailing edge <strong>of</strong> the wing used<br />

<strong>for</strong> roll/bank control<br />

δ F Flaps<br />

Hinged surfaces on the trailing edge <strong>of</strong> the wings used <strong>for</strong> braking<br />

<strong>and</strong> bank-to-turn<br />

δ R Rudder Vertical control surface at the rear <strong>of</strong> the aircraft used <strong>for</strong> turning<br />

A conventional aircraft has control surfaces such as ailerons δ A , elevators δ E , flaps δ F <strong>and</strong><br />

rudders δ R . A positive deflection <strong>of</strong> the control surfaces are to give a negative aerodynamic<br />

moment on the aircraft. Some <strong>of</strong> the control surfaces, like the ailerons, deflects simultaneously<br />

in an asymmetric manner. If an aircraft has two ailerons δ AL <strong>and</strong> δ AR , the aileron<br />

deflection is computed by:<br />

δ A = 1 2 (δ A R<br />

− δ AL ) (2.62)<br />

Other control surfaces, such as elevators, deflect symmetrically. If an aircraft has two elevators<br />

δ EL <strong>and</strong> δ ER , the elevator deflections become:<br />

δ E = 1 2 (δ E R<br />

+ δ EL ) (2.63)<br />

Linear theory<br />

The number <strong>of</strong> aerodynamic coeffi cients is reduced if linear theory is assumed. The control<br />

inputs <strong>and</strong> aerodynamic <strong>for</strong>ces <strong>and</strong> moments can be written as:<br />

τ = −M F ˙ν − N F ν + Bu (2.64)<br />

where −M F is aerodynamic added mass, −N F is aerodynamic damping <strong>and</strong> B is a matrix<br />

describing the actuator configuration including the <strong>for</strong>ce coeffi cients. The actuator dynamics<br />

is modeled by a 1st-order system:<br />

˙u = T −1 (u c − u) (2.65)<br />

where u c is comm<strong>and</strong>ed input, u is the actual control input produced by the actuators <strong>and</strong><br />

T = diag{T 1 , T 2 , ..., T r } is a diagonal matrix <strong>of</strong> positive time constants. Substitution <strong>of</strong> (2.64)<br />

into the model (2.42) gives:<br />

(M RB + M F ) ˙ν + (N RB<br />

+ N F<br />

)ν + Gη = Bu<br />

M ˙ν + Nν + Gη = Bu<br />

⇕ (2.66)<br />

The matrices M <strong>and</strong> N are defined as M = M RB + M F <strong>and</strong> N = N RB + N F<br />

. The linearized<br />

kinematics takes the following <strong>for</strong>m:<br />

The resulting state-space models are:<br />

˙η = J ν ν + J η η (2.67)


16 CHAPTER 2. AIRCRAFT MODELING<br />

Linear state-space model with actuator dynamics<br />

2<br />

4<br />

˙η<br />

˙ν<br />

˙u<br />

3<br />

2<br />

5 = 4<br />

3<br />

J η J ν 0<br />

−M −1 G −M −1 N M −1 B 5<br />

0 0 −T −1<br />

2<br />

4<br />

η<br />

ν<br />

u<br />

3<br />

2<br />

5 + 4<br />

3<br />

0<br />

0 5 u c (2.68)<br />

T −1<br />

Linear state-space model neglecting the actuator dynamics<br />

<br />

J<br />

= η<br />

˙η˙ν −M −1 G<br />

J ν<br />

−M −1 N<br />

η<br />

ν<br />

<br />

+<br />

0<br />

M −1 B<br />

<br />

u (2.69)<br />

2.6.1 Longitudinal aerodynamic <strong>for</strong>ces <strong>and</strong> moments<br />

McLean [7] expresses the longitudinal <strong>for</strong>ces <strong>and</strong> moments as:<br />

2<br />

4<br />

dX<br />

dZ<br />

dM<br />

3<br />

2<br />

5 = 4<br />

3<br />

X ˙u Xẇ X ˙q<br />

Z ˙u Zẇ Z ˙q<br />

5<br />

M ˙u Mẇ M ˙q<br />

2<br />

4<br />

˙u<br />

ẇ<br />

˙q<br />

3<br />

5 + 4<br />

2<br />

4<br />

2<br />

3<br />

X u X w X q<br />

Z u Z w Z q<br />

5<br />

M q M w M q<br />

3 2<br />

X δT X δE X δF<br />

Z δT<br />

Z δE Z δF<br />

5 4<br />

M δT M δE M δF<br />

2<br />

4<br />

u<br />

w<br />

q<br />

3<br />

δ T<br />

3<br />

δ E<br />

5<br />

5 +<br />

(2.70)<br />

which corresponds to the matrices −M F , −N F <strong>and</strong> B in (2.64). If the aircraft cruise speed<br />

U 0 = constant, then δ T = 0. Altitude can be controlled by using the elevators δ E . Flaps δ F<br />

can be used to reduce the speed during l<strong>and</strong>ing. The flaps can also be used to turn harder<br />

<strong>for</strong> instance by moving one flap while the other is kept at the zero position. This is common<br />

in bank-to-turn maneuvers. For conventional aircraft the following aerodynamic coeffi cients<br />

can be neglected:<br />

X ˙u , X q , Xẇ, X δE , Z ˙u , Zẇ, M ˙u (2.71)<br />

Hence, the model <strong>for</strong> altitude control reduces to:<br />

2<br />

4<br />

dX<br />

dZ<br />

dM<br />

3<br />

2<br />

5 = 4<br />

3<br />

0 0 X ˙q<br />

0 0 Z ˙q<br />

5<br />

0 Mẇ M ˙q<br />

2<br />

4<br />

˙u<br />

ẇ<br />

˙q<br />

3<br />

2<br />

5 + 4<br />

3<br />

X u X w 0<br />

Z u Z w Z q<br />

5<br />

M q M w M q<br />

2<br />

4<br />

u<br />

w<br />

q<br />

3<br />

2<br />

5 + 4<br />

3<br />

X δE<br />

Z δE<br />

5 δ E (2.72)<br />

M δE<br />

If the actuator dynamics is important, aerodynamic coeffi cients such as X˙δT<br />

, X˙δE<br />

, ... must<br />

be included in the model.<br />

2.6.2 Lateral aerodynamic <strong>for</strong>ces <strong>and</strong> moments<br />

The lateral model takes the <strong>for</strong>m [7]:


2.7. PROPELLER THRUST 17<br />

2<br />

4<br />

dY<br />

dL<br />

dN<br />

3<br />

2<br />

5 = 4<br />

Y ˙v Yṗ Yṙ<br />

L ˙v Lṗ Lṙ<br />

N ˙v Nṗ Nṙ<br />

3 2<br />

5 4<br />

˙v<br />

ṗ<br />

ṙ<br />

3<br />

2<br />

5 + 4<br />

3<br />

Y v Y p Y r<br />

L v L p L r<br />

5<br />

N v N p N r<br />

2<br />

4<br />

v<br />

p<br />

r<br />

3<br />

2<br />

5 + 4<br />

Y δA Y δR<br />

3<br />

L δA L δR<br />

<br />

5 δA<br />

δ R<br />

(2.73)<br />

which corresponds to the matrices −M F , −N F <strong>and</strong> B in (2.64). For conventional aircraft<br />

the following aerodynamic coeffi cients can be neglected:<br />

This gives:<br />

Y ˙v , Y p , Yṗ, Y r , Yṙ, Y δA ,L ˙v , Lṙ, N ˙v , Nṙ (2.74)<br />

2<br />

4<br />

dY<br />

dL<br />

dN<br />

3<br />

2<br />

5 = 4<br />

0 0 0<br />

0 Lṗ 0<br />

0 Nṗ 0<br />

3 2<br />

5 4<br />

˙v<br />

ṗ<br />

ṙ<br />

3<br />

2<br />

5 + 4<br />

3<br />

Y v 0 0<br />

L v L p L r<br />

5<br />

N v N p N r<br />

2<br />

4<br />

v<br />

p<br />

r<br />

3<br />

2<br />

5 + 4<br />

3<br />

0 Y δR<br />

<br />

L δA<br />

L δR<br />

5 δA<br />

δ<br />

N δA N R<br />

δR<br />

(2.75)<br />

2.7 Propeller Thrust<br />

The thrust from a propeller can be modeled as:<br />

where<br />

T Thrust <strong>for</strong>ce [N]<br />

K t Thrust coeffi cient<br />

ρ Density <strong>of</strong> air [kg/m 3 ]<br />

n Propeller angular velocity [rad/s]<br />

D Propeller diameter [m]<br />

T = K t ρn 2 D 4 (2.76)<br />

Assuming that the trust <strong>for</strong>ce acts in the horizontal plane <strong>of</strong> {b}, <strong>and</strong> parallel with the<br />

x b -axis, it can be expressed in {b} as:<br />

2 3 2 3<br />

X t T<br />

ft b = 4 Y t<br />

5 = 4 0 5 (2.77)<br />

Z t 0<br />

3<br />

3<br />

If the line <strong>of</strong> action <strong>of</strong> the thrust is not directed through CO, it will in addition generate an<br />

applied moment given by<br />

2<br />

L t<br />

N t<br />

2<br />

0<br />

m b t = 4 M t<br />

5 = r b t/b × ft b = S(r b t/b)ft<br />

b<br />

−T r ty<br />

= 4 T r tz<br />

5 (2.78)


18 CHAPTER 2. AIRCRAFT MODELING<br />

where r b t/b = [r t x<br />

, r ty , r tz ] ⊤ is the location <strong>of</strong> the propeller with respect to CO.<br />

Sometimes the effects due to the rotating mass caused by the propeller must be considered<br />

as an applied moment. This is because the equations <strong>of</strong> motion has been derived by assuming<br />

that the aircraft is a rigid body with no internal moving parts. The applied moment due to<br />

the rotating mass caused by the propeller is given by:<br />

m b p =<br />

i d<br />

dt hb p (2.79)<br />

where h b p is the angular momentum <strong>of</strong> the rotating mass, <strong>and</strong> is given in {b} as:<br />

2<br />

h b p = 4<br />

I p Ω p<br />

0<br />

0<br />

3<br />

5 (2.80)<br />

where I p is the inertia <strong>of</strong> the rotating mass <strong>and</strong> Ω p is the angular velocity. Assuming the<br />

angular velocity <strong>of</strong> the rotating mass is constant:<br />

2<br />

m b p = 4<br />

L p<br />

3<br />

M p<br />

5 =<br />

i d<br />

dt hb p =<br />

b d<br />

dt hb p + ω b b/n × h b p<br />

= ω b b/n × h b p = S(ω b b/n)h b p<br />

2 3<br />

0<br />

= 4 I p Ω p R<br />

−I p Ω p Q<br />

5 (2.81)<br />

The total <strong>for</strong>ces <strong>and</strong> moments due to thrust, τ t are:<br />

<br />

τ t =<br />

f b t<br />

S(r b t/g )f b t + S(ω b b/n )hb p<br />

<br />

(2.82)<br />

2.8 Aerodynamic Coeffi cients<br />

Different methods can be used to estimate the aerodynamic coeffi cients such as wind tunnel<br />

experiments or system identification based on logged experimental data from the aircraft.<br />

The aircraft equations <strong>of</strong> motion can be quite simple if the aircraft is not highly maneuverable.<br />

The more maneuverable aircraft, the more coeffi cients are needed to accurately<br />

describe the aerodynamic <strong>for</strong>ces <strong>and</strong> moments.<br />

The aerodynamic coeffi cients are usually linear at small angles <strong>of</strong> angle <strong>of</strong> attack <strong>and</strong><br />

sideslip. The most important parameters used to describe an airfoil are:<br />

b<br />

c<br />

¯c<br />

S<br />

AR = b 2 /S<br />

Wing span (tip to tip)<br />

Wing chord (varies along span)<br />

Mean aerodynamic chord<br />

Total wing area<br />

Aspect ratio


2.8. AERODYNAMIC COEFFICIENTS 19<br />

<strong>Aircraft</strong> with delta-shaped wings <strong>of</strong>ten have a low-aspect ratio, AR. A low-aspect ratio<br />

permits very high roll rates, but have a greatly reduced lift-over-drag. High lift-over-drag<br />

gives an aircraft that has effective cruise per<strong>for</strong>mance, that is passenger jets.<br />

The aerodynamic <strong>for</strong>ces <strong>and</strong> moments are <strong>of</strong>ten proportional to the freestream mass<br />

density, ρ a , <strong>and</strong> the square <strong>of</strong> the freestream airspeed, V T . It is thus convenient to define the<br />

dynamic pressure, ¯q, which is later used to calculate the aerodynamic <strong>for</strong>ces <strong>and</strong> moments:<br />

2.8.1 Aerodynamic <strong>for</strong>ces <strong>and</strong> moments<br />

¯q = 1 2 ρ aV 2<br />

T (2.83)<br />

The aerodynamic <strong>for</strong>ces <strong>and</strong> moments acting on an aircraft can be parametrized as:<br />

2 3 2 3<br />

X a C X<br />

fa b = 4 Y a<br />

5 = ¯qS 4 C Y<br />

5 (2.84)<br />

Z a C z<br />

2<br />

m b a = 4<br />

L a<br />

3<br />

M a<br />

5 = ¯qS<br />

2<br />

4<br />

bC l<br />

3<br />

¯cC m<br />

5 (2.85)<br />

where the aerodynamic coeffi cients C X <strong>and</strong> C Z <strong>of</strong>ten are replaced by expressions that are<br />

functions <strong>of</strong> the drag coeffi cient C D <strong>and</strong> the lift coeffi cient C L (see Figure 2.4). Positive drag<br />

<strong>and</strong> lift coeffi cients are directed along the −x s <strong>and</strong> −z s stability axes, respectively, such that:<br />

2<br />

4<br />

C D<br />

3<br />

0<br />

C<br />

5 = R s b<br />

2<br />

= 4<br />

2<br />

4<br />

3<br />

−C X<br />

0 5<br />

−C Z<br />

cos(α) 0 sin(α)<br />

0 1 0<br />

− sin(α) 0 cos(α)<br />

3 2<br />

5 4<br />

−C X<br />

3<br />

0 5 (2.86)<br />

Hence, the aerodynamic <strong>for</strong>ces can then be rewritten as:<br />

2 3 2 3 2 3<br />

X a<br />

0<br />

−C D<br />

fa b = 4 Y a<br />

5 = ¯qS 4 C Y<br />

5 + ¯qSR b 4<br />

s 0 5<br />

Z a 0<br />

−C L<br />

2<br />

3<br />

−C D cos(α) + C L sin(α)<br />

= ¯qS 4 C Y<br />

−C D sin(α) − C L cos(α)<br />

5 (2.87)<br />

The generalized aerodynamic <strong>for</strong>ces become:<br />

f<br />

b<br />

τ a = a<br />

m b a<br />

<br />

(2.88)<br />

The aerodynamic coeffi cients C i , i = D, Y, L, l, m, n, are in general functions <strong>of</strong> angle <strong>of</strong><br />

attack α, sideslip angle β, Mach number M, altitude h, control surface deflections δ S <strong>and</strong><br />

the thrust coeffi cient:<br />

T C =<br />

T<br />

(2.89)<br />

¯qS D


20 CHAPTER 2. AIRCRAFT MODELING<br />

where S D is the area <strong>of</strong> the disc swept out by a propeller blade. There<strong>for</strong>e, the aerodynamic<br />

coeffi cients are functions:<br />

C i = C i (α, β, M, h, δ S , T C ) (2.90)<br />

where i = D, Y, L, l, m, n. Consequently, the aerodynamic coeffi cients may have complicated<br />

dependences, which are hard to model with great accuracy. If the aircraft has restricted<br />

flight modes, thts is restricted to low Mach numbers <strong>and</strong> small angles <strong>of</strong> attack <strong>and</strong> sideslip,<br />

the aerodynamic coeffi cients may be greatly simplified. The complicated dependency may<br />

then be broken into a sum <strong>of</strong> simpler linear terms.<br />

2.8.2 Drag coeffi cient<br />

The drag <strong>of</strong> an aircraft is one <strong>of</strong> the most important features to model. By reducing drag,<br />

the fuel consumption is reduced, thus giving higher range, <strong>and</strong> in addition higher maximum<br />

cruise speed. The total drag <strong>of</strong> an aircraft is usually taken as the sum <strong>of</strong> the following drag<br />

components:<br />

• Parasite drag = friction drag + <strong>for</strong>m drag<br />

• Induced drag (effect <strong>of</strong> wing-tip vortices)<br />

• Wave drag (effect <strong>of</strong> shock waves)<br />

where<br />

Friction drag also known as skin friction, describes the friction <strong>of</strong> the fluid against the<br />

surface <strong>of</strong> the aircraft. The friction drag is proportional to the area in contact with<br />

the fluid, also known as the wetted area.<br />

Form drag is the pressure drag caused by flow separation at high alpha.<br />

Induced Drag also known as vortex drag, is the pressure drag caused by the tip vortices<br />

<strong>of</strong> a finite wing when it is producing lift.<br />

Wave drag is the pressure drag caused if shock waves are present over the surface <strong>of</strong> the<br />

aircraft. The wave drag becomes significant at Mach numbers M > 0.4 <strong>for</strong> a fighter<br />

aircraft, where the flow reaches supersonic speed.<br />

The friction drag <strong>of</strong>ten varies parabolically with the lift coeffi cient, while the induced<br />

drag can <strong>of</strong>ten be modeled as:<br />

C Di =<br />

C2 L<br />

(2.91)<br />

πeAR<br />

where e is the effi ciency factor, close to unity, <strong>and</strong> AR is the aspect ratio. The total drag <strong>of</strong><br />

an aircraft takes the <strong>for</strong>m:<br />

C D (C L , M) = k(M) (C L − C LDM ) 2 + C DM (M) (2.92)


2.8. AERODYNAMIC COEFFICIENTS 21<br />

where k(M) is a proportionality constant that varies with the Mach number. In addition,<br />

the drag coeffi cient may be a function <strong>of</strong> the control surfaces <strong>and</strong> l<strong>and</strong>ing gear. A more<br />

general equation <strong>for</strong> the drag coeffi cient is:<br />

C D = C D (α, β, M, h) + ∆C D (M, δ E ) + ∆C D (M, δ R ) + ∆C D (δ F ) + ∆C D (gear) + · · · (2.93)<br />

A linear approximation <strong>of</strong> this expressions is:<br />

2.8.3 Lift coeffi cient<br />

C D = C D0 + C Dα α + C DδE δ E + C DδR δ R + C DδF δ F (2.94)<br />

The lift coeffi cient, C L , is in general mainly determined by the fuselage, wings <strong>and</strong> their<br />

interference effects. It is also a function <strong>of</strong> Mach <strong>and</strong> the angle <strong>of</strong> attack. The lift coeffi cient<br />

is usually quite linear with a until near stall, where it drops sharply, <strong>and</strong> then may rise again<br />

be<strong>for</strong>e falling to zero at large angles <strong>of</strong> attack (see Figure 2.4). A general equation <strong>for</strong> the<br />

lift coeffi cient is given in<br />

C L = C L (α, β, M, T C ) + ∆C L (δ F ) + ∆C Lge (h) (2.95)<br />

where ∆C Lge (h) is the increment <strong>of</strong> lift due to ground effect. A typical simple linear model<br />

structure <strong>for</strong> the lift coeffi cient is given in<br />

C L = C L0 + C Lα α + C LδF δ F (2.96)<br />

2.8.4 Side<strong>for</strong>ce coeffi cient<br />

The side<strong>for</strong>ce coeffi cient, C Y , is in general mainly determined by sideslip <strong>and</strong> rudder deflection.<br />

It is <strong>of</strong>ten quite linear in sideslip angle β, <strong>and</strong> a positive sideslip will give a negative<br />

side<strong>for</strong>ce. A typical expression <strong>for</strong> the side<strong>for</strong>ce coeffi cient is:<br />

C Y = C Y (α, β, M)+∆C YδR (α, β, M, δ R )+∆C YδA (α, β, M, δ A )+ b<br />

2V T<br />

<br />

CYp (α, M)P + C Yr (α, M)R <br />

A typical simple linear model structure <strong>for</strong> the side<strong>for</strong>ce coeffi cient is given in<br />

2.8.5 Rolling moment coeffi cient<br />

(2.97)<br />

C Y = C Y0 + C Yβ β + C YδR δ R + C YδA δ A (2.98)<br />

The rolling moment coeffi cient, C l , is in general mainly determined by sideslip <strong>and</strong> by control<br />

action <strong>of</strong> the ailerons <strong>and</strong> the rudder. It is <strong>of</strong>ten quite linear in sideslip angle <strong>and</strong> a positive<br />

sideslip will give a negative rolling moment. A typical expression <strong>for</strong> the rolling moment<br />

coeffi cient is:<br />

C l = C l (α, β, M)+∆C lδA (α, β, M, δ A )+∆C lδR (α, β, M, δ R )+ b <br />

Clp (α, M)P + C lr (α, M)R <br />

2V T<br />

(2.99)


22 CHAPTER 2. AIRCRAFT MODELING<br />

Figure 2.4: Lift <strong>and</strong> drag as a function <strong>of</strong> angle <strong>of</strong> attack.<br />

A linear model <strong>for</strong> the rolling moment coeffi cient is:<br />

b b<br />

C l = C l0 + C lβ β + C lδA δ A + C lδR δ R + C lp P + C lr R (2.100)<br />

2V T 2V T<br />

2.8.6 Pitching moment coeffi cient<br />

The pitching moment coeffi cient, C m , is in general mainly determined by angle <strong>of</strong> attack<br />

α <strong>and</strong> by control action <strong>of</strong> the elevators. It may be quite linear in angle <strong>of</strong> attack, <strong>and</strong><br />

a positive angle <strong>of</strong> attack will give a negative pitching moment. A typical model <strong>for</strong> the<br />

pitching moment coeffi cient is:<br />

C m = C m (α, M, h, δ F , T C ) + ∆C mδe (α, M, h, δ E )<br />

+ ¯c C mq (α, M, h)Q + C m<br />

2V<br />

˙α<br />

(α, M, h) ˙α + x R<br />

T<br />

¯c C L<br />

+ ∆C mthrust (δ T , M, h) + ∆C mgear (h) (2.101)<br />

The term x C<br />

R¯c L is used to correct <strong>for</strong> any x-displacement <strong>of</strong> the aircraft CG from the aerodynamic<br />

data reference position. The second last term represents the effect <strong>of</strong> the engine thrust<br />

vector not passing through the aircraft CG, while the last term represents the moment due<br />

to the l<strong>and</strong>ing-gear doors <strong>and</strong> the l<strong>and</strong>ing gear. A linear model <strong>for</strong> the pitching moment<br />

coeffi cient takes the <strong>for</strong>m:<br />

¯c<br />

C m = C m0 + C mα α + C mδe δ e + C mq Q (2.102)<br />

2V T


2.9. STANDARD AIRCRAFT MANEUVERS 23<br />

2.8.7 Yawing moment coeffi cient<br />

The yawing moment coeffi cient, C n , is in general mainly determined by the sideslip angle β<br />

<strong>and</strong> by control action <strong>of</strong> the rudders. It is quite linear in sideslip β <strong>for</strong> small angles <strong>of</strong> attack<br />

α, where a positive sideslip β will give a positive yawing moment. A general equation <strong>for</strong><br />

the yawing moment coeffi cient is given by:<br />

C n = C n (α, β, M, T C ) + ∆C nδR (α, β, M, δ R ) + ∆C nδA (α, β, M, δ A )<br />

+ b<br />

2V T<br />

<br />

Cnp (α, M)P + C nr (α, M)R (2.103)<br />

A linear model <strong>for</strong> the yawing moment coeffi cient is gven by:<br />

b b<br />

C n = C n0 + C nβ β + C nδR δ R + C nδA δ A + C np P + C nr R (2.104)<br />

2V T 2V T<br />

2.8.8 Equations <strong>of</strong> motion including aerodynamic coeffi cients<br />

The resulting equations <strong>of</strong> motion including the aerodynamic coeffi cients are:<br />

P˙<br />

− I xz<br />

I x<br />

Ṙ − I xz<br />

I z<br />

˙U = V R − W Q − g sin(Θ) + T m + ¯qS m C X (2.105)<br />

˙V = W P − UR + g cos(Θ) sin(Φ) + ¯qS m C Y (2.106)<br />

Ẇ = UQ − V P + g cos(Θ) cos(Φ) + ¯qS m C Z (2.107)<br />

Ṙ = − I z − I y<br />

QR + I xz<br />

P Q + ¯qbS C l (2.108)<br />

I x I x I x<br />

˙Q = − I x − I z<br />

P R − I xz<br />

(P 2 − R 2 ) + T r t z<br />

I y I y I y<br />

P ˙ = − I y − I x<br />

I z<br />

P Q − I xz<br />

I z<br />

QR − T r t y<br />

I z<br />

2.9 St<strong>and</strong>ard <strong>Aircraft</strong> Maneuvers<br />

The nominal values depends on the aircraft maneuver. For instance:<br />

1. Straight flight: Θ 0 = Q 0 = R 0 = 0<br />

2. Symmetric flight: Ψ 0 = V 0 = 0<br />

3. Flying with wings level: Φ 0 = P 0 = 0<br />

Constant angular rate maneuvers can be classified according to:<br />

1. Steady turn: R 0 = constant<br />

2. Steady pitching flight: Q 0 = constant<br />

3. Steady rolling/spinning flight: P 0 = constant<br />

+ I p<br />

I y<br />

Ω p R + ¯qS¯c<br />

I y<br />

C m (2.109)<br />

− I p<br />

I z<br />

Ω p Q + ¯qSb<br />

I z<br />

C n (2.110)


24 CHAPTER 2. AIRCRAFT MODELING<br />

2.9.1 Dynamic equation <strong>for</strong> coordinated turn (bank-to-turn)<br />

A frequently used maneuver is coordinated turn where the acceleration in the y-direction is<br />

zero ( ˙β = 0), sideslip β = 0 <strong>and</strong> zero steady-state pitch <strong>and</strong> roll angles—that is,<br />

β = ˙β = 0 (2.111)<br />

Θ 0 = Φ 0 = 0 (2.112)<br />

Furthermore it is assumed that V T = U 0 = constant. Since β = ˙β = 0 <strong>and</strong> β = V/U 0 it<br />

follows that V = ˙V = 0. This implies that the external <strong>for</strong>ces δY = 0. From (2.28) it is seen<br />

that:<br />

m[ ˙V + UR − W P − g cos(Θ) sin(Φ)] = Y (2.113)<br />

m[(U 0 + u)(R 0 + r) − (W 0 + w)(P 0 + p) − g cos(θ) sin(φ)] = Y 0 (2.114)<br />

Assume that the longitudinal <strong>and</strong> lateral motions are decoupled—that is, u = w = q = θ = 0.<br />

If perturbation theory is applied under the assumption that the 2nd-order terms ur = pw =<br />

0,we get:<br />

m(U 0 R 0 + U 0 r − W 0 P 0 − W 0 p − g sin(φ)) = Y 0 (2.115)<br />

The equilibrium equation (2.35) gives the steady-state condition:<br />

Substitution <strong>of</strong> (2.116) into (2.115) gives:<br />

or<br />

m(U 0 R 0 − W 0 P 0 ) = Y 0 (2.116)<br />

m(U 0 r − W 0 p − g sin(φ)) = 0 (2.117)<br />

r = W 0<br />

U 0<br />

p + g U 0<br />

sin(φ) (2.118)<br />

The aircraft is <strong>of</strong>ten trimmed such that the angle <strong>of</strong> attack α 0 = W 0 /U 0 = 0. This implies<br />

that the yaw rate can be expressed as:<br />

r = g U 0<br />

sin(φ) φ small<br />

≈ g U 0<br />

φ (2.119)<br />

which is a very important result since it states that a roll angle angle φ different from zero<br />

will induce a yaw rate r which again turns the aircraft (bank-to-turn). With other words, we<br />

can use a moment in roll, <strong>for</strong> instance generated by the ailerons, to turn the aircraft. The<br />

yaw angle is given by:<br />

˙ψ = r (2.120)<br />

An alternative method is <strong>of</strong> course to turn the aircraft by using the rear rudder to generate<br />

a yaw moment. The bank-to-turn principle is used in many missile control systems since it<br />

improves maneuverability, in particular in combination with a rudder controlled system.<br />


2.9. STANDARD AIRCRAFT MANEUVERS 25<br />

Example 2 (Augmented turning model using rear rudders)<br />

Turning autopilots using the rudder δ R as control input is based on the lateral state-space<br />

model. The differential equation <strong>for</strong> ψ is augmented on the lateral model as shown below:<br />

□<br />

2<br />

6<br />

4<br />

˙v<br />

ṗ<br />

ṙ<br />

˙φ<br />

˙ψ<br />

3 2<br />

7<br />

5 = 6<br />

4<br />

a 11 a 12 a 13 a 14 0<br />

a 21 a 22 a 23 0 0<br />

a 31 a 32 a 33 0 0<br />

0 1 0 0 0<br />

0 0 1 0 0<br />

3 2<br />

7 6<br />

5 4<br />

v<br />

p<br />

r<br />

φ<br />

ψ<br />

3 2<br />

7<br />

5 + 6<br />

4<br />

b 11<br />

b 21<br />

b 31<br />

0<br />

0<br />

3<br />

7<br />

5 δ R (2.121)<br />

Example 3 (Augmented bank-to-turn model using ailerons)<br />

Bank-to-turn autopilots are designed using the lateral state-space model with ailerons as<br />

control inputs, <strong>for</strong> instance δ A = 1/2(δ AL + δ AR ). This is done by augmenting the bank-toturn<br />

equation (2.119) to the state-space model according to:<br />

□<br />

2<br />

6<br />

4<br />

˙v<br />

ṗ<br />

ṙ<br />

˙φ<br />

˙ψ<br />

3 2<br />

7<br />

5 = 6<br />

4<br />

a 11 a 12 a 13 a 14 0<br />

a 21 a 22 a 23 0 0<br />

a 31 a 32 a 33 0 0<br />

0 1 0 0 0<br />

0 0 0<br />

g<br />

U 0<br />

0<br />

3 2<br />

7 6<br />

5 4<br />

v<br />

p<br />

r<br />

φ<br />

ψ<br />

3 2<br />

7<br />

5 + 6<br />

4<br />

b 11<br />

b 21<br />

b 31<br />

0<br />

0<br />

3<br />

7<br />

5 δ A (2.122)<br />

2.9.2 Dynamic equation <strong>for</strong> altitude control<br />

<strong>Aircraft</strong> altitude control systems are designed by considering the equation <strong>for</strong> the vertical<br />

acceleration in the center <strong>of</strong> gravity expressed in NED coordinates; see (2.28). This gives:<br />

a zCG = Ẇ + V P − QU − g cos(Θ) cos(Φ) (2.123)<br />

If the acceleration is perturbed according to a zCG = a z0 + δa z <strong>and</strong> we assume that Ẇ0 = 0,<br />

the following equilibrium condition is obtained:<br />

Equation (2.123) can be perturbed as:<br />

a z0 = V 0 P 0 − Q 0 U 0 − g cos(Θ 0 ) cos(Φ 0 ) (2.124)<br />

a z0 + δa z = ẇ + (V 0 + v)(P 0 + p) − (Q 0 + q)(U 0 + u) − g cos(Θ 0 + θ) cos(Φ 0 + φ) (2.125)<br />

Furthermore, assume that the altitude is changed by symmetric straight-line flight with<br />

horizontal wings such that V 0 = Φ 0 = Θ 0 = P 0 = Q 0 = 0. This gives:<br />

a z0 + δa z = ẇ + vp − q(U 0 + u) − g cos(θ) cos(φ) (2.126)


26 CHAPTER 2. AIRCRAFT MODELING<br />

Assume that the 2nd-order terms vp <strong>and</strong> uq can be neglected <strong>and</strong> subtract the equilibrium<br />

condition (2.124) from (2.126) such that:<br />

δa z = ẇ − U 0 q (2.127)<br />

Differentiating the altitude twice with respect to time gives the relationship:<br />

ḧ = −δa z = U 0 q − ẇ (2.128)<br />

If we integrate this expression under the assumption that ḣ(0) = U 0θ(0) − w(0) = 0, we get:<br />

ḣ = U 0 θ − w (2.129)<br />

The flight path is defined as:<br />

γ := θ − α (2.130)<br />

where α = w/U 0 . This gives the resulting differential equation <strong>for</strong> altitude control:<br />

ḣ = U 0 γ (2.131)<br />

Example 4 (Augmented model <strong>for</strong> altitude control using ailerons)<br />

An autopilot model <strong>for</strong> altitude control based on the longitudinal state-space model with<br />

states u, w(alt. α), q <strong>and</strong> θ is obtained by augmenting the differential equation <strong>for</strong> h to the<br />

state-space model according to:<br />

2<br />

6<br />

4<br />

˙u<br />

ẇ<br />

˙q<br />

˙θ<br />

ḣ<br />

3 2<br />

7<br />

5 + 6<br />

4<br />

a 11 a 12 a 13 a 14 0<br />

a 21 a 22 a 23 a 24 0<br />

a 31 a 32 a 33 0 0<br />

0 0 1 0 0<br />

0 −1 0 U 0 0<br />

3 2<br />

7 6<br />

5 4<br />

u<br />

w<br />

q<br />

θ<br />

h<br />

3 2<br />

7<br />

5 + 6<br />

4<br />

b 11 b 12<br />

b 21 b 22<br />

b 31 b 32<br />

0 0<br />

0 0<br />

3<br />

7<br />

5<br />

<br />

δT<br />

δ E<br />

<br />

(2.132)<br />

□<br />

2.10 <strong>Aircraft</strong> Stability Properties<br />

The aircraft stability properties can be investigated by computing the eigenvalues <strong>of</strong> the<br />

system matrix A using:<br />

For a matrix A <strong>of</strong> dimension n × n there is n solutions <strong>of</strong> λ.<br />

det(λI − A) = 0 (2.133)


2.10. AIRCRAFT STABILITY PROPERTIES 27<br />

2.10.1 Longitudinal stability analysis<br />

For conventional aircraft the characteristic equation is <strong>of</strong> 4th order. Moreover,<br />

(λ 2 + 2ζ ph ω ph λ + ω 2 ph)(λ 2 + 2ζ sp ω sp λ + ω 2 sp) = 0 (2.134)<br />

where the subscripts ph <strong>and</strong> sp denote the following modi:<br />

• Phugoid mode<br />

• Short-period mode<br />

The phugoid mode is observed as a long period oscillation with little damping. In some<br />

cases the Phugoid mode can be unstable such that the oscillations increase with time. The<br />

Phugoid mode is characterized by the natural frequency ω ph <strong>and</strong> relative damping ratio ζ ph .<br />

The short-period mode is a fast mode given by the natural frequency ω sp <strong>and</strong> relative<br />

damping factor ζ sp . The short-period mode is usually well damped.<br />

Classification <strong>of</strong> eigenvalues<br />

• Conventional aircraft: Conventional aircraft have usually two complex conjugated<br />

pairs <strong>of</strong> Phugoid <strong>and</strong> short-period types. For the B-767 model in Chapter 4, the<br />

eigenvalues can be computed using damp.m in Matlab:<br />

a = [<br />

-0.0168 0.1121 0.0003 -0.5608<br />

-0.0164 -0.7771 0.9945 0.0015<br />

-0.0417 -3.6595 -0.9544 0<br />

0 0 1.0000 0]<br />

damp(a)<br />

Eigenvalue Damping Freq. (rad/sec)<br />

-0.0064 + 0.0593i 0.1070 0.0596<br />

-0.0064 - 0.0593i 0.1070 0.0596<br />

-0.8678 + 1.9061i 0.4143 2.0943<br />

-0.8678 - 1.9061i 0.4143 2.0943<br />

From this is seen that ω ph = 0.0596, ζ ph = 0.1070, ω sp = 2.0943 <strong>and</strong> ζ sp = 0.4143.<br />

• Tuck Mode: Supersonic aircraft may have a very large aerodynamic coeffi cient M u .<br />

This implies that the oscillatory Phugoid equation gives two real solutions where one<br />

is positive (unstable) <strong>and</strong> one is negative (stable). This is referred to as the tuck mode<br />

since the phenomenon is observed as a downwards pointing nose (tucking under) <strong>for</strong><br />

increasing speed.


28 CHAPTER 2. AIRCRAFT MODELING<br />

Figure 2.5: <strong>Aircraft</strong> longitudinal eigenvalue configuration plotted in the complex plane.<br />

• A third oscillatory mode: For fighter aircraft the center <strong>of</strong> gravity is <strong>of</strong>ten located<br />

behind the neutral point or the aerodynamic center—that is, the point where the trim<br />

moment M w w is zero. When this happens, the aerodynamic coeffi cient M w takes a<br />

value such that the roots <strong>of</strong> the characteristic equation has four real solutions. When<br />

the center <strong>of</strong> gravity is moved backwards one <strong>of</strong> the roots <strong>of</strong> the Phugoid <strong>and</strong> shortperiod<br />

modes become imaginary <strong>and</strong> they <strong>for</strong>m a new complex conjugated pair. This<br />

is usually referred to as the 3rd oscillatory mode. The locations <strong>of</strong> the eigenvalues are<br />

illustrated in Figure 2.5.<br />

2.10.2 Lateral stability analysis<br />

The lateral characteristic equation is usually <strong>of</strong> 5th order:<br />

λ 5 + α 4 λ 4 + α 3 λ 3 + α 2 λ 2 + α 1 λ + α 0 = 0 (2.135)<br />

λ(λ + e)(λ + f)(λ 2 + 2ζ D ω D λ + ω 2 D) = 0 (2.136)<br />

where the term λ in the last equation corresponds to a pure integrator in roll—that is, ˙ψ = r.<br />

The term (λ+e) is the aircraft spiral/divergence mode. This is usually a very slow mode.<br />

Spiral/divergence corresponds to horizontally leveled wings followed by roll <strong>and</strong> a diverging<br />

spiral maneuver.<br />

The term (λ+f) describes the subsidiary roll mode while the 2nd-order system is referred<br />

to as Dutch roll. This is an oscillatory system with a small relative damping factor ζ D . The<br />

natural frequency in Dutch roll is denoted ω D .<br />

If the lateral B-767 Matlab model in Chapter 4 is considered, the Matlab comm<strong>and</strong><br />

damp.m gives:<br />


2.11. DESIGN OF FLIGHT CONTROL SYSTEMS 29<br />

a = [<br />

-0.1245 0.0350 0.0414 -0.9962<br />

-15.2138 -2.0587 0.0032 0.6450<br />

0 1.0000 0 0.0357<br />

1.6447 -0.0447 -0.0022 -0.1416<br />

damp(a)<br />

Eigenvalue Damping Freq. (rad/sec)<br />

-0.0143 1.0000 0.0143<br />

-0.1121 + 1.4996i 0.0745 1.5038<br />

-0.1121 - 1.4996i 0.0745 1.5038<br />

-2.0863 1.0000 2.0863<br />

In this example we only get four eigenvalues since the pure integrator in yaw is not<br />

include in the system matrix. It is seen that the spiral mode is given by e = −0.0143 while<br />

the subsidiary roll mode is given by f = −2.0863. Dutch roll is recognized by ω D = 0.0745<br />

<strong>and</strong> ζ D = 1.5038.<br />

2.11 Design <strong>of</strong> flight control systems<br />

A detailed introduction to design <strong>of</strong> flight control systems are given by [7], [11], [8] <strong>and</strong> [12].<br />

The control system are based on linear design techniques using linearized models similar to<br />

those discussed in the sections above.


30 CHAPTER 2. AIRCRAFT MODELING


Chapter 3<br />

Satellite Modeling<br />

When stabilizing satellites in geostationary orbits only the attitude <strong>of</strong> the satellite is <strong>of</strong><br />

interest since the position is given by the Earth’s rotation. Figure 3.1 shows the NUTS<br />

CubeSat project at <strong>NTNU</strong>.<br />

3.1 Attitude Model<br />

3.1.1 Euler’s 2nd Axiom applied to satellites<br />

The rigid-body kinetics (2.21) gives:<br />

I CG ˙ω + ω × (I CG ω) = τ (3.1)<br />

3<br />

were ω = [p, q, r] ⊤ <strong>and</strong> I CG = I ⊤ CG > 0 is the inertia tensor about the center <strong>of</strong> gravity given<br />

by:<br />

2<br />

I x −I xy −I xz<br />

−I xz −I yz I z<br />

I CG = 4 −I xy I y −I yz<br />

5 (3.2)<br />

Both the Euler angles <strong>and</strong> quaternions can be used to represent attitude. Moreover,<br />

˙q = T q (q)ω (3.3)<br />

Figure 3.1: The NUTS CubeSat project at <strong>NTNU</strong>- Courtesy to http://nuts.cubesat.no/<br />

31


32 CHAPTER 3. SATELLITE MODELING<br />

where q = [η, ε 1 , ε 2 , ε 3 ] ⊤ , θ = [Φ, Θ, Ψ] ⊤ <strong>and</strong><br />

T θ (θ) =<br />

2<br />

4<br />

T q (q) = 1 6<br />

24<br />

˙θ = T θ (θ)ω (3.4)<br />

1 sin(Φ) tan(Θ) cos(Φ) tan(Θ)<br />

0 cos(Φ) − sin(Φ)<br />

0 sin(Φ)/ cos(Θ) cos(Φ)/ cos(Θ)<br />

2<br />

−ε 1 −ε 2 −ε 3<br />

η −ε 3 ε 2<br />

ε 3 η −ε 1<br />

−ε 2 ε 1 η<br />

3<br />

3<br />

5 , cos(Θ) ≠ 0 (3.5)<br />

7<br />

5 (3.6)<br />

The satellite has three controllable moments τ = [τ 1 , τ 2 , τ 3 ] ⊤ , which can be generated using<br />

different actuators. Magnetic actuators are described in Section 3.2.<br />

3.1.2 Skew-symmetric representation <strong>of</strong> the satellite model<br />

The dynamic model (3.1) can also be written:<br />

I CG ˙ω − (I CG<br />

ω) × ω = τ (3.7)<br />

where we have used the fact that a × b = −b × a. Furthermore, it is possible to find a<br />

matrix S(ω) such that S(ω) = −S ⊤ (ω) is skew-symmetric. With other words:<br />

The matrix S(ω) must satisfy the condition:<br />

I CG ˙ω + S(ω)ω = τ (3.8)<br />

S(ω)ω ≡ −(I CG<br />

ω) × ω (3.9)<br />

A matrix satisfying this condition is:<br />

2<br />

0 −I yz q − I xz p + I z r I yz r + I xy p − I y q<br />

3<br />

S(ω) = 4 I yz q + I xz p − I z r 0 −I xz r − I xy q + I x p 5 (3.10)<br />

−I yz r − I xy p + I y q I xz r + I xy q − I x p 0<br />

3.2 Actuator Dynamics<br />

Magnetic actuators are used to control the NUTS cube-satellite, which is a student satellite<br />

project at <strong>NTNU</strong>. The magnetic actuators are cheap, solid-state <strong>and</strong> energy-effi cient. Due<br />

to variations <strong>of</strong> the magnetic field <strong>for</strong> a moving satellite, modeling <strong>and</strong> control may be<br />

challenging. An analogy could be an airplane moving through a various density atmospheres,<br />

suddenly leaving you with limited control <strong>of</strong> either roll, pitch or yaw.<br />

When applying an electrical current through the windings <strong>of</strong> the coils, a magnetic dipole<br />

moment is generated. When this field reacts with the magnetic field surrounding the Earth,<br />

a magnetic torque is generated, which makes the satellite move. Let the Earth´s magnetic


3.2. ACTUATOR DYNAMICS 33<br />

field in BODY coordinates be denoted by B b = [B x , B y , B z ] ⊤ while the magnetic field m b =<br />

[m x , m y , m z ] ⊤ set up by the actuators is:<br />

m b =<br />

2<br />

6<br />

4<br />

N xA xV x<br />

R x<br />

N yA yV y<br />

R y<br />

N zA zV z<br />

R z<br />

3<br />

7<br />

5 (3.11)<br />

= K coil V (3.12)<br />

where<br />

<strong>and</strong> <strong>for</strong> n = x, y, z,<br />

K coil =<br />

2<br />

6<br />

4<br />

N xA x<br />

R x<br />

0 0<br />

N<br />

0 yA y<br />

R y<br />

0<br />

N<br />

0 0 zA z<br />

R z<br />

3<br />

7<br />

5 , V =<br />

2<br />

4<br />

V x<br />

3<br />

V y<br />

5 (3.13)<br />

N n<br />

A n<br />

V n<br />

R n<br />

is the number <strong>of</strong> turns with copper thread in the coil<br />

is the area <strong>of</strong> the coil<br />

is the voltage over the coil<br />

is the resistance <strong>of</strong> the coil<br />

Then, the torque becomes:<br />

where<br />

S(−B b ) = 4<br />

τ = m b × B b<br />

= S(−B b )m b (3.14)<br />

2<br />

0 B z −B y<br />

−B z 0 B x<br />

B y −B x 0<br />

3<br />

5 (3.15)<br />

For control purposes, we usually assign τ , however, the satellite only accepts voltage signals<br />

as input, <strong>and</strong> not torque. It is straight<strong>for</strong>ward to find the magnetic field m b due to the<br />

actuators:<br />

m b = τ × B b (3.16)<br />

In view <strong>of</strong> (3.13), one may find the relation between voltages <strong>and</strong> the torque vector:<br />

V = K −1<br />

coil<br />

τ × B b (3.17)<br />

The matrix K coil , usually referred to as the control allocation matrix, is a constant matrix.<br />

Note that B b will change throughout the orbit.<br />

According to (3.14), τ b always lies in the plane perpendicular to B b . Suppose that τ d is<br />

the torque that the controller produces. It is required that τ d lies in the plane perpendicular


34 CHAPTER 3. SATELLITE MODELING<br />

to B b . It implies that τ only represents the component <strong>of</strong> τ d that is perpendicular to B b . To<br />

avoid unnecessary use <strong>of</strong> electricity, the length <strong>of</strong> τ d is scaled down by 1/ B b 2 . Accordingly,<br />

m b =<br />

⇓<br />

τ =<br />

1<br />

‖B b ‖ 2 (τ d × B b ) (3.18)<br />

1<br />

‖B b ‖ 2 (τ d × B b ) × B b (3.19)<br />

Let γ be the angle between τ d <strong>and</strong> B b . The magnitude <strong>of</strong> τ is given by:<br />

‖τ ‖ = 1<br />

‖B b ‖ 2 ‖τ d‖ B<br />

b sin(γ)<br />

B b <br />

= ‖τ d ‖ sin(γ) (3.20)<br />

3.3 Design <strong>of</strong> Satellite Attitude <strong>Control</strong> Systems<br />

Design <strong>of</strong> nonlinear trajectory-tracking control systems <strong>for</strong> Hamiltonian systems in the <strong>for</strong>m<br />

(3.8) is quite common in the literature. One example is the nonlinear <strong>and</strong> passive adaptive<br />

attitude control system <strong>of</strong> Slotine <strong>and</strong> Di Benedetto [10]. An alternative representation<br />

is proposed by Fossen [3], which simplifies the representation <strong>of</strong> the regressor matrix.<br />

Trajectory-tracking controllers require that the model parameters are known or estimated<br />

using parameter adaptation. For set-point regulation is possible to derive controllers that<br />

do not require model parameters. This is presented below.<br />

3.3.1 Nonlinear quaternion set-point regulator<br />

Consider the Lyapunov function c<strong>and</strong>idate:<br />

V = 1 2 ω⊤ I CG ω + h(˜q) (3.21)<br />

where h(˜q) is a positive definite function depending <strong>of</strong> the attitude error, <strong>for</strong> instance:<br />

h(˜q) = 1 2˜q⊤ K p˜q, ˜q = q − q d (3.22)<br />

where q d = constant <strong>and</strong> K p > 0 is a design matrix. This means that the desired attitude<br />

must be specified as a unit quaternion, which usually is computed by specifying the desired<br />

Euler angles <strong>and</strong> trans<strong>for</strong>ming these to a desired unit quaternion.<br />

Differentiation <strong>of</strong> V with respect to time gives:<br />

˙V = ω T I CG ˙ω + ˙q ⊤ ∂h<br />

∂˜q<br />

= ω ⊤ (−S(ω)ω + τ ) + ω ⊤ T ⊤ q (q) ∂h<br />

∂˜q<br />

(3.23)


3.3. DESIGN OF SATELLITE ATTITUDE CONTROL SYSTEMS 35<br />

Since S(ω) = −S ⊤ (ω), it follows that:<br />

This suggests that the control input τ should be chosen as:<br />

ω ⊤ S(ω)ω = 0 ∀ω (3.24)<br />

τ = −K d ω − T ⊤ q (q) ∂h<br />

∂˜q<br />

= −K d ω − T ⊤ q (q)K p˜q (3.25)<br />

where K d > 0. such that<br />

<br />

˙V = ω ⊤ τ + T ⊤ q (q) ∂h <br />

∂˜q<br />

= −ω ⊤ K d ω ≤ 0 (3.26)<br />

<strong>and</strong> stability <strong>and</strong> convergence follow from st<strong>and</strong>ard Lyapunov techniques, <strong>for</strong> instance by<br />

using Barbalat’s lemma.<br />

3.3.2 PD-controller <strong>of</strong> Wen <strong>and</strong> Kreutz-Delago (1991)<br />

The classical PD-controller <strong>for</strong> satellite attitude regulation uses feedback from the imaginary<br />

part ˜ε = [˜ε 1 , ˜ε 2 , ˜ε 3 ] ⊤ <strong>of</strong> the unit quaternion error ˜q = [˜η, ˜ε 1 , ˜ε 2 , ˜ε 3 ] ⊤ where ˜η denotes the<br />

real part <strong>of</strong> the unit quaternion. The attitude error dynamics ˜q can be described by unit<br />

quaternions according to:<br />

˙˜η = 1 2 ω⊤˜ε (3.27)<br />

˙˜ε = − 1 2˜ηω − 1 ω × ˜ε (3.28)<br />

2<br />

The following PD-controller is proposed by Wen <strong>and</strong> Kreutz-Delgado [13]:<br />

τ = k p˜ε − k d ω (3.29)<br />

where k p > 0 <strong>and</strong> k d > 0 are the controller gains. Notice that the control law does not<br />

depends on the model <strong>of</strong> the system, <strong>and</strong> it makes ω <strong>and</strong> ˜ε asymptotically converge to zero<br />

<strong>and</strong> hence ˜η converges to zero. Consider the following scalar function:<br />

V = (k p + ck d ) (˜η − 1) 2 + ˜ε ⊤˜ε + 1 2 ω⊤ I CG ω − cε ⊤ I CG ω (3.30)<br />

One may show that there exists a suffi ciently small c > 0 such that V > 0 <strong>for</strong> all ˜η ≠ 1,<br />

˜ε ≠ 0, <strong>and</strong> ω ≠ 0. In fact, V is bounded from below as<br />

V ≥ (k p + ck d )(˜η − 1) 2 + 1 2<br />

<br />

˜ε ω<br />

2(k p + ck d ) cγ I<br />

cγ I µ I<br />

˜ε<br />

ω<br />

<br />

(3.31)


36 CHAPTER 3. SATELLITE MODELING<br />

where γ I = ‖I CG ‖ <strong>and</strong> µ I = inf ‖v‖=1 v ⊤ I CG v. µ I > 0 since the inertia matrix I CG is positive<br />

definite. We take the time derivative <strong>of</strong> V :<br />

˙V = 2(k p + ck d ) (˜η − 1)˙˜η + ˜ε ⊤ ˙˜ε + ω ⊤ I CG ˙ω − c˙˜ε ⊤ I CG ω − c˜ε ⊤ I CG ˙ω (3.32)<br />

It is straight<strong>for</strong>ward to find that (˜η − 1)˙˜η + ˜ε ⊤ ˙˜ε = −(1/2)ω ⊤˜ε since ˜ε ⊤ (ω × ˜ε) = 0. Also,<br />

ω ⊤ I CG ˙ω = ω ⊤ (τ − ω × I CG ω) = ω ⊤ τ . Moreover, we can find that<br />

<br />

−c˙˜ε ⊤ I CG ω − c˜ε ⊤ I CG ˙ω = −c − 2˜ηω 1 − 1 ⊤<br />

2 ω × ˜ε I CG ω − c˜ε ⊤ (τ − ω × I CG ω) (3.33)<br />

Thus, together with the control law ((3.29)), the time derivative will be<br />

˙V = −ck p ‖˜ε‖ 2 − k 2 d ‖ω‖ 2 − c −<br />

1 2˜ηω − 1 2 ω × ˜ε ⊤<br />

I CG ω − c˜ε ⊤ (−ω × I CG ω)<br />

≤ − ˜ε<br />

ω <br />

⊤ ckp 0 ˜ε<br />

0 k d − 2cγ I ω<br />

⊤<br />

(3.34)<br />

˙V ≤ 0; then, from Barbalat’s lemma, it follows that ˜ε → 0 <strong>and</strong> ω → 0 as time tends to<br />

infinity.<br />

As explained by (3.11)—(3.13), the input signal is voltage V <strong>of</strong> the coil. Thus, we need<br />

to find the relation between τ <strong>and</strong> V. In practical situations, scaling is also required so as<br />

to avoid unnecessary electricity load. There<strong>for</strong>e, one may derive the following input signal:<br />

V = K −1<br />

coil<br />

τ × B b<br />

= k p<br />

‖B b ‖ 2 K−1 coil<br />

˜ε × B b − k d<br />

‖B b ‖ 2 K−1 coil<br />

ω × B b (3.35)


Chapter 4<br />

Matlab Simulation <strong>Models</strong><br />

4.1 Boeing-767<br />

The longitudinal <strong>and</strong> lateral B-767 state-space models are given below. The state vectors<br />

are:<br />

2 3 2 3<br />

u (ft/s)<br />

β (deg)<br />

x lang = 6 α (deg)<br />

7<br />

4 q (deg/s) 5 , x lat = 6 p (deg/s)<br />

7<br />

4 φ (deg/s) 5 (4.1)<br />

θ (deg)<br />

r (deg)<br />

Equilibrium point:<br />

u lang =<br />

<br />

δE (deg)<br />

δ T (%)<br />

Speed V T = 890 ft/s = 980 km/h<br />

Altitude h = 35 000 ft<br />

Mass m = 184 000 lbs<br />

Mach-number M = 0.8<br />

4.1.1 Longitudinal model<br />

<br />

, u lat =<br />

a = [ -0.0168 0.1121 0.0003 -0.5608<br />

-0.0164 -0.7771 0.9945 0.0015<br />

-0.0417 -3.6595 -0.9544 0<br />

0 0 1.0000 0];<br />

b = [ -0.0243 0.0519<br />

-0.0634 -0.0005<br />

-3.6942 0.0243<br />

0 0 ];<br />

<br />

δA (deg)<br />

δ R (deg)<br />

<br />

(4.2)<br />

37


38 CHAPTER 4. MATLAB SIMULATION MODELS<br />

Eigenvalues:<br />

lam = [<br />

-0.8678 + 1.9061i<br />

-0.8678 - 1.9061i<br />

-0.0064 + 0.0593i<br />

-0.0064 - 0.0593i];<br />

4.1.2 Lateral model<br />

a = [<br />

-0.1245 0.0350 0.0414 -0.9962<br />

-15.2138 -2.0587 0.0032 0.6458<br />

0 1.0000 0 0.0357<br />

1.6447 -0.0447 -0.0022 -0.1416];<br />

b = [<br />

-0.0049 0.0237<br />

-4.0379 0.9613<br />

0 0<br />

-0.0568 -1.2168];<br />

Eigenvalues:<br />

lam = [<br />

-0.1121 + 1.4996i<br />

-0.1121 - 1.4996i<br />

-2.0863<br />

-0.0143];<br />

4.2 F-16 Fighter<br />

The lateral model <strong>of</strong> the F-16 fighter aircraft is based on [11], pages 370—371. The state<br />

vectors are:<br />

2 3<br />

β (ft/s)<br />

φ (ft/s)<br />

2 3<br />

p (rad/s)<br />

<br />

r w (deg)<br />

x lat =<br />

r (rad)<br />

uA (rad)<br />

, u lat =<br />

, y 6 δ A (rad)<br />

u R (rad) lat = 6 p (deg/s)<br />

7<br />

4 β (deg) 5 (4.3)<br />

7<br />

4 δ R (rad) 5<br />

φ (deg)<br />

r w (rad)<br />

Equilibrium point:<br />

Speed V T = 502 ft/s = 552 km/h<br />

Mach-number M = 0.45


4.2. F-16 FIGHTER 39<br />

4.2.1 Longitudinal model<br />

a = [<br />

-0.3220 0.0640 0.0364 -0.9917 0.0003 0.0008 0<br />

0 0 1 -0.0037 0 0 0<br />

-30.6492 0 -3.6784 0.6646 -0.7333 0.1315 0<br />

8.5396 0 -0.0254 -0.4764 -0.0319 -0.0620 0<br />

0 0 0 0 -20.2 0 0<br />

0 0 0 0 0 -20.2 0<br />

0 0 0 57.2958 0 0 -1 ];<br />

b = [<br />

0 0<br />

0 0<br />

0 0<br />

0 0<br />

20.2 0<br />

0 20.2<br />

0 0 ];<br />

c= [<br />

0 0 0 57.2958 0 0 -1<br />

0 0 57.2958 0 0 0 0<br />

57.2958 0 0 0 0 0 0<br />

0 57.2958 0 0 0 0 0 ];<br />

Eigenvalues:<br />

lam = [<br />

-1.0000<br />

-0.4224+ 3.0633i<br />

-0.4224- 3.0633i<br />

-0.0167<br />

-3.6152<br />

-20.2000<br />

-20.2000 ];<br />

Notice that the last two eigenvalues correspond to the actuator states.


40 CHAPTER 4. MATLAB SIMULATION MODELS<br />

Figure 4.1: Schematic drawing <strong>of</strong> the Bristol F.2B Fighter (McRuer et al 1973).<br />

4.3 F2B Bristol Fighter<br />

The lateral model <strong>of</strong> the F2B Bristol fighter aircraft is given below [8]. This is a British<br />

aircraft from World War I (see Figure 4.1). The aircraft model is <strong>for</strong> β = 0 (coordinated<br />

turn). The state-space vector is:<br />

x lat =<br />

2<br />

6<br />

4<br />

p (deg/s)<br />

r (deg/s)<br />

φ (deg)<br />

ψ (deg)<br />

where the dynamics <strong>for</strong> ψ satisfies the bank-to-turn equation:<br />

Equilibrium point:<br />

˙ψ = g U 0<br />

φ =<br />

Speed V T = 138 ft/s = 151.4 km/h<br />

Altitude h = 6 000 ft<br />

Mach-number M = 0.126<br />

3<br />

7<br />

5 , u lat = [δ A (deg)] (4.4)<br />

9.81<br />

φ = 0. 233φ (4.5)<br />

138 · 0.3048


4.4. <strong>NTNU</strong> STUDENT CUBE-SATELLITE 41<br />

4.3.1 Lateral model<br />

a = [<br />

-7.1700 2.0600 0 0<br />

-0.4360 -0.3410 0 0<br />

1.0000 0 0 0<br />

0 0 0.2330 0];<br />

b = [<br />

26.1000<br />

-1.6600<br />

0<br />

0];<br />

Eigenvalues:<br />

lam = [<br />

0<br />

0<br />

-0.4752<br />

-7.0358];<br />

4.4 <strong>NTNU</strong> student cube-satellite<br />

The <strong>NTNU</strong> student cube-satellite shown in Figure 3.1 has the following inertia matrix:<br />

I_CG = diag{[0.0108,0.0113,0.0048]};<br />

% kg m^2


42 CHAPTER 4. MATLAB SIMULATION MODELS


Bibliography<br />

[1] Blakelock, J. H. (1991). <strong>Aircraft</strong> <strong>and</strong> Missiles (John Wiley & Sons Ltd.)<br />

[2] Etkin, B. <strong>and</strong> L. D. Reid (1996). Dynamics <strong>of</strong> Flight: Stability <strong>and</strong> <strong>Control</strong> (John<br />

Wiley & Sons Ltd.)<br />

[3] Fossen, T. I. (1993). Comments on ”Hamiltonian Adaptive <strong>Control</strong> <strong>of</strong> Spacecraft”,<br />

IEEE Transactions on Automatic <strong>Control</strong>, TAC-38(5):671—672.<br />

[4] Fossen, T. I. (1994). Guidance <strong>and</strong> <strong>Control</strong> <strong>of</strong> Ocean Vehicles (John Wiley & Sons<br />

Ltd.)<br />

[5] Fossen, T. I. (2011). H<strong>and</strong>book <strong>of</strong> Marine Craft Hydrodynamics <strong>and</strong> Motion <strong>Control</strong>.<br />

(John Wiley & Sons Ltd.)<br />

[6] Hughes, P. C. (1986). Spacecraft Attitude Dynamics (John Wiley & Sons Ltd.)<br />

[7] McLean, D. (1990). Automatic Flight <strong>Control</strong> Systems (Prentice Hall Inc.)<br />

[8] McRuer, D., D. Ashkenas og A. I. Graham (1973). <strong>Aircraft</strong> Dynamics <strong>and</strong><br />

Automatic <strong>Control</strong>. (Princeton University Press, New Jersey 1973).<br />

[9] Nelson R. C. (1998). Flight Stability <strong>and</strong> Automatic <strong>Control</strong> (McGraw-Hill Int.)<br />

[10] Slotine, J. J. E. og M. D. Di Benedetto (1990). Hamiltonian Adaptive <strong>Control</strong><br />

<strong>of</strong> Spacecraft, IEEE Transactions on Automatic <strong>Control</strong>, TAC-35(7):848—852.<br />

[11] Stevens, B. L. og F. L. Lewis (1992). <strong>Aircraft</strong> <strong>Control</strong> <strong>and</strong> Simulation (John Wiley<br />

& Sons Ltd.)<br />

[12] Roskam, J. (1999). Airplane Flight Dynamics <strong>and</strong> Automatic Flight <strong>Control</strong>s (Darcorporation)<br />

[13] Wen, J. T.-Y. <strong>and</strong> K. Kreutz-Delgado (1991). The Attitude <strong>Control</strong> Problem.<br />

IEEE Transactions on Automatic <strong>Control</strong>, Vol. 36, No. 10. October 1991.<br />

43

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!