Sie sind auf Seite 1von 172

GROUP FOR AERONAUTICAL RESEARCH AND TECHNOLOGY IN EUROPE

ORIGINAL: ENGLISH GARTEUR/TP-088-27


April 4, 1997

GARTEUR Open

Description and Analysis of the


Research Civil Aircraft Model (RCAM)

by

Gertjan Looye and Samir Bennani

GARTEUR aims at stimulating and co-ordinating


co-operation between Research Establishments and Industry
in the areas of Aerodynamics, Flight Mechanics, Helicopters,
Structures & Materials and Propulsion Technology
GROUP FOR AERONAUTICAL RESEARCH AND TECHNOLOGY IN EUROPE

ORIGINAL: ENGLISH GARTEUR/TP-088-27


April 4, 1997

GARTEUR Open

Description and Analysis of the


Research Civil Aircraft Model (RCAM)

by

Gertjan Looye and Samir Bennani

This report has been prepared under auspices of


the Responsables for Flight Mechanics, Systems
and Integration of the Group for Aeronautical
Research and Technology in EURope (GARTEUR)

Group of Resp. : FM-GoR Action Group : FM(AG08)


Report Resp. : G. Looye/ Version : 1
Project Man. : J.C. Terlouw/ Completed : April 4, 1997
Monitoring Resp. : J.T.M. van Doorn/ c GARTEUR 1997
Version: 1
i Date: April 4, 1997
GARTEUR/TP-088-27

List of authors
Gertjan Looye1 ,
Samir Bennani2

1 Delft University of Technology, fac. of Aerospace Engineering, sect. Stability & Control
from 1/10/97: DLR Oberpfa enhofen, German Aerospace Research Establishment,
Institute for Robotics and System Dynamics
P.O. Box 1116, D-82230 Wessling, Germany
E-mail: Gertjan.Looye@dlr.de
2 Delft University of Technology, fac. of Aerospace Engineering, sect. Stability & Control
Kluyverweg 1, 2629 HS Delft, The Netherlands
E-mail: S.Bennani@lr.tudelft.nl
Version: 1
Date: April 4, 1997 ii
GARTEUR/TP-088-27

Summary
This report contains a description and an analysis of the Research Civil Aircraft Model
(RCAM). The RCAM is based on a modi ed civil aircraft during the nal approach. It
is used for controller design and simulation in the civil variant of the two Robust Flight
Control design challenges, prepared by the GARTEUR Action Group FM(AG08). A set
of design requirements for the design challenge is given in the RCAM problem formulation
manual [2].
This document has two main objectives:
 documentation of the aircraft modeling and disturbance modeling (windshear, gust),
including implementation in Matlab/Simulink;
 open loop analysis of the RCAM as a basis for the design of a ight controller. This
analysis includes performance limitations, dynamic characteristics and co-ordination
of controls in speci c situations, like turns, windshear, and engine failure.
The analysis is carried out using linear system theory, point-mass approximations and
steady state analysis in several ight situations. Especially energy concepts and Korham-
mer diagrams have proven to be very illuminating.
The longitudinal performance and dynamics of the RCAM are representative for a civil
airliner, while the lateral dynamics have strange properties, like a very fast spiral mode.
In fact, in spite of many simpli cations, the most interesting ight situations are covered
by the model, including backside operation and stall.
Version: 1
iii Date: April 4, 1997
GARTEUR/TP-088-27

Distribution list
FM(AG08) Principal Persons
Ahmed S (UK) CCL 1 copy
Ambrosino G (IT) UNAP 1 copy
Bernussou J (FR) LAAS 1 copy
Cruz J de la (ES) UCM 1 copy
Delgado I (ES) INTA 1 copy
Dormido S (ES) UNED 1 copy
Duda H (DE) DLR 1 copy
Escande B (FR) ONERA 1 copy
Faleiro L (UK) LUT 1 copy
Fielding C (UK) BAe-MA 1 copy
Game G (UK) BAeDD 1 copy
Gautrey J (UK) CUN 1 copy
Helmersson A (SE) LiTH 1 copy
Joos H (DE) DLR 1 copy
Laidlaw D (UK) AVRO 1 copy
Luckner R (DE) DASA 1 copy
Maciejowski J (UK) UCAM 1 copy
Magni J (FR) ONERA 1 copy
Muir E (UK) DRA 1 copy
Postlethwaite I (UK) ULES 1 copy
Schram G (NL) DUT-EE 1 copy
Stahl-Gunnarsson K (SE) SMA 1 copy
Terlouw J (NL) NLR 1 copy
Tonon A (IT) ALN 1 copy
Vaart J van der (NL) DUT-AE 1 copy
Verde L (IT) CIRA 1 copy
Members of the Flight Mechanics, Systems and Integration Group of
Responsables
Brannstrom B (SE) FMV 1 copy
Doorn J van (NL) NLR 1 copy
England P (UK) DRA 1 copy
Mu~noz F (ES) INTA 1 copy
Rodlo R (DE) DLR 1 copy
Verbrugge R (FR) ONERA 1 copy
Version: 1
Date: April 4, 1997 iv
GARTEUR/TP-088-27

Members of the Executive Committee


Cheret J (FR) ONERA 1 copy
Coleman G (UK) DRA 1 copy
Gustafsson A (SE) FFA 1 copy
Haupt R (DE) DLR/PD-L 1 copy
Moreno Labata G (ES) INTA 1 copy
Sloo J (NL) NLR 1 copy
Secretary GARTEUR
Jonasson K (SE) FFA 1 copy
Version: 1
v Date: April 4, 1997
GARTEUR/TP-088-27

Contents

List of gures viii


List of tables xi
List of symbols and abbreviations xii
1 Introduction 1
2 Description of the aircraft simulation model 3
2.1 The nonlinear aircraft model . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.1 Assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.2 Axis systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.3 The generic state equations of the aircraft motions . . . . . . . . 10
2.1.4 The aerodynamic model of the RCAM . . . . . . . . . . . . . . . 13
2.1.5 Thrust contribution to forces and moments . . . . . . . . . . . . 19
2.2 The actuator and engine models . . . . . . . . . . . . . . . . . . . . . . 19
2.3 The wind models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3.1 Constant wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3.2 Turbulence models . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.3 Windshear model . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.4 The implementation of the models in the software . . . . . . . . . . . . 27
2.4.1 The simulation environment . . . . . . . . . . . . . . . . . . . . 27
2.4.2 Trimming and linearization routines . . . . . . . . . . . . . . . . 28
3 Point-mass and steady state analysis of the aircraft dynamics 31
3.1 Point-mass ight mechanics . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.1.1 The reduced equations of motion . . . . . . . . . . . . . . . . . . 31
3.1.2 Energy considerations and the Korhammer diagram . . . . . . . 33
3.1.3 Aircraft responses to control inputs . . . . . . . . . . . . . . . . 37
3.2 Performance in steady symmetric ight . . . . . . . . . . . . . . . . . . 43
3.2.1 Basic relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2.2 Analytical performance calculations . . . . . . . . . . . . . . . . 44
3.2.3 Performance diagrams . . . . . . . . . . . . . . . . . . . . . . . . 47
3.2.4 Analytical expressions for weight variations . . . . . . . . . . . . 51
3.3 Turning performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.3.1 Equilibrium in co-ordinated turns . . . . . . . . . . . . . . . . . 53
3.3.2 Parameters for turning performance . . . . . . . . . . . . . . . . 54
3.3.3 The minimum turning radius . . . . . . . . . . . . . . . . . . . . 56
3.3.4 The maximum turning rate . . . . . . . . . . . . . . . . . . . . . 58
Version: 1
Date: April 4, 1997 vi
GARTEUR/TP-088-27

3.3.5 Climb performance in turns . . . . . . . . . . . . . . . . . . . . . 59


3.3.6 Control de ections in co-ordinated turns . . . . . . . . . . . . . 60
3.4 The e ect of steady wind . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.4.1 Tail wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.4.2 Vertical wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.4.3 Cross wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.4.4 Time responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.5 The e ect of windshear . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.5.1 The F -factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.5.2 Energy altitude reserve . . . . . . . . . . . . . . . . . . . . . . . 69
3.5.3 Time responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.6 Aircraft control during engine failure . . . . . . . . . . . . . . . . . . . . 71
3.6.1 Control strategies . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.6.2 The minimum control speed . . . . . . . . . . . . . . . . . . . . 77
3.6.3 Climb performance with engine failure . . . . . . . . . . . . . . . 78
3.6.4 Time responses . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4 Open loop analysis of the dynamic aircraft model 82
4.1 The linearized aircraft models . . . . . . . . . . . . . . . . . . . . . . . 82
4.2 Transfer functions and block diagrams . . . . . . . . . . . . . . . . . . . 84
4.3 Approximations for the longitudinal modes . . . . . . . . . . . . . . . . 86
4.3.1 The linearized longitudinal model . . . . . . . . . . . . . . . . . 86
4.3.2 Approximations for the short period . . . . . . . . . . . . . . . . 87
4.3.3 Approximations for the phugoid . . . . . . . . . . . . . . . . . . 90
4.4 Approximations for the lateral modes . . . . . . . . . . . . . . . . . . . 101
4.4.1 The linearized lateral model . . . . . . . . . . . . . . . . . . . . 101
4.4.2 Approximation of the aperiodic roll motion . . . . . . . . . . . . 103
4.4.3 Approximation of the Dutch roll . . . . . . . . . . . . . . . . . . 103
4.4.4 Spiral stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.5 Control capacity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 106
4.5.1 Longitudinal responses . . . . . . . . . . . . . . . . . . . . . . . 106
4.5.2 Longitudinal non-minimum phase responses . . . . . . . . . . . . 109
4.5.3 Lateral responses . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.6 Aircraft responses to disturbance inputs . . . . . . . . . . . . . . . . . . 114
4.6.1 Longitudinal wind input . . . . . . . . . . . . . . . . . . . . . . . 114
4.6.2 Vertical wind input . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.6.3 Lateral wind input . . . . . . . . . . . . . . . . . . . . . . . . . . 115
5 Conclusions 117
References 118
Version: 1
vii Date: April 4, 1997
GARTEUR/TP-088-27

A The e ect of the center of gravity location 120


B Software for the RCAM nonlinear model 123
B.1 Implementation in Simulink-blocks . . . . . . . . . . . . . . . . . . . . . 123
B.2 The routine TRIMRCAM.M . . . . . . . . . . . . . . . . . . . . . . . . 127
B.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
B.2.2 Trimming in a wind eld . . . . . . . . . . . . . . . . . . . . . . . 127
B.2.3 linearization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
B.2.4 Simulations for veri cation . . . . . . . . . . . . . . . . . . . . . 130
C linearized RCAM models 132
D Coecients for linear models 135
E Characteristics of linearized models 136
F RCAM simulations 142
G About the authors 153
Version: 1
Date: April 4, 1997 viii
GARTEUR/TP-088-27

List of gures
2.1 The body- xed, vehicle carried-vertical and the earth- xed axes systems 5
2.2 The vehicle-carried vertical and the ightpath axes systems . . . . . . . 6
2.3 The body- xed and the ightpath axes systems . . . . . . . . . . . . . . 7
2.4 The wind, stability and body- xed axes systems . . . . . . . . . . . . . 7
2.5 Transformation from the vehicle-carried system to the wind axes system 8
2.6 The de nition of wind angles . . . . . . . . . . . . . . . . . . . . . . . . 9
2.7 De nitions of lift and drag . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.8 Curves of CL and CD vs. . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.9 The lift-drag polar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.10 Location of moment points . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.11 Power spectral densities of Dryden lters . . . . . . . . . . . . . . . . . 23
2.12 Location of vortices of windshear model . . . . . . . . . . . . . . . . . . 25
2.13 Wind eld due to windshear and downburst . . . . . . . . . . . . . . . . 25
2.14 Encountered wind speeds in windshear/downburst . . . . . . . . . . . . 26
2.15 Simulink block diagram of the RCAM simulation environment . . . . . . 27
3.1 Longitudinal aircraft motion . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2 Force equilibrium for unpowered ight in undisturbed atmosphere . . . 32
3.3 Korhammer Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.4 Korhammer Diagram; accelerated ight . . . . . . . . . . . . . . . . . . 36
3.5 E ect of changing aerodynamic trim point in Korhammer diagram: angle
of attack and ap de ection . . . . . . . . . . . . . . . . . . . . . . . . 38
3.6 Korhammer Diagram; ight below minimum drag speed . . . . . . . . . 39
3.7 Korhammer Diagram; e ect of thrust commands . . . . . . . . . . . . . 40
3.8 Performance related to the lift-drag polar . . . . . . . . . . . . . . . . . 44
3.9 Modi cation of the lift-drag polar . . . . . . . . . . . . . . . . . . . . . . 45
3.10 Performance diagrams for di erent masses (forces) . . . . . . . . . . . . 47
3.11 RCAM performance diagrams (power) . . . . . . . . . . . . . . . . . . . 48
3.12 Korhammer Diagram; maximum ightpath angle . . . . . . . . . . . . . 49
3.13 Maximum ightpath angle vs speed . . . . . . . . . . . . . . . . . . . . . 50
3.14 Minimum ightpath angle vs speed . . . . . . . . . . . . . . . . . . . . . 50
3.15 Aircraft in co-ordinated turn . . . . . . . . . . . . . . . . . . . . . . . . 53
3.16 RCAM performance diagrams for di erent bank angles (force) . . . . . . 55
3.17 RCAM performance diagrams for di erent bank angles (power) . . . . . 56
3.18 Limitations on the turning radius . . . . . . . . . . . . . . . . . . . . . . 57
3.19 Limitations on the turning rate . . . . . . . . . . . . . . . . . . . . . . . 59
3.20 Modi ed Korhammer diagram for a steady turn . . . . . . . . . . . . . . 60
3.21 Maximum ightpath angle vs speed in turn . . . . . . . . . . . . . . . . 61
3.22 De nition of wind vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Version: 1
ix Date: April 4, 1997
GARTEUR/TP-088-27

3.23 Winds shear F -factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68


3.24 Aircraft response to windshear . . . . . . . . . . . . . . . . . . . . . . . 71
3.25 Asymmetric equilibrium for zero sideslip . . . . . . . . . . . . . . . . . . 73
3.26 Asymmetric equilibrium for zero bank angle . . . . . . . . . . . . . . . . 75
3.27 Asymmetric equilibrium for zero rudder de ection . . . . . . . . . . . . 76
3.28 Performance diagram for engine failure; bank angles . . . . . . . . . . . 79
3.29 Maximum ightpath angle vs speed, engine failure . . . . . . . . . . . . 80
3.30 Aircraft response to right engine failure . . . . . . . . . . . . . . . . . . 81
4.1 Block diagram of example system . . . . . . . . . . . . . . . . . . . . . . 84
4.2 Block diagram of the longitudinal state equations . . . . . . . . . . . . . 88
4.3 Phugoid simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
4.4 Animation of phugoid . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.5 Speci c energy rate and distribution rate . . . . . . . . . . . . . . . . . 95
4.6 Korhammer diagram example with labelled elements . . . . . . . . . . . 95
4.7 Korhammer diagrams at di erent time points during phugoid period (A) 96
4.8 Korhammer diagrams at di erent time points during phugoid period (B) 97
4.9 Energy rate and distribution rate errors for optimal elevator-throttle co-
ordination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.10 Block diagram of the lateral state equations . . . . . . . . . . . . . . . . 102
4.11 RHP-zero due to backside operation . . . . . . . . . . . . . . . . . . . . 112
A.1 Trim curves for di erent CoG locations . . . . . . . . . . . . . . . . . . 122
A.2 Trim curves for di erent masses . . . . . . . . . . . . . . . . . . . . . . . 122
B.1 Simulink block diagram of the RCAM simulation environment . . . . . . 123
B.2 Simulink block diagram of the nonlinear RCAM model . . . . . . . . . . 124
B.3 Simulink block diagram of the RCAM actuator models . . . . . . . . . . 125
B.4 Simulink block diagram of the RCAM elevator model . . . . . . . . . . . 125
B.5 Simulink block diagram of the RCAM engine model . . . . . . . . . . . 125
B.6 Simulink block diagram of the RCAM wind models . . . . . . . . . . . . 126
B.7 Simulink block diagram of the RCAM windshear model . . . . . . . . . 126
B.8 Corrections for trimming in constant wind eld . . . . . . . . . . . . . . 128
B.9 Results linear/nonlinear simulations . . . . . . . . . . . . . . . . . . . . 131
E.1 Bode magnitude plots for varying airspeed . . . . . . . . . . . . . . . . . 138
E.2 Bode magnitude plots for varying center of gravity location . . . . . . . 138
E.3 Bode magnitude plots for varying aircraft mass . . . . . . . . . . . . . . 139
E.4 Bode magnitude plots for varying airspeed . . . . . . . . . . . . . . . . . 139
E.5 Bode magnitude plots for varying center of gravity location . . . . . . . 140
E.6 Bode magnitude plots for varying aircraft mass . . . . . . . . . . . . . . 140
E.7 Bode magnitude plots for disturbance inputs . . . . . . . . . . . . . . . 141
F.1 Aircraft response to aileron step . . . . . . . . . . . . . . . . . . . . . . 143
F.2 Aircraft response to tailplane step . . . . . . . . . . . . . . . . . . . . . 144
Version: 1
Date: April 4, 1997 x
GARTEUR/TP-088-27

F.3 Aircraft response to rudder step . . . . . . . . . . . . . . . . . . . . . . . 145


F.4 Aircraft response to throttle step . . . . . . . . . . . . . . . . . . . . . . 146
F.5 Aircraft response to lon. wind step . . . . . . . . . . . . . . . . . . . . . 147
F.6 Aircraft response to lat. wind step . . . . . . . . . . . . . . . . . . . . . 148
F.7 Aircraft response to vert. wind step . . . . . . . . . . . . . . . . . . . . 149
F.8 Aircraft response to right engine failure . . . . . . . . . . . . . . . . . . 150
F.9 Aircraft response to right engine failure . . . . . . . . . . . . . . . . . . 151
F.10 Aircraft response to windshear . . . . . . . . . . . . . . . . . . . . . . . 152
Version: 1
xi Date: April 4, 1997
GARTEUR/TP-088-27

List of tables
2.1 Aircraft geometry and inertial parameters . . . . . . . . . . . . . . . . . 12
2.2 Numerical values of the RCAM aerodynamic model parameters . . . . . 20
2.3 Numerical values control saturations . . . . . . . . . . . . . . . . . . . . 21
2.4 States de nitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.5 Inputs de nitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.6 Outputs de nitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.1 Characteristics of the longitudunal and the lateral modes . . . . . . . . 83
4.2 Initial and steady state gains for step inputs . . . . . . . . . . . . . . . . 108
C.1 Trim settings of the controls . . . . . . . . . . . . . . . . . . . . . . . . . 132
C.2 Flight data for trim condition . . . . . . . . . . . . . . . . . . . . . . . . 132
C.3 RCAM state space model for nominal trim condition . . . . . . . . . . . 133
C.4 Longitudinal linear model, with transformed state vector . . . . . . . . . 134
C.5 Lateral linear model, with transformed state vector . . . . . . . . . . . . 134
E.1 Characteristics of lon. modes, speed variations . . . . . . . . . . . . . . 136
E.2 Characteristics of the lateral modes, speed variations . . . . . . . . . . . 136
E.3 Characteristics of the lon. modes, center of gravity shifts . . . . . . . . . 137
E.4 Characteristics of the lateral modes, center of gravity shifts . . . . . . . 137
E.5 Characteristics of the lon. modes, mass variations . . . . . . . . . . . . . 137
E.6 Characteristics of the lateral modes, mass variations . . . . . . . . . . . 137
Version: 1
Date: April 4, 1997 xii
GARTEUR/TP-088-27

List of symbols and abbreviations


Symbols
The used symbols are in accordance with the nomenclature de ned in the Communication
Handbook [1]. Lots of derivative symbols have been used; their meaning will appear from
the context.
Additional de nitions:
Aircraft related quantities
A wing aspect ratio {
iT angle thrust axis w.r.t. xB -axis rad
 wing sweep rad
 taper ratio {
Air data quantities
Vtrim airspeed in trim condition m=s
Energy related quantities
D speci c energy distribution m
E energy J
Es speci c energy E=W m
Etot total speci c energy m
Pa available power W
Pr required power W
T total thrust level N
E energy ightpath angle rad
gl angle of glide (= CD =CL ) rad
Variables, ight mechanical quantities
E vector with Euler angles rad
R position vector CoG in inertial frame rad
D0 zero-lift drag N
Di induced drag N
k constant for parabolic approx. of lift-drag polar {
kq correction factor lift distribution due to q {
lsens accelerometer xM -position m
m120 nominal mass (m = 120000 kg) kg
RA total aerodynamic force vector N
rf turning radius m
V0 speed at linearization point m=s
Vground ground speed m=s
0 zero-lift angle of attack rad
t angle of attack stabilizer rad
Version: 1
xiii Date: April 4, 1997
GARTEUR/TP-088-27

Derivatives of variables
D_ speci c energy distribution rate m=s
E_s speci c energy rate m=s
E_ tot total speci c energy rate m=s
Control surface de ections
f ap position rad
Wind related quantities
B b {
2Lg
Lg scale of turbulence m
ug gust speed in x-direction m=s
vg gust speed in y-direction m=s
wg gust speed in z -direction m=s
uwx spacial der. wind speed in x-direction s 1
vwx spacial der. wind speed in y-direction s 1
wwx spacial der. wind speed in z -direction s 1
g angle of attack due to gust rad
w angle of attack due to wind rad
g sideslip angle due to gust rad
w sideslip angle due to wind rad
misc. symbols
H (s) transfer function {
!0 eigen frequency rad=s
Aerodynamic coecients
CLwb dCLwb =d {
CLt dCLt =d t {
CR total aerodynamic force coecient {
CW weight normalized with qS {
Accelerations and load factors
ax measured acceleration in xF -direction m=s2
az measured acceleration in zF -direction m=s2
nx load factor, nx = (T D)=W {
nz load factor, nz = L=W {
Subscripts
a w.r.t. airmass
ac aerodynamic center
cg center of gravity
f w.r.t. ightpath
g gust, turbulence
B relative to body axes
Version: 1
Date: April 4, 1997 xiv
GARTEUR/TP-088-27

E relative to earth axes


F relative to ightpath axes
S relative to stability axes
ss steady state
V relative to vehicle carried vertical axes
W relative to wind axes

Abbreviations
ANDECS ANalysis and DEsign of Controlled Systems
ALN Alenia Aeronautica
AVRO Avro International Aerospace
BAeDD MATRA - British Aerospace Defence Dynamics
BAe-MA British Aerospace Military Aircraft
CC Classical Control
CCL Cambridge Control
CERT Centre d'Etudes et de Recherches de Toulouse
CIRA Centro Italiano Ricerche Aerospaziali
CUN Cran eld University
DASA Daimler-Benz Aerospace Airbus
DE Germany
DLR Deutsche Forschungsanstalt fur Luft- und Raumfahrt
DRA Defence Research Agency
DUT-AE Delft University of Technology, Department of
Aerospace Engineering
DUT-EE Delft University of Technology, Department of
Electrical Engineering
EA Eigenstructure Assignment
ES Spain
FFA The Aeronautical Research Institute of Sweden
FL Fuzzy Logic control
FMAG Flight Mechanics Action Group
FM-GoR Flight Mechanics, Systems and Integration Group of Responsables
FMV Defense Material Administration
FR France
GARTEUR Group for Aeronautical Research and Technology in EURope
HI H1-Synthesis
INTA Instituto Nacional de Tecnica Aerospacial
IT Italy
LAAS Laboratoire d'Analyse et d'Architecture des Systemes
Version: 1
xv Date: April 4, 1997
GARTEUR/TP-088-27

LiTH Linkoping University


LUT Loughborough University
LY Lyapunov technique
MF Model Following control
MIMO Multi-Input Multi-Output
MM Modal Multi-Model synthesis
MO Multi Model Multi Objective Optimization
MS -Synthesis
NL The Netherlands
NLR National Aerospace Laboratory
ONERA Oce National d'Etudes et de Recherches Aerospatiales
SE Sweden
SISO Single-Input Single-Output
SMA Saab AB, Saab Military Aircraft
UCAM University of Cambridge
UCM Universidad Complutense Madrid
UK United Kingdom
ULES University of Leicester
UNAP Universita degli Studi di Napoli Federico II
UNED Universidad Nacional de Educacion a Distancia
WSPO Worst-Stability Parameter search by numerical Optimization

a.c. aerodynamic center


accel. acceleration
c.g. center of gravity
clp. closed loop
coe . coecient
der. derivative
dir. direction
displ. displacement
ext. external
i if and only if
lat. lateral
lon. longitudinal
olp. open loop
ref. reference
rhp right half plane
transf. fnc. transfer function
vert. vertical
w.r.t. with respect to
Version: 1
Date: April 4, 1997 xvi
GARTEUR/TP-088-27
Version: 1
1 Date: April 4, 1997
GARTEUR/TP-088-27

1 Introduction
This document contains a description and an analysis of the Research Civil Aircraft Model
(RCAM). This nonlinear model is based on a large civil transport aircraft in landing
con guration. It is used for controller design and simulation in one of the two Robust
Flight Control design challenges, prepared by GARTEUR Action Group FM(AG08), see
[3, 4]. The idea behind these design challenges is to invite the European industry, research
establishments and universities involved with ight control to apply their modern and
classical ight control techniques to realistic ight control problems and to demonstrate in
this way the capabilities, applicability and limitations of these methods. It is also intended
that these problems will raise the awareness and con dence of the European aeronautical
industry in the use of robust control techniques.
The design challenges are based on a military and a civil ight control problem. RCAM
is the basis for the civil variant. An extensive set of design objectives and additional
requirements can be found in the manual prepared by GARTEUR, see [2].
This document has two main objectives:
 detailed documentation of the aircraft modeling and disturbance modeling (winds-
hear, gust) for the RCAM, including implementation in Matlab/Simulink;
 open loop analysis of the RCAM as a basis for the design of a ight controller. This
analysis includes performance limitations, dynamic characteristics and co-ordination
of controls in speci c situations, like turns, windshear, engine failure etc.
The main task of the controller is guidance of the aircraft along trajectories with turns,
climbs and descents, while for example also wind disturbances or an engine failure may
occur. Along these trajectories, physical aircraft performance (e.g. turning performance,
maximum climb angles etc.) is a limiting factor, that of course can not be overcome by the
implementation of a "robust" controller. For this reason aircraft performance is analysed,
mainly using point-mass ight mechanics.
From the nonlinear model, linearizations can be determined at di erent operating points.
One operating point is de ned as nominal, where standard tools from linear system theory
are applied, like bode plots, step responses, and eigenvalues to analyse dynamic charac-
teristics of the RCAM. Furthermore, the e ects of airspeed variations, center of gravity
shifts and mass variations are considered.
Aircraft control is basically a multi-input multi-output problem. Therefore an analysis is
carried out to derive some a priori knowledge about proper co-ordination of controls, like
for example elevator/throttle co-ordination in climbing ight. This study is based on the
equations of motion, point-mass approximations, and steady state analysis.
Although in many modern multivariable design techniques control laws are synthesized
automatically, the results can be used as a reference in the assessment of the controller.
Version: 1
Date: April 4, 1997 2
GARTEUR/TP-088-27

This document may serve as a "database" with RCAM data, concerning the modeling,
dynamic characteristics and performance, as far as these can be derived from the model.
Therefore, most of the results are presented in gures or in tabular form.
The outline is as follows: in the chapter 2 the modeling and implementation of the RCAM
is described. In the third chapter the analysis work based on point-mass and steady state
ight mechanics is described, including Korhammer diagrams, energy principles and steady
state aircraft performance. In chapter 4 the analysis work based on linearized models is
described.
Version: 1
3 Date: April 4, 1997
GARTEUR/TP-088-27

2 Description of the aircraft simulation model


This chapter describes the modeling of the Research Civil Aircraft Model (RCAM) and
the implementation in the Matlab/Simulink-environment. The model is based on a large
transport aircraft, propelled by two turbofan engines. This aircraft is a modi ed conven-
tional passenger aircraft: the fuselage has a modi ed shape and a very large diameter. At
the ends of the horizontal tailplane additional vertical surfaces have been tted.
The nonlinear rigid body equations of motion will be discussed in the rst section. In
the second section the actuator models will be described and the wind models will be
discussed in section 3. All sensors are assumed to be perfect. Therefore, no sensor models
are available for the RCAM. Finally, the software for RCAM will be described in section 5.

2.1 The nonlinear aircraft model


In this section the rigid body equations of motion will be described. The derivation
presented here is based on the book Flugregelung, [6]. All speci c aircraft data, like the
aircraft geometry and the aerodynamic model have been obtained from [2].

2.1.1 Assumptions

In order to make the nonlinear equations of motion not unnecessarily complicated, some
assumptions for the generic nonlinear model will be made rst. The aerodynamic model
and the engine model, which are speci c for the RCAM, are based on more detailed
assumptions.
 The earth is in rest and is considered to be at. This implies that any reference frame
xed with respect to the earth (position and orientation), is an inertial reference
system.
 The equations are valid for a rigid body, which means that all elastic degrees of
freedom of any part of the aircraft are neglected. This assumption holds as long
as the eigenfrequencies of the elastic modes are much higher than those of the rigid
body motions.
 The mass of the aircraft is constant. This is a reasonable assumption since the time
constant of fuel consumption is much higher than the time constants of the rigid
body modes. Simulation times will be in the order of minutes.
 The aircraft is symmetric with respect to its xB -zB {plane. For this reason the Ixy
and Iyz in the inertia tensor disappear.
 Interactions between the engines and the other parts of the airframe are not taken
into account. This means that cross coupling induced by the rotation of the engines
Version: 1
Date: April 4, 1997 4
GARTEUR/TP-088-27

is neglected. Furthermore, interactions between the exhaust stream of the engines


and the air ow around the airframe are neglected.
 All forces acting on the di erent parts of the aircraft are simply taken together and
considered with respect to the center of gravity. This assumption is valid as long as
the aerodynamics can be considered to be quasi-stationary, see the next assumption.
 The generation of aerodynamic forces by the air ow around the wings, the tail
surfaces and the fuselage, is assumed to be quasi-stationary. This assumption is
valid as long as the wavelengths of variations in the air ow are at least eight times
the mean aerodynamic chord length.

The forces and moments acting on the aircraft can be divided into three groups:
 aerodynamic forces and moments,
 forces and moments induced by the power plant,
 gravity forces.
The rst two are complex, multidimensional nonlinear functions and they are usually ap-
proximated by polynomial expansions. These forces and moments arise in di erent axes
systems. Although these systems are described in the Communication Handbook, [1], they
will also be discussed in this section, so that some important ight mechanical quantities
can be introduced simultaneously.

2.1.2 Axis systems


The earth- xed reference frame is the most important, since it is the inertial frame required
for application of Newton's law. The xE axis is directed north, the yE axis is directed
east and the zE axis is in the direction of the earth, as shown in gure 2.1. This reference
frame is used for purpose of navigation, since it de nes the absolute position of the aircraft
with respect to a xed point on the earth. All other reference systems considered here are
vehicle-carried, which means that their origin is always at a xed point of the aircraft. If
the earth- xed frame is translated to the aircraft center of gravity (over its position vector
R = [x; y; z]TE ), the vehicle-carried vertical frame is obtained, see gure 2.1. If the origin
of the inertial system is chosen on the earth surface, the third component of R is equal to
minus the altitude of the aircraft:
z= h
The vehicle-carried vertical frame has the same orientation as the earth- xed system. The
components of the velocity vector relative to the earth along the xV , yV and zV axes are
uV , vV and wV respectively.
Version: 1
5 Date: April 4, 1997
GARTEUR/TP-088-27

xB

xV
θ k1
ψ
yV
CoG
ψ
xV, yV φ k2

xE R xB , z v
yB
φ
yE
yB , z B
zB
θ
k3

zV
zE

Fig.2.1 The body- xed, vehicle carried-vertical and the earth- xed axes systems
The second vehicle-carried system, the body- xed axes system, is obtained through succes-
sive rotations of the vertical vehicle-carried system with heading angle , pitch angle  and
roll angle , see gure 2.1. These angles form the vector of Euler angles E = [ ; ; ]T ,
which describes the attitude of the aircraft. The xB -axis is directed towards the nose of
the aircraft, the yB -axis towards the right wing and the zB -axis towards the bottom of
the aircraft. The components of the velocity vector along these axes are uB , vB and wB ,
the components of the vector of angular rates around the axes,
B , are respectively: the
roll rate p, the pitch rate q and the yaw rate r.
The x-axis of the ightpath axes system is in the direction of the velocity vector V. The
system is obtained by rotating the vehicle-carried vertical system about the zV -axis over
the ight path azimuth angle () and about the yF -axis over the ightpath (elevation)
angle ( ) respectively, see gure 2.2.
Both angles can be calculated using the velocity components in the vehicle-carried vertical
frame from respectively:

tan = q 2 wV 2 (2.1)
uV + vV
and:
tan  = uvV (2.2)
V
Version: 1
Date: April 4, 1997 6
GARTEUR/TP-088-27

xF

xV χ γ

V yV

χ
CoG
yF

zF
zV

Fig.2.2 The vehicle-carried vertical and the ightpath axes systems


The body- xed frame is obtained from the ightpath axes system through successive
rotations over the ightpath sideslip angle ( f ), the ightpath angle of attack ( f ) and
the ightpath bank angle (f ), see gure 2.3
Aerodynamic forces and moments arise in the wind or aerodynamic axes system. The
xW -axis is in the direction of the aircraft velocity relative to the air ow, the airspeed
(Va = V VW , where VW is the windspeed vector). The wind axes system is obtained
by successive rotations of the body axes system over (minus) the sideslip angle (- ) and
the angle of attack ( ), see gure 2.4.
The components of Va along the body axes are uaB , vaB and waB respectively. The angle
of attack and the sideslip angle are de ned as:
tan = wu aB (2.3)
aB
and
sin = vVaB (2.4)
A q
with VA = jVa j = u2aB + va2B + wa2B .
Di erentiating equations 2.3 and 2.4 the following expressions for the time derivatives of
and are found:
_ = aaz uua2 + awa2x wa = aaz cosV cosa ax sin (2.5)
a a A
2 2
_ = aay ua + wa 2 p va (aax ua + aaz wa ) (2.6)
VA u2a + wa2
=
aay cos sin (aax cos + aaz sin ) (2.7)
VA
Version: 1
7 Date: April 4, 1997
GARTEUR/TP-088-27

xB

αf
βf xF
V

CoG
βf
yF
µf
yB

µf
αf
zB
zF

Fig.2.3 The body- xed and the ightpath axes systems

xB

xS
α xW
β

VA
CoG Y
D
β
yB , yS
yW

zB
zW , z S

Fig.2.4 The wind, stability and body- xed axes systems


Version: 1
Date: April 4, 1997 8
GARTEUR/TP-088-27

Where aax , aay and aaz are the time derivatives of the airspeed components along the body
axes (e.g. aax = dudta ). The aerodynamic forces in the model to be derived are expressed
with respect to the stability, or experimental axes system. The term \experimental" comes
from the fact that this system is also used in wind tunnel experiments. The system should
not be confused with the "quasi-stationary" stability axes, used in many textbooks. In
stationary ight the systems are identical, but when the air ow becomes unsteady, the
orientation of the stability axes in these books is considered to be xed, while the stability
axes used here vary with the direction of the airspeed, Va . Although it would be more
convenient to use the term "experimental", the nomenclare from [1] will be adopted here.
The stability axes system is obtained from the body axes system in almost the same way
as the wind axes; the body axes system however, is only rotated about the yB -axis over
the angle of attack ( ), see gure 2.4.
Parallel to the body axes, the wind reference system can obtained from the vehicle carried
vertical system by successive rotations over the angles a , a and a respectively, as can
be seen in gure 2.5. The azimuth angle a can be calculated from:
xW

xV χa γa

VA yV

χa
CoG

µa

yW

µa

γa

zW
zV

Fig.2.5 Transformation from the vehicle-carried system to the wind axes system

 va 
a = arctan u V (2.8)
aV
The elevation angle is obtained as follows:
0 1
a = arctan @ q 2 waV A (2.9)
uaV + va2V
For the small sideslip angle and small angle of attack the bank angle a is approxi-
Version: 1
9 Date: April 4, 1997
GARTEUR/TP-088-27

mately equal to the roll angle:


a   (2.10)
Finally, three important angles for the wind in uence are de ned by transformation from
the wind axes to the ightpath axes by successive rotations over the wind sideslip angle
( crab angle) w , the wind angle of attack w and the wind bank angle w . The angles
can be seen in gure 2.6.
The components of vectors relative to one axes system can be expressed in co-ordinates
xB
xW
βw
VA
VW αw
xF

V
CoG
yF
yB
β
µw yW
βw

zF
zV

Fig.2.6 The de nition of wind angles

with respect to another system using transformation matrices. The transformation matrix
from axes system M to axes system N is expressed using the notation RNM , where
RNM = RMN T .
Usually one system is obtained from the other via three successive rotations, like the body-
xed system is obtained from the vehicle-carried vertical system via the angles ,  and
 respectively, see gure 2.1. In a more general form, the transformation matrix resulting
from rotations over 1 , 2 and 3 respectively, can be written as:
RNM = R
2 3(3 ) R2(2 ) R1(1 ) 3 2 32 (2.11)
3
6 1 0 0 7 6 cos 2 0 sin 2 7 6 cos 1 sin 1 0 7
= 4 0 cos 3 sin 3 75 64 0 1
6 0 75 64 sin 1 cos 1 0 75
0 sin 3 cos 3 sin 2 0 cos 2 0 0 1
For the vehicle-carried systems, the most important transformation matrices are:
Version: 1
Date: April 4, 1997 10
GARTEUR/TP-088-27

 wind to body axes ( g 2.6):


RBW = R2 3 (0) R2( ) R1( )3 2 3 (2.12)
6 cos 0 sin 76 cos sin 0 7
= 64 0 1 0 75 64 sin cos 0 75
sin 0 cos 0 0 1
 stability to body axes ( g 2.4):
2 3
6 cos 0 sin 7
RBS 6
= R3 (0) R2 ( ) R1 (0) = 4 0 1 0 75 (2.13)
sin 0 cos
 vehicle-carried vertical axes to body axes ( g 2.1):
RBV = R
2 3 () R2 () R3( ) 3 2 32 3(2.14)
6 1 0 0 7 6 cos  0 sin  7 6 cos sin 07
= 4 0 cos  sin  75 64 0 1
6 0 75 64 sin cos 0 75
0 sin  cos  sin  0 cos  0 0 1
 vehicle-carried vertical axes to ightpath axes ( g 2.2):
RFV = 2R3 (0) R2 ( ) R1 () 3 2 3 (2.15)
6 cos 0 sin 7 6 cos  sin  0 77
6
= 4 0 1 0 75 64 sin  cos  0 5
sin 0 cos 0 0 1

2.1.3 The generic state equations of the aircraft motions


For description of the states of the aircraft, twelve variables have already been introduced
with the axes systems: x, y, and z for position in the inertial system, ,  and  for
attitude with respect to the inertial system, uB , vB and wB , the velocity components
relative to the earth and p, q and r, the angular rates. The state variables u, v, w, p, q, r
are along the body axes, because the equations of motion will be derived with respect to
this reference frame. The reason for choosing this system is that the moments of inertia
remain constant when the aircraft position and attitude vary. To solve the equations of
motion with twelve states, twelve (scalar) equations are needed. Using Newton's law, six
equations are obtained. The remaining equations come from kinematic relationships.

Force equations
The force and moment equations for a rigid body in an axes system that is xed to the
Version: 1
11 Date: April 4, 1997
GARTEUR/TP-088-27

center of gravity of this body, were rst derived by Euler. Using the second assumption
in section 2.1.1, the force equations in the body- xed system become:
2 3 2 3 2 3 2 3
 dV  6 u_ 77 = 1 R 66 FxA 77 + 1 66 FxT 77 +R 66 0 77 g
V (2.16)
6
= 4 v_ 5 m BS 4 FyA 5 m 4 FyT 5 BV 4 0 5 B B
dt B
w_ B FzA S FzT
1 V B
(vector subscripts indicate axes system) The term induced by the coupling between velocity
and rotation can be expanded as follows:
2 3 2 3 2 3
6 p7 6u 77 = 66 qw rv 7

B  VB = 4 q 75  64 v
6 5 4 ru pw 75 (2.17)
r B w B pv qu B
Expanding the gravity term gives:
2 3 2 3
6 0 7 6 g sin  77
RBV 64 0 75 g = 64 g sin  cos  5 (2.18)
1 V g cos  cos 
where g = 9:81 m=s.

Moment equations
The moment equations can be written in a similar general form as the force equations:
2 3 02 3 2 3 1
 d
 6 p_ 7 LA 77 + 66 LT
= 64 q_ 75 = I 1B 6
B@64 MA 77
B  I
B CCA
dt B 5 4 MT 5 (2.19)
r_ B NA B NT B
The inertia tensor I contains the moments of inertia with respect to the body- xed axes
system. With Ixy = 0 and Iyz = 0 (symmetry assumed) the tensor can be written as:
2 3
6 Ix 0 Ixz 7
I = 64 0 Iy 0 75 (2.20)
Ixz 0 Iz
The values of the principal moments of inertia IX , IY and IZ can be found in table 2.1.
These are the moments of inertia with respect to the principal axes, in which the cross
products are zero. The principal axes are obtained by rotating the body axes about the
yB -axis over the angle p, see also for example [10]. The inertia tensor in body axes is
calculated using standard formulas:
2 (IX +IZ ) 3
6 2 + (IX 2 IZ ) cos( 2p ) 0 (IX IZ ) sin(
2 2p ) 77
I = 64 0 IY 0 5 (2.21)
(IX IZ ) sin( 2 ) 0 (IX +IZ ) + (IX IZ ) cos( 2p )
2 p 2 2
Version: 1
Date: April 4, 1997 12
GARTEUR/TP-088-27

Symbol Name Default Min. value Max. value Unit


Mass Parameters
m aircraft total mass 120 000 100 000 150 000 kg
Inertia Parameters
IX moment of inertia w.r.t. body principal x- 40.0 { { m2
m
axis per unit mass
IY moment of inertia w.r.t. body principal y- 64.0 { { m2
m
axis per unit mass
IZ moment of inertia w.r.t. body principal z- 100.0 { { m2
m
axis per unit mass
p rotation angle from body axes to body -2.0 { { deg
principal axes
Engine Parameters
iT thrust vector angle w.r.t. xB -axis 0.0 { { deg
xAPT 1 x pos. of application point of thrust of N/A { { m
engine 1 in FM
yAPT 1 y pos. of application point of thrust of en- -7.94 { { m
gine 1 in FM
zAPT 1 z pos. of application point of thrust of en- -1.9 { { m
gine 1 in FM
xAPT 2 x pos. of application point of thrust of N/A { { m
engine 2 in FM
yAPT 2 y pos. of application point of thrust of en- 7.94 { { m
gine 2 in FM
zAPT 2 z pos. of application point of thrust of en- -1.9 { { m
gine 2 in FM
Geometric parameters
c mean aerodynamic chord 6.6 { { m
b wing span 45.0 { { m
A aspect ratio 7.73 { { {
 taper ratio 0.29 { { {
 wing sweep 28.0 { { deg
lt distance between AC of the wing-body 24.8 { { m
(ACwb ) and AC of the tail (ACt )
S wing planform area 260.0 { { m2
St tail unit planform area 64.0 { { m2
xcg x-position of the CoG 0.23 c 0.15 c 0.31 c m
ycg y-position of the CoG 0.0 c -0.03 c 0.03 c m
zcg z-position of the CoG 0.0 c 0.0 c 0.21 c m
xacwb x-position of ACwb in FM 0.12 c { { m
Table 2.1 Aircraft geometry and inertial parameters
Version: 1
13 Date: April 4, 1997
GARTEUR/TP-088-27

Kinematic relations
The remaining six equations needed to solve the equations of motion for twelve states are
obtained through kinematic relations:
2 3 2 3
 dR  6 x_ 7 6 u 77
6 7 6
= 4 y_ 5 = RV B 4 v 5 (2.22)
dt E
z_ E w B
with RV B = T .
RBV
For the Euler angular rates holds:
2 3
 dE  6 p7
dt = RB 64 q 75 (2.23)
r B
with the transformation matrix:
2 3
6 1 sin  tan  cos  tan  77
RB = 64 0 cos  sin  5 (2.24)
0 sin = cos  cos = cos 

2.1.4 The aerodynamic model of the RCAM


The aerodynamic forces and moments are nonlinear functions of the motion of the aircraft
relative to the air ow for a given aircraft con guration (e.g. ap setting f , gear up
or down) and a given ight condition. The forces and moments from the aerodynamic
model (in the stability axes system) are expressed using nondimensional coecients Ci .
All numerical values of the model parameters that are mentioned in this section, can be
found in table 2.2.
The aerodynamic forces are calculated as follows:
2 3 2 3 2 3
66 FxA 77 = 66 D 77 =  V 2 S 66 C D ( M; ; ; ;
_ q;  E ::: ) 7
4 FyA 5 4 Y 5 2 A 4 Y C ( M; ; ; _
; p; r; A R :::) 7
; 5 (2.25)
FzA S L S CL (M; ; ; ;_ q; E :::) S
The aerodynamic model of the RCAM is relatively simple. Most of the coecients are
constant; no look-up tables have been implemented. Another simpli cation is that no
atmospheric model is used; the atmospheric conditions at h = 0 m are taken:
0 = 1:225 kg=m3
p0 = 101325:0 N=m2
T0 = 288:15 K
The airspeed, Va is calculated by subtracting wind components from the inertial speed,
V. The model has two sets of wind inputs: one set for wind components generated with
Version: 1
Date: April 4, 1997 14
GARTEUR/TP-088-27

respect to the body axes ([WXB WY B WZB ]T ) and another set for wind generated in FE
([WXE WY E WZE ]T ). The gust models used in the simulation software generate gusts
w.r.t. the body axes while the constant wind eld and windshear are generated in FE .
The airspeed components along the body axes can be calculated from:
2 3 2 3 2 3
6 u 7 6 WXB 77 + R 66 WXE 77
6 7 6
VaB = 4 v 5 + 4 WY B 5 BV 4 WY E 5 (2.26)
w B WZB WZE

The lift coecient is calculated from (see gure 2.7):


S t
 Vt 2
CL = CLwb + CLt S V (2.27)
For the RCAM model the contribution of the stabilizer is only corrected for its reference
surface (St=S ). The in uence of loss of dynamic pressure at the horizontal tailplane is
2
neglected, VVt = 1.
The lift coecient of the wing-body combination, CLwb , is given by:

CL
wb

xB
CoG CL
CD t
α
Va ε δT
AC
AC t

αt

q=0 lt

Fig.2.7 De nitions of lift and drag


8
< C ( )  180
14:5 rad
CLwb = : Lwb 3 0 2 =
14:5 rad (2.28)
768:5 + 609:2 155:2 + 15:2 > 180 =
where 0 is the angle of attack for which CLwb = 0 and CLwb  is the derivative of CLwb
with respect to . For angles of attack up to = 14:5 deg the relation between the lift
coecient and the angle of attack is linear. In order to implement a stall in the model, at
higher angles of attack the CL curve has a modi ed shape. The lift coecient has a
maximum at = 18 deg, so that at this point the lift coecient is CL = 2:75.
In gure 2.8 a plot of the CL curve can be found. The contribution of the stabilizer

Unless stated di erently, with CLwb is always meant the derivative in the linear range of the Cwb
relation ( < 14:5 deg).
Version: 1
15 Date: April 4, 1997
GARTEUR/TP-088-27

2.5 f = 32:5 deg, gear down


2 CL

CL , CD
1.5

0.5 CD
0
−10 −5 0 5 10 15 20

(deg)
Fig.2.8 Curves of CL and CD vs.
to the lift coecient, CLt can be calculated from:
CLt = CLt t (2.29)
where CLt is the derivative of CLt with respect to t , the angle of attack of the horizontal
tailplane:
t = " + T + kq Vqlt (2.30)
A
with the downwash angle:
d" ( )
" = d (2.31)
0
For pitch control of the RCAM, the whole tailplane is rotated (T ); this de ection directly
a ects t . Although the original aircraft on which the RCAM is based, is certainly con-
trolled using the elevator, no aerodynamic data of this device are availabley.
The last term of equation 2.30 is introduced by the rotation of the aircraft about the yB -
axis, which makes that at the stabilizer an extra vertical airspeed component is felt, equal
to q  lt . The factor lt is the distance between the aerodynamic centers of the wing-body
combination and the stabilizer. This is equivalent to a contribution to the angle of attack
of q lt =VA . The constant kq is a correction factor to take into account that during rotation
a similar change in is felt along the whole aircraft, a ecting the total lift and the lift
distribution over the aircraft.

In the literature the contributions of elevator/tailplane de ection, pitch rate, angle of


attack etc. to the total lift coecient are usually modeled explicitly. The derivative of CL
w.r.t. for the complete aircraft can be calculated from:
 d"  St
CL = CLwb + CLt 1 d S (2.32)
y
In this thesis we will usually speak of elevator (E ). Only for speci c RCAM analysis tailplane (T ) is
used.
Version: 1
Date: April 4, 1997 16
GARTEUR/TP-088-27

The in uence of the tailplane de ection on the lift coecient (CLT ) can for the RCAM
model be written as:
CLT = CLt St  Vt 2 (2.33)
S V
The contribution to the lift coecient due to a pitch rate is modeled with:
 2
C qc = k C St Vt q lt
Lq V q Lt S V VA (2.34)
A
Thus:
 2
CLq = kq CLt SStclt VVt (2.35)

The drag coecient CD is calculated from:


CD = CD0 + k (CLwb ( 0 ) CL0 )2 (2.36)
Up to CL = 2:5 this equation is equivalent to a parabolic lift{drag relation:
CD = CD0 + k (CLwb CL0 )2 (2.37)
k is a constant and CL0 is the lift coecient at the top of the parabola, where CD = CD0 .
The lift-drag polar is plotted in gure 2.9 and the CD plot can be found in gure 2.8.
3

2.5

2
CL

1.5

0.5
f = 32:5 deg, gear down
0
0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6

CD
Fig.2.9 The lift-drag polar

The contribution of the stabilizer to the drag coecient is neglected. However, if the
aircraft is trimmed, the stabilizer has a secondary e ect on the total aircraft drag. If the
equilibrium state requires a negative tailplane de ection, the device will have a negative
contribution to the total lift. This has to be compensated by the wing, so that a higher
Version: 1
17 Date: April 4, 1997
GARTEUR/TP-088-27

angle of attack is required. However, this also means a higher drag coecient, as can
easily be seen from equation 2.37.
In practice this extra drag is very important, because over a longer period of time it might
cost a considerable amount of fuel. This is one of the reasons why the application of
relaxed stability is considered for civil aircraft. By moving the location of the center of
gravity to a more aft position, the angle of incidence of the stabilizer will become positive,
or at least less negative, resulting in less drag. On the other hand, it is also possible
to decrease the size of tail plane in the design of an aircraft (see [11]). The decreased
longitudinal stability needs to be compensated by application of active controls.
The lateral component of the aerodynamic force vector, CY , is calculated from:
CY = CY + CYR R (2.38)

With equation 2.25 it is now possible to calculate the aerodynamic forces.


The aerodynamic moments are modeled in the body axes as follows:
2 3 2 3
66 LA 77 =  V 2S 66 Cl 77 +
4 MA 5 2 A 4 Cm 5 (2.39)
NA C
2 xncg xac 3 0 2 31
CD
 V 2 S c 66 c ycg c 77  BBR 66 7C
A 4
2 c
zcg
5 @ BS 4 CY 75 CA
c B CL S B
where S is de ned asz :
2 3
6 Sb 0 0 7
S = 64 0 S c 0 75
0 0 Sb
The last term in equation 2.39 is the contribution of the aerodynamic forces, see gure 2.10.

Expanding the rst of the righthand terms of equation 2.39 gives:


2 3 2 3 2 32 3
66 Cl 77 = 66 0 77 + 66 0 Cl 0 76 7
4 Cm 5 4 Cm0 5 4 Cm 0 0 75 64 75 + (2.40)
Cn 0 0 C Cn
2 3 2n pb 3
66 Clp 0 Clr 7 6 2VA 77 +
4 0 Cmq 0 75 64 VqAc 5
Cnp 0 Cnr rb
2VA
z
Lateral moment coecients are normalised with 21 VA2 b. However, in the software c in stead of b is
used.
Version: 1
Date: April 4, 1997 18
GARTEUR/TP-088-27

zM
xn fix
x cg
x ac wb lt
xB
CoG NP AC t

z cg
Ref. Point AC wb xM

z APT1,2

zB

Fig.2.10 Location of moment points


2 32 3
66 ClA 0 ClR 77 66 A 77
4 0 CmT 0 5 4 T 5
CnA
0 CnR R
A, T and R are the de ections of the aileron, tailplane and rudder. The numerical
values of the moment derivatives can be found in table 2.2, where also the coecients of
the aerodynamic forces have been summarized.
The aerodynamic derivatives for the moment about the lateral axis have been derived
according to [10]. The aerodynamic moment with respect to the aerodynamic center can
be written as:
 Vt 2 St lt
Cm = Cmac CLt V S c (2.41)
After substitution of equation 2.29 and 2.30 this expression can be written as:
C = C + C ( ) + C qc + C 
m m0 m 0 mq VA mT T (2.42)
with:
 Vt 2 Stlt
Cm0 = Cmac CLt 0 V S c (2.43)
 d"   Vt 2 Stlt
Cm = CLt 1 d V S c (2.44)
 2 2
Cmq = kq CLt VVt SStcl2t (2.45)
and:
 2
CmT = CLt VVt SStclt (2.46)
Version: 1
19 Date: April 4, 1997
GARTEUR/TP-088-27

These moment derivatives are mainly a ected by the center of gravity location. In ap-
pendix A the e ect of center of gravity shifts on the moment derivatives and static stability
is discussed.
2.1.5 Thrust contribution to forces and moments
The aircraft is powered by two engines. The thrust Ti of each of the engines is calculated
by:
Ti = THi m120 g; i = 1; 2 (2.47)
g = 9:81 m=s2 is the gravity acceleration, m120 = 120000 kg, a mean value for the aircraft
mass. The factor m120 g is only a scale factor in order to get reasonable values for throttle
settings. The thrust is aligned with the xB -axis; the components along the yB - and zB -axis
are zero. The vector of the thrust contributions to the force equations becomes:
2 3 2 3
66 FxT 77 66 T1 + T2 77
4 FyT 5 =4 0 5 (2.48)
FzT B 0 B
The contribution of the thrust of each of the engines to the moment equations can be
calculated from:
2 3 2 3
6 xcg xAPTi 77  66 Ti 77
6
MTi = 4 ycg + yAPTi 5 4 0 5 (2.49)
zcg zAPTi B 0 B
xcg , ycg , zcg , xAPTi, yAPTi and zAPTi (i = 1; 2) can be found in table 2.1.
2.2 The actuator and engine models
Control surface de ections are generated using actuators that have limited bandwidth,
limited de ections and limited de ection rates. The engine dynamics have identical lim-
itations. All control devices are modeled using rst order models. These are described
below:

The aileron actuators has a time constant of  = 0:15 s:


HA (s) = 0:151s + 1 (2.50)
The tailplane actuator also has a time constant of  = 0:15 s:
HT (s) = 0:151s + 1 (2.51)
The rudder has a time constant of  = 0:30s:
HR (s) = 0:301s + 1 (2.52)
Version: 1
Date: April 4, 1997 20
GARTEUR/TP-088-27

Aerodynamic model
h = 0 m, f = 32:5 deg
gear down
Non-coecients
k 0.07
kq 1.3
0 -0.21 rad
d" 0.25
d 2
Vt 1.0
V2
Force equations
CD0 0.13 CL0 0.45 {
{ CL 6.07 {
{ CLwb 5.5 {
{ CLt 3.1 {
{ CLq 3.73 {
{ CLT 0.76 {
{ { CY -1.6
{ { CYR 0.24
Moment equations
{ Cm0 -0.015 {
{ Cm -2.15 {
Cl -0.205 { Cn 0.147
{ { Cn -0.560
Clp -0.473 { Cnp 0.073
{ Cmq -14.0 {
Clr 0.215 { Cnr -0.495
ClA -0.088 { CnA 0.0
{ CmT -2.87 {
ClR 0.032 { CnR -0.092
Note: all coecients w.r.t. aerodyn. center,
force coe . in FS , moment coe . in FB .

Table 2.2 Numerical values of the RCAM aerodynamic model parameters


Version: 1
21 Date: April 4, 1997
GARTEUR/TP-088-27

The engine dynamics have a time constant of  = 1:5s:


HTH (s) = 1:5s1+ 1 (2.53)
Of course the maximum de ections of the control inputs are limited. These limits are
given in table 2.3.
Because only simple rst order models are used, no acceleration limits are taken into

input min. sat. max. sat. unit rate limit unit


A -25.0 25.0 deg 25 deg/s
T -25.0 10.0 deg 15 deg/s
R -30.0 30.0 deg 25 deg/s
TH 1;2 0.5 10.0 deg 1.6 deg/s

Table 2.3 Numerical values control saturations

account. The limit of 10 degrees for the throttles means that the maximum T=W -ratio is
approximately 10=57:3  2 = 0:35, since the throttle de ections are scaled with a nominal
weight, m120 g, see section 2.1.5.
In the case of an engine failure the throttle setting of the failed engine reduces to THi =
0:5 deg (i = 1; 2) with rst order system dynamics given by the transfer function:
Heng:failure(s) = 3:3s1+ 1 (2.54)
This transfer function is chosen such that the maximum rate of the decreasing throttle
setting starting from a trim position (4 5 deg) is somewhat less than 1:6 deg=s. When
a controller is used to compensate automatically by increasing the throttle position of the
working engine, this can be done at a rate just below the maximum allowed throttle rate.
2.3 The wind models
In order to make both open loop and closed loop nonlinear simulations more realistic,
the nonlinear model is equiped with inputs for wind and turbulence. In this section the
models are described that will be used to generate these wind inputs.
2.3.1 Constant wind
The rst group of three wind inputs to the aircraft model enter the equations of motion
in the inertial reference system. The windspeeds are simply subtracted from the inertial
speed, resulting in the airspeed. The constant wind needs no further modeling, only a
realistic windspeed with repect to the earth should be set in a desired direction.
Before a simulation is started where a constant wind eld is applied, it is necessary to trim
the aircraft for this wind. With the trim routine it is possible to nd equilibrium with the
Version: 1
Date: April 4, 1997 22
GARTEUR/TP-088-27

correct crab angle, ground speed etc. For controller evaluation in time simulations, wind
inputs of 10{20 m=s are applied.
2.3.2 Turbulence models
Turbulence is a stochastic process that can be de ned by velocity spectra. Commonly
used velocity spectra for turbulence modeling are the Dryden spectra. These spectra are
based on a number of assumptions; the most important are:
 Atmospheric turbulence is a random process with Gaussian distribution
 Atmospheric turbulence is a stationary process, hence the distribution of wind ve-
locities in a wind eld is considered to be independent of time
 Atmospheric turbulence is homogeneous along the ightpath
 Atmospheric turbulence is an isentropic process
For an aircraft ying at a speed V through a `frozen' turbulence eld with a spatial
frequency of
rad/m, the circular frequency of the turbulence can be calculated as:
! = V 
rad/s (2.55)
With this, the spectra of the encountered gust velocities can be described as follows:
2L
ug (!) = u2g Vug (1 + (L1 ! )2 )
ug V
L 1 + 3( L ! )2
vg (!) = v2g Vvg (1 + (L vg!V)2 )2 (2.56)
vg V
L 1 + 3(L ! )2
wg (!) = w2 g Vwg (1 + (L wg!V)2 )2
wg V
A Bode-plot of these power spectral densities can be found in gure 2.11. For this plot
is taken: Lug = Lvg = Lwg = Lg = 305 m and ug = vg = wg =  = 1:5 m=s. A
turbulence scale length (Lug , Lvg , Lwg ) can be interpreted as a measure of spacial extent
of signi cant correlation. The turbulence scale lengths and turbulence standard deviations
ug , vg , wg are dependent on altitude and atmospheric conditions. As an indicator for
the atmospheric conditions it is possible to take the wind speed at 20 ft above the ground
(W20 ). For `moderate' conditions, W20 = 15:4 m/s (30 kts) is selected. The turbulence
standard deviation wg is then given as follows:
wg = 0:1W20 (2.57)
ug and vg are assumed to be functions of wg and the altitude h. For h < 305 m
(1000 ft):
ug = vg = (0:177 +0w:0027
g
h)0:4 (2.58)
Version: 1
23 Date: April 4, 1997
GARTEUR/TP-088-27

10
2 lon (|{), lat/vert (- - - -)

1
10

0
10
ug ; wg
−1
10

−2
10

−3
10

Lg = 305 m,  = 1:5 m=s


−4
10 −2 −1 0 1 2
10 10 10 10 10

! (rad/s)
Fig.2.11 Power spectral densities of Dryden lters
and for h > 305 m (1000 ft):
ug = vg = wg (2.59)
The turbulence scale lengths Lug , Lvg and Lwg are calculated as a function of altitude: for
3 < h < 305 m:
Lug = Lvg = (0:177 + 0h:0027h)1:2 (2.60)
Lwg = h (2.61)
and for h > 305 m is taken:
Lug = Lvg = Lwg = 305 m (2.62)
With this procedure, the gust velocities ug , vg and wg are de ned in the stability reference
frame. However, as an approximation the RCAM inputs WXB , WY B and WZB are used.
To simulate turbulence, white noise is ltered through forming lters. A great advantage
of the Dryden lters is that these lters can easily be derived from equation 2.56. As
an example, the transfer function of the lter for simulating the gust velocity wg will be
considered.
Given white noise w, the spectrum of wg can be obtained as:
wg = jHwg w (!)j2 w (2.63)
Where w = 1, and Hwg w (!) is the frequency response function of the forming lter.
Therefore,
3(Lwg V! )2
w2 g LVwg (11 +
+ (Lwg V! )2 )2 = jHwg w (!)j = Hwg w (!)Hwg w ( !)
2 (2.64)
Version: 1
Date: April 4, 1997 24
GARTEUR/TP-088-27

To obtain a stable and minimum phase lter, the following frequency response function is
selected:
s p Lwg
Hwg w (!) = wg VL wg 1 + 3 V j!
(2.65)
(1 + LVwg j!)2
Replacing the variable j! by s, the following transfer function is obtained:
s p
Hwg w (s) = wg Lwg 1 + 3 LVwg s (2.66)
V (1 + LVwg s)2
The transfer function for generating vg is equivalent. The transfer function for generating
ug can be found as:
s
Hug w (s) = ug 2LVug 1Lug (2.67)
1+ V s
The three forming lters have been implemented in Simulink-blocks. It is important
to note that for correct application of these lters the white noise inputs need to be
independent.
As can be seen from 2.2, the RCAM aerodynamic model is relatively simple. An important
limitation is that no gust penetration e ects have been modeled. Furthermore, the gust
wind eld is considered to be two dimensional. Due to these simpli cations the turbulence
windspeeds only a ect the airspeed vector Va at the aerodynamic center. In order to
obtain more realistic gust responses of the aircraft model, a more complex aerodynamic
model is needed. With the available data of the aircraft it is possible to implement the
gust penetration e ect and to model e ects of three dimensional turbulence wind elds in
a future release of the model.
For a more detailed discussion on turbulence modeling, the reader is referred to for example
[6].
2.3.3 Windshear model
For the simulation of windshear, an existing model is used from [16], which is 2-dimensional
and symmetrical with respect to the y{z plane. It is basically a combination of two counter
rotating vortices. In order to obtain zero vertical windpeeds on the ground, there are also
two virtual vortices mirrored under the earth surface, see gure 2.12.
The model is implemented as a look-up table of positions x and z in FE with corresponding
windspeeds wx and wz . The symmetry axis can be placed at a desired position on the
trajectory to be own. The windspeeds WXE and WZE at a location in the wind eld are
calculated by interpolating using the x and z coordinates of the aircraft with respect to
the inertial system in the look-up table. The wind eld is plotted in gure 2.13.
When ying through the wind eld, a registration can be created of the encountered wind
speeds. An example is given in gure 2.14. This registration was made during a simulated
Version: 1
25 Date: April 4, 1997
GARTEUR/TP-088-27

symmetry axis

vortices

ground

imag. vortices

Fig.2.12 Location of vortices of windshear model

symmetry axis
2000

1800

1600

1400
altitude (m)

1200

1000

800

600

400

200

0
−2000 −1500 −1000 −500 0 500 1000 1500 2000

x-position (m)
Fig.2.13 Wind eld due to windshear and downburst
Version: 1
Date: April 4, 1997 26
GARTEUR/TP-088-27

10
WXE
8 WZE
wind velocities WXE and WZE [m/s]
6

−2

−4

−6

−8

−10
−11000−10000−9000 −8000 −7000 −6000 −5000 −4000 −3000 −2000 −1000
x−position (XE) [m]

Fig.2.14 Encountered wind speeds in windshear/downburst

ight starting at h = 700 m with a ightpath angle = 3 deg. Windspeeds of the


windshear model enter the aircraft equations of motion via the inertial reference system.
The horizontal tailplane feels changing windspeeds a small period of time later than the
wing, resulting in a pitching moment. However, as in the case of turbulence, wind pene-
tration e ects are not taken into account.
Version: 1
27 Date: April 4, 1997
GARTEUR/TP-088-27

2.4 The implementation of the models in the software


In this section some general aspects of the implementation of the the nonlinear aircraft
state equations, the actuator models and the wind models will be discussed.

2.4.1 The simulation environment

The RCAM simulation environment consists of the nonlinear state equations, the actuators
and engine dynamics, the wind and gust models and the controller that has to be designed.
This environment has been implemented in Simulink, see gure 2.15. There is one input
for references (REFERENCES-block). Other inputs can be set to either 0 or 1, in order
to command an engine to fail, or to add constant wind or gust.
On the right hand side there are six blocks that write simulation outputs to the Matlab
workspace.

LOAD FILE EXECUTE


WITH NEW
TRIM DATA TRIM
time
DOUBLE − DOUBLE − Clock time
CLICK ME CLICK ME

traject
reference
signals
Controls
controls
controller control−
outputs inputs

REFERENCES sim
ouputs for
simulation
CONTROLLER

Engine failure
(1/0)
lon
ACTUATORS
longitudinal
0 measurements
Const. windfield
(1/0)

lat
0 lateral
Turbulence measurements
WIND &
(1/0) TURBULENCE
AIRCRAFT
wind
total wind−
inputs

Fig.2.15 Simulink block diagram of the RCAM simulation environment

The contents of the sub-blocks in gure 2.15 are explained in more detail in appendix B.
In section 2.1.3 the nonlinear state equations have been derived. These have been im-
plemented in the AIRCRAFT sub-block in gure 2.15. The twelve states used in the
equations of motion are summarized in table 2.4.
As already mentioned in section 2.1.3, the control inputs for the nonlinear state equations
consist of throttles, elevator, rudder and ailerons. There are six more inputs for wind and
turbulence. The rst group of three wind inputs is in the inertial reference system. The
Version: 1
Date: April 4, 1997 28
GARTEUR/TP-088-27

Symbol Name Unit


p x(1) = roll rate (in FB ) rad/s
q x(2) = pitch rate (in FB ) rad/s
r x(3) = yaw rate (in FB ) rad/s
 x(4) = roll angle (Euler angle) rad
 x(5) = pitch angle (Euler angle) rad
x(6) = heading angle (Euler angle) rad
uB x(7) = x component of inertial velocity in FB m/s
vB x(8) = y component of inertial velocity in FB m/s
wB x(9) = z component of inertial velocity in FB m/s
x x(10) = x position of aircraft CoG in FE m
y x(11) = y position of aircraft CoG in FE m
z x(12) = z position of aircraft CoG in FE m
Table 2.4 States de nitions

second group is for turbulence, it is de ned with repect to the body axes. The inputs for
the simulation model are given in table 2.5.
For the RCAM simulation model three groups of outputs have been de ned, see gure 2.15.
The rst group consists of longitudinal measurements, the second group consists of the
lateral measurements. The longitudinal and lateral measurements may be used as inputs
for the controller. The third group consists of simulation outputs that may only be used
for evaluation purposes. The model outputs are given in table 2.6. All inputs, outputs
and states have SI units.

The nonlinear RCAM model with the de ned states, inputs and outputs has been imple-
mented in Dymola, a software package for object-oriented modeling [8, 19, 9]. Dymola
features automatic code-generation for several simulation environments. For RCAM code
was generated for use with Matlab/Simulink, in both standard Matlab-format and C-
format (CMEX). For more information, refer to appendix F in [3].

2.4.2 Trimming and linearization routines

Trimming
Before linearizing the model or starting a simulation, the aircraft needs to be trimmed. To
nd the trim settings of the controls, the RCAM software-package is equiped with a trim
routine, in which constraints and conditions at the equilibrium point can be speci ed. For
nding equilibrium the routine makes use of the existing Simulink trim functions, based
on constrained optimization routines.
The aircraft model can be trimmed in symmetric ight in the undisturbed atmosphere (no
Version: 1
29 Date: April 4, 1997
GARTEUR/TP-088-27

Symbol Name Unit


A u(1) = aileron de ection rad
T u(2) = tailplane de ection rad
R u(3) = rudder de ection rad
TH1 u(4) = throttle position of engine 1 rad
TH2 u(5) = throttle position of engine 2 rad
WxE u(6) = Wind velocity in the x-axis of FE m/s
WyE u(7) = Wind velocity in the y-axis of FE m/s
WzE u(8) = Wind velocity in the z -axis of FE m/s
WxB u(9) = Wind velocity in the x-axis of FB m/s
WyB u(10) = Wind velocity in the y-axis of FB m/s
WzB u(11) = Wind velocity in the z -axis of FB m/s
Table 2.5 Inputs de nitions

wind). The main parameters that may be varied are:


 the aircraft mass, m,
 the center of gravity location, [xcg ycg zzg ]T ,
 the airspeed, VA ,
 the ightpath angle, ,
These parameters should only be varied within their limits, which can be found in table 2.1.
The trim routine has been implemented in the le \TRIMRCAM.M". More details can be
found in B.2.

linearization
For analysis and controller design purposes it is necessary to obtain linearizations of the
nonlinear aircraft model at desired operating conditions. In many text books on aircraft
stability and control linearized models are derived analytically, see for example [10, 6,
21]. Describing the analytical linearizations goes beyond the scope of this chapter. The
linear models of the RCAM can be obtained numerically using the linearization routine
provided with Simulink. In this way a linearization of the trimmed nonlinear model can
be calculated for small perturbations around the equilibrium point found using the trim
routine. Some more details about linearization can be found in section B.2.
Version: 1
Date: April 4, 1997 30
GARTEUR/TP-088-27

Symbol Name Unit


Measured
q y(1) = pitch rate (in FB ) = x(2) rad/s
nx y(2) = horizontal load factor (in FB ) = mgF xA + F xT -
nz y(3) = F
vertical load factor (in FB ) = mgzA -
wV y(4) = z component of inertial velocity in FV m/s
z y(5) = z position of aircraft CoG in FE = x(12) m
VA y(6) = air speed m/s
V y(7) = total inertial velocity m/s
y(8) = angle of sideslip rad
p y(9) = roll rate (in FB ) = x(1) rad/s
r y(10) = yaw rate (in FB ) = x(3) rad/s
 y(11) = roll angle (Euler angle) = x(4) rad
uV y(12) = x component of inertial velocity in FV m/s
vV y(13) = y component of inertial velocity in FV m/s
y y(14) = y position of aircraft CoG in FE = x(11) m
 y(15) = inertial track angle rad
Simulation
y(16) = heading angle (Euler angle) = x(6) rad
 y(17) = pitch angle (Euler angle) = x(5) rad
y(18) = angle of attack rad
y(19) = inertial ight path angle rad
x y(20) = x position of aircraft CoG in FE = x(10) m
ny y(21) = lateral load factor (in FB )= Fmg
yA -
Table 2.6 Outputs de nitions
Version: 1
31 Date: April 4, 1997
GARTEUR/TP-088-27

3 Point-mass and steady state analysis of the aircraft dynamics


The problem formulation for the RCAM ight control benchmark is mainly involved with
guidance along (landing) trajectories. The ight controller must be able to keep the aircraft
on the reference path and maintain the required airspeed, while ying through wind elds,
turbulence, downbursts and windshears. Tracking performance may not degrade in the
case of an engine failure or due to mass variations.
Basically, the controlled aircraft innerloop dynamics are concerned with generating and
directing the aerodynamic and thrust forces, necessary for guidance in the outerloop dy-
namics. Due to the time scale separation the slower outerloop dynamics may be considered
using point-mass approximations and steady state analysis. This is done in the following
sections.
For ight control system design it is very important to know at least the basic aspects of the
capabilities of the aircraft, like climbing and turning performance. From the analysis in this
chapter also a priori knowledge about control strategies is obtained, like elevator/throttle
co-ordination and aircraft control during an engine failure.
3.1 Point-mass ight mechanics
In this section the aircraft is considered as a point-mass. This simpli cation reduces the
complexity of the equations of motion considerably. From the analysis physical insight is
obtained in the trajectory motion of the aircraft, control strategies along trajectories and
physical limitations of the aircraft. In many cases the application of energy considerations
and Korhammer diagrams is very illuminating.
3.1.1 The reduced equations of motion
As a rst step to the analysis of stationary aircraft behaviour and energy management, the
symmetrical equations of motion in a constant wind eld with respect to the ightpath axes
(xF , zF ) will be derived. This is done from the force equilibrium along and perpendicular
to the ightpath, see gure 3.1.
Using gure 3.1, the equations can be derived:
" _ # " #" # " # " #
m V = 1 w D + 1 T+ sin W (3.1)
V
_ w 1 L ( f + iT ) cos
The measured accelerations in the aircraft can be written as:
" # " _ # " #
ax = V + g sin = m1 D w L + T (3.2)
az F V
_ g cos w D L ( w + iT )T
The aircraft load factor nz along the ightpath can be expressed as:
n = az = V
z g g
_ cos = L w D + ( f + iT )T  L
W W (3.3)
Version: 1
Date: April 4, 1997 32
GARTEUR/TP-088-27

xB

αw L
xA
RA
q, M
VA
θ αf Vw α
xF
αw
T
γ V γa
xV
D

β = βw = φ = 0

p=r=0 W

zB zA zF zV

Fig.3.1 Longitudinal aircraft motion


This approximately holds when the drag and thrust contributions to the vertical equilib-
rium are small compared to the total lift. By dividing the horizontal acceleration by g
and substituting L = Wnz a very important equation is obtained that characterizes
the energy management of the aircraft in a wind eld:
anx = V_ + sin = T D L w = T +  CD +  n (3.4)
g g W W C w z
L
CD
The term determines the aerodynamic aircraft state. For a particular angle of attack
CL
the equilibrium condition is uniquely de ned by a point on the lift-drag polar (see g-
ure 3.2), determining the value of the aerodynamic coecients CL and CD . The coecients
CL

CD
CL

γgl = - γ

xv C W= - CR Lift-drag
polar

xa VA CD

zv za

Fig.3.2 Force equilibrium for unpowered ight in undisturbed atmosphere

are parallel to the forces L and D induced by dynamic pressure on the reference surface
S . This implies that the axes system of the lift-drag polar and the aerodynamic (or wind)
Version: 1
33 Date: April 4, 1997
GARTEUR/TP-088-27

axes system are the same. The resulting aerodynamic force coecient is CR = 1=2RVAA2 S ,
which is parallel to the aerodynamic force vector RA .
Equilibrium in unpowered ight is only satis ed if:
RA + W = 0 (3.5)
or, in terms of coecients, CR must be equal to CW ( gure 3.2), where CW is the coecient
for the aircraft weight: CW = 1=2W VA2 S . The weight vector determines the direction of the
zV -axis, so that the angle between the zV axis and the zF -axis ( ) is determined. In steady
unpowered ight without wind the ightpath angle will thus be equal to the glide angle
gl (always negative), see gure 3.2:
tan gl = CCD = D L  gl (3.6)
L
Substitution in equation 3.4 for small ightpath angles and with the assumption that
nz  1, gives:
T = + + + V_ (3.7)
W gl w g
With the aerodynamic ightpath de ned by a = + w , the equation can be written as
follows:
T = + + V_ (3.8)
W gl a g
This expression shows the relation between the thrust input T , the aerodynamic state gl
and what is available for manoeuvring: a and V_ . By writing it in terms of pure ightpath
quantities, this equation has become very simple.
3.1.2 Energy considerations and the Korhammer diagram
In this section the aircraft motions will be considered from an energetical point of view.
If rotational energy can be neglected, the total energy can be written as:
E = Ekin + Epot = m2 V 2 + mgh (3.9)
and dividing by the aircraft weight, the speci c energy is obtained:
E = V2 +h =h (3.10)
W 2g E
which is equal to the energy height hE .
hE is the height that the aircraft can achieve with the available energy level, when all
kinetic energy is transferred to potential energy. A dual de nition can be made for the
energy speed VE .
The rate of change of the energy height can be written as (it is assumed that the aircraft
mass is constant):
E_ = V_ V + h_ = V ( V_ + sin )
h_ E = W (3.11)
g g
Version: 1
Date: April 4, 1997 34
GARTEUR/TP-088-27

The right hand side of this equation is also called the speci c surplus energy, which is
often related to the climb performance. Therefore, similarly to the ightpath angle , the
energy ightpath angle E is de ned as:
h_ E = V sin E (3.12)
where
_ E_
sin E = Vg + sin = WV (3.13)
This quantity is a measure for the energy surplus with respect to the aircraft weight and it
is the total amount of energy that can be extracted from the system during a manoeuvre.
This term is often referred to as potential energy ightpath angle. However, E is a
measure for the total energy and not only for the potential energy.
With equation 3.4 the energy rate can be written as:
   
T + CD + n = V T V ( )L = P P
E_ = WV W CL w z w gl a r (3.14)
The right hand side of this equation describes the available power provided by thrust (Pa )
and the power consumed by the aerodynamic force (Pr ). The di erence between these
quantities, the surplus power, is available for manoeuvring.
This energy di erence can be considered with respect to any energy level. Altitude is
mostly de ned with respect to the earth surface, so that the potential energy equals mgh.
The aerodynamic state of the aircraft is related to the airspeed VA rather than the inertial
speed V . As a control variable the airspeed is also more important and therefore plays
a key role in the analysis of aircraft motion. For these reasons in literature the kinetic
energy is mostly related to the airspeed: Eakin = 21 mVA2 . In this respect a mixed energy
quantity, the aero-kinetic energy Ea , can be de ned as:
Ea = Eakin + Epot = m2 VA2 + mgh (3.15)
However, in this analysis the de nition of the total system energy according to equation 3.9
is used.
Both force equilibrium and energy management can be visualized in the so called
Korhammer-diagram. An example of this diagram is shown in gure 3.3. In a Korhammer
diagram all forces and accelerations acting in the aircraft center of gravity can directly be
related to the CL CD polar. In order to improve the visibility of some force components,
the angles are exaggerated. The only quantities that are drawn in the right proportions
with respect to each other, are the components of equation 3.1.
The diagram is self explanatory, but to understand the dynamics involved it is necessary
to proceed in a prescribed fashion. In rst instance the aerodynamic trim point on the
lift-drag polar is chosen, this determines the value of the lift coecient CL and also the
coecient CR . The direction of RA has now been determined. With the given wind angle
w and ightpath angle the directions of the ightpath axes and the vehicle carried
Version: 1
35 Date: April 4, 1997
GARTEUR/TP-088-27

CL
α

γgl Lift Drag polar

xT
xa VA
CR CD
α + iT
Vw αw
xf V γ

xv
RA
W
γa
γgl
αw γ

α f + iT T
xf W sin( γ )
RAsin( α w+ γ gl
)
zv
zb zf
za

Fig.3.3 Korhammer Diagram


vertical axes are determined. The direction of the weight vector W in the diagram is now
also known. The angle + iT gives the direction of the thrust T required for equilibrium
ight. T and RA can be adjusted to close the force polygone. In this way the required
T and RA are obtained and with RA = CR 21 VA2 S the required airspeed for trim can be
calculated.

Some further system characteristics can be deduced geometrically from this diagram. In
the case of constant thrust, the sum of the glidepath angle and aerodynamic ightpath
angle remains constant when a new trim point is chosen. This can also directly be seen
from equation 3.8 (where gl is positive and V_ = 0). The aerodynamic ightpath angle
a , which characterizes the ightpath with respect to the airmass, does not depend on
the wind but depends on the di erence between thrust and drag. A stationary air motion
has an e ect on forces and energy management only through the wind angle w . Lift is
a ected by this angle such that a positive w (down wind) results in a smaller in order
to keep a constant. A part of the available power is thus absorbed by the wind. Finally,
thrust compensates the energy loss induced aerodynamically by w gl and provides
the additional power that is required for a speci ed ightpath angle (speci c potential
energy rate).

Aircraft performance is often studied through power analysis. This power analysis is
carried out with respect to the ightpath axes. The available power provided by thrust is
V T cos ( f + iT ). The resulting potential energy increase is V W sin while loss of power
due to aerodynamics is V RA sin( w gl ). If the thrust is aligned with the horizontal the
Version: 1
Date: April 4, 1997 36
GARTEUR/TP-088-27

following exactly holds:


T = sin( + )
W w gl

For an accelerated ight the force equilibrium has to be extended with the acceleration
term ( ddtV )g , as shown in gure 3.4. This term results in a kinetic energy rate of E_ kin =
mV V_ and a potential energy rate of E_ pot = WV sin . Further, the total energy rate
or power surplus becomes: W sin E = W sin + mV_ . The energy ightpath angle can
be deduced through rotation of the weight vector until the ightpath contribution of the
acceleration term is compensated. The energy angle is the maximum ightpath angle that
can be achieved at the current thrust setting.
In the same diagram it is also possible to show the acceleration contributions to external
forces in the ightpath axis system. For the horizontal and vertical contributions we have:
maxF = Wsin E = Wnx (3.16)
mazF = Wnz
both vectors determine the direction of the sensed acceleration (virtual gravity accelera-
tion) which takes a value  with respect to the zF axis:
tan  = axF = sin E
azF nz (3.17)
All the quantities as mentioned here which characterize the aircraft performance are now
put into one framework describing and providing insight in the physical process of ma-
noeuvred ight. This means that in the Korhammer diagram all gliding, accelerated or
decelerated ights with or without wind can be represented, provided that the force and
speed vectors remain within the aircraft symmetry plane (no lateral e ects).

xa
xa

ρ
W
γ γ γ
gl αw gl αw γ

γa
RA γa RA
W γE
T m a zf T
= W n tz
m dV αk+ ιT
dt g m dV
dt g
x .
f
mV
W sin γ
sin γ E γ gl)
R Asin( α w W sin γ E
+
x m a xf= W
f
za zf zv za zf

Fig.3.4 Korhammer Diagram; accelerated ight


Version: 1
37 Date: April 4, 1997
GARTEUR/TP-088-27

3.1.3 Aircraft responses to control inputs


Elevator input
For the conventional control of the longitudinal motion elevator and thrust are available as
e ectors. The e ect of these controls on the aircraft motion and the underlying physical
principles will be discussed now.
The aerodynamic state of the trimmed aircraft is determined by a point on the lift-drag
polar. The corresponding glide angle determines the direction of the total aerodynamic
force (RA ) with magnitude given by
RA = CR 21 VA2 S  nz W (3.18)
The stationary equilibrium point on the polar is determined by the angle of attack and
is achieved when there is moment equilibrium about the yB -axis, thus:
MuVA + M + ME E zAPT T = 0 (3.19)
Applying an elevator step input results in an angular acceleration about the yB -axis with
an initial response given by:
2
q_ = V2AI S cCmE E (3.20)
y
The steady state value of the angle of attack for the elevator step input is given by:
C
 ss = CmE E (3.21)
m
This means that an elevator de ection directly a ects the location of the trim point on
the polar. For a negative elevator de ection the total aerodynamic force coecient CR
will increase. The total aerodynamic force (RA ) will remain approximately constant, so
that the airspeed VA has to decrease. In normal operation the glide angle gl will decrease
in absolute sense. When the thrust setting is held constant, this will result in an increase
in ightpath angle .
The de ection of spoilers or aps directly a ects lift and drag. Depending on the ap
or spoiler setting, the polar moves to a range of higher lift and drag coecients, see
gure 3.5. If the contribution of the spoiler or ap setting to the pitching moment can
be neglected, the angle of attack will remain constant. When aps are de ected, CR will
increase and depending on the ratio CCDL the glide angle gl will vary accordingly. Since
the total aerodynamic force RA remains unchanged, airspeed has to decrease, explaining
the primary function of aps to allow for a lower landing speed. A decrease of the glide
angle can be used to allow a lower throttle setting or to increase ightpath angle and vice
versa.
Generally, a spoiler or ap de ection also induces a large disturbance in the moment
equilibrium, a ecting the aircraft response. This has to be compensated through an
elevator de ection.
Version: 1
Date: April 4, 1997 38
GARTEUR/TP-088-27

CL CL

δf

α α

CR

CR

xa VA CD xa VA CD

RA RA
W W

γ − γ gl γ − γ gl

T T
za zv zv za

Fig.3.5 E ect of changing aerodynamic trim point in Korhammer diagram: angle of attack
and ap de ection
Backside operation
If the angle of attack is further increased until the total aerodynamic force coecient
is tangent to the polar (see gure 3.6), then also the point is reached where the glide
path angle has achieved its minimal value in absolute sense. The trim speed at which this
operating point is achieved is called minimum drag speed. This trim point could also be
called the speed for minimum glide angle and is aerodynamically the most ecient point
for cruising. The slope of the polar in this situation is equal to the ratio CCDL , therefore:
CD = @CD = CD (3.22)
CL @CL CL
In this case the glide angle remains approximately constant for small variations in angle of
attack. This implies that at constant thrust the ightpath angle also becomes independent
of the angle of attack. An elevator de ection then implies, through its e ect on CR , only
a change in airspeed. If the angle of attack is further increased, the glide angle starts to
increase again. However, variations in angle of attack for this region on the polar imply
CD > CD . Since the total force coecient grows, the speed will decrease further. This
CL CL
speed region is called "below the minimum drag speed", where the e ect of angle of attack
on the ightpath angle is reversed.
The increase of the angle of attack induces an increase in the glide angle, which at a
constant thrust can only be compensated by a decrease in ightpath angle. This means
that a negative elevator de ection (up) results in an unexpected decrease of the ightpath
angle. At the same time CR increases so that the speed decreases, as in the normal
situation.
The origin of this phenomenon can directly be seen from the Korhammer diagram. The
(constant) thrust vector can not compensate the overproportional growth of the drag
Version: 1
39 Date: April 4, 1997
GARTEUR/TP-088-27

CL CD CL
CD
CL
CL

α
CD
α CL

CD
CL
γ gl
γ gl
min

CR

CR
xa VA CD xa VA CD

RA W
RA W

γ max γ

T max T
zv za za zv

Fig.3.6 Korhammer Diagram; ight below minimum drag speed


contribution. Therefore, in this region of the ight envelope a decrease of the airspeed at
constant ightpath angle is only possible by increasing the thrust.
In practice, pilots operating in this part of the ight envelope have the tendensy still to
control the ightpath angle by means of the elevator, resulting in a continuously decreasing
airspeed. For this reason this phenomenon has been called \speed instability". However,
it is not instability of the aircraft itself, because it has stationary operating points in this
part of the ight envelope.
The origin of the phenomenon lies in the all-pass behaviour of the transfer function between
elevator and ightpath angle (i.e. change of sign in the gain). In section 4.5.2 this will be
shown by plotting that speci c zero in the complex plane at di erent trim speeds. For the
RCAM the minimum drag speed is approximately 70 m=s (m = 120000 kg).
Generally, ying on the backside of the power curve may occur at low speeds when the
aircraft is in the approach to land or when the aircraft is operating at high altitudes, due
to the required large value of the lift coecient to compensate for the lower air density.

Throttle input
To analyse the e ect of thrust changes the equation of motion along the ightpath can be
written in the form:
T =  CD + WZE  n + sin + V_ (3.23)
W C V z g
L A
which reveals that the required thrust is a function o :
 the vertical load factor nz
 the aerodynamic cost due to glide angle gl
 the vertical component of the wind eld WZE
Version: 1
Date: April 4, 1997 40
GARTEUR/TP-088-27

 the ightpath angle


 the acceleration along the ightpath V_ .
the e ect of thrust changes has a geometric interpretation in the Korhammer diagram,
see gure 3.7. For example, in the case that airspeed VA has to remain constant during

γ gl

γ gl

xT
xa

α + iT
xa

W
RA
W
RA

∆γ
∆γ
T
T
∆ iT

θ + iT
xv
za zv za zv

Fig.3.7 Korhammer Diagram; e ect of thrust commands

an increasing tailwind WXE , then, using the relation


V_ = V_ A + W_ XE ; (3.24)
and with nz  1, the following holds:
T = + + WZE + W_ XE (3.25)
W gl V g A
Obviously, with increasing tail wind WXE (horizontal windshear) thrust needs to be in-
creased, meaning that the aircraft has to be accelerated in order to keep airspeed constant.
In order to be able to develop a control strategy for the ightpath control, the steady
state e ect of a thrust increment will be analysed. Only the stationary case without wind
is considered. The immediate response of the ightpath angle and the angle of attack
(and therefore the aerodynamic state of the aircraft characterized through gl ) on a thrust
increment cannot occur without accelerations. The thrust increment results in an initial
response of the acceleration:
V_ (t=0) = mT (3.26)
= V_ At=0
in the case that V_ w = 0. This complies with Newton's Law. For the steady state behaviour
an additional assumption has to be made on the moment about the yB -axis. As long as
the pitching moment equilibrium does not change, the angle of attack will return to its
original value after a perturbation (static stability). This means that the aircraft always
Version: 1
41 Date: April 4, 1997
GARTEUR/TP-088-27

returns to the original trim point on the polar. Therefore the glide angle gl and the
airspeed VA remain constant. Under this condition the steady state response of speed to
thrust is:
VAss = Vss = 0 (3.27)
Thus only the ightpath angle will be a ected:
 = T
ss W (3.28)
This means that in steady state all energy input by thrust is completely transformed into
potential energy, while airspeed remains constant.
In the case that T sin( f + iT ) is taken into account, a slight modi cation has to be made.
In the Korhammer diagram in gure 3.7 the glide angle gl remains xed, therefore also
the direction of the resulting aerodynamic force R remains constant. Mostly, the thrust
has also a small contribution in the negative zV -direction, thus upwards when  + iT > 0.
Since the gravity contribution is constant, a thrust increase not only results in \opening"
the Force triangle, but also results in a shorter vector R, and therefore a smaller airspeed
VA . This system response is opposite to what one may expect. Furthermore, it is also
opposite to the initial response of the system.
The magnitude of the total aerodynamic force vector, and therefore also the airspeed, will
be invariant to small thrust increments only if the thrust vector is perpendicular to the
gravity vector. Only in that case thrust can be changed while RA remains unaltered,
resulting only in opening the force polygone and thus in increasing the ightpath angle.
Another e ect of thrust is when the thrust vector induces a contribution to the pitching
moment equilibrium. In a trim point this moment due to thrust is compensated by a
steady state aerodynamic moment (induced by  ):
1 V 2 SC
2 A m0 zAPT T0 = 0 (3.29)
If there is a thrust increment new moment equilibrium is achieved only aerodynamically
by:
1 V 2 SC  z T = 0 (3.30)
2 A m APT
which means that
 ss = + 1 zAPT T (3.31)
2 VA SCm
2
This means that a positive thrust moment induces an increase in angle of attack , there-
fore an increase in CR and thus a decrease in the airspeed VA . As a consequence, when
ying on the right side of the power curve, the glide angle gl decreases (in absolute sense).
The steady state airspeed change due to a thrust increment remains small. It will decrease
when the thrust vector induces a pitch up moment. The initial response to a stepwise
thrust increment is:
q_ = zAPTI T (3.32)
y
Version: 1
Date: April 4, 1997 42
GARTEUR/TP-088-27

Suppose that it is possible to change the thrust vector by means of iT . In that case both
force and moment equilibrium are a ected. A positive iT results in a decrease of the
total aerodynamic force and thus in an decrease of the airspeed. Furthermore, the weight
vector will rotate to compensate for the acceleration contribution, so that the ightpath
angle also increases.
Version: 1
43 Date: April 4, 1997
GARTEUR/TP-088-27

3.2 Performance in steady symmetric ight


In this section some basic aircraft performance parameters will be reviewed and calculated
for the RCAM. In the equations it will be assumed that the aircraft weight is constant and
the air density is equal to 0 at sea level. E ects of compressibility will not be taken into
account, since the RCAM aircraft will be ying at approach speeds. These assumptions
will simplify most of the analysis considerably.
Most of the analytic expressions in this section were obtained from [20, 7].
3.2.1 Basic relations
In section 3.1.1 equations of motion with respect to the ightpath axes have been derived.
From these equations the force equilibrium along the ightpath has been obtained in the
following form (equation 3.4):
V_ + sin = T +  CD +  n (3.33)
g W C w z
L
Rewriting this expression, with w = 0:
T = CD n + V_ + sin (3.34)
W C z g
L
Now only stationary ight will be considered, thus all accelerations are set to zero. Thus,
with V_ =g = 0 and nz = 1 is obtained:
T = CD + sin (3.35)
W CL
In many textbooks on aircraft performance, this expression is the basis for analysis of
stationary symmetric ight.
In section 3.1.2 also the expression has been derived for the available power for manoeu-
vring (eq. 3.14, no wind, no accelerations):
T CD  = V T V D = WV sin
Pa Pr = WVA W CL A A A (3.36)
The term Pa = TVA is the available power from the thrust, the term is Pr = DVA is
the required power to compensate for the energy loss due to drag and the right hand
term contains the rate of climb, RC = h_ = VA sin . From the equations 3.36 and 3.35
three basic performance parameters can be deduced (with L = W = CL 21 Vair 2 S and
D = CD 12 Vair
2 S ):
s
VA = W2 1 (3.37)
S  CL
D = CCD W (3.38)
Ls
Pr = W W 2 CD2 (3.39)
S  CL3
Version: 1
Date: April 4, 1997 44
GARTEUR/TP-088-27

At a given altitude (given air density ) and a given aircraft weight, the airspeed, drag and
required power are a function of the angle of attack and thus a function of the equilibrium
point on the lift-drag polar. The rate of climb and ightpath angle also depend on the
engine control setting.
From the equations minimum values can be calculated for the airspeed (at CLmax ), for

the drag (at (CL =CD )max ) and for the required power (at CL3 =CD2 max ). The minimum
airspeed, or stall speed, is important for safety reasons. In horizontal ight the longest
possible range is achieved when ying at the minimum drag speed, while maximum en-
durance ight requires a minimum fuel ow and therefore minimum required power.
In gure 3.8 these quantities are easiliy related to the lift-drag polar.

3 3

C CL 3
2.5 2.5
CDL CD 2
minimum Pr
2 2

minimum D
CL

1.5 1.5

1 1

0.5 0.5

CL 3
20 40 60 80CD 2100
0 0
0 0.2
CD
0.4 0 2 4 6 8
CL 10
CD
Fig.3.8 Performance related to the lift-drag polar

3.2.2 Analytical performance calculations


It will be clear that longitudinal performance of the aircraft is mainly determined by its
aerodynamic characteristics (the lift-drag polar) and the characteristics of the powerplant.
For the RCAM both are modelled in an extremely simple manner. The thrust is constant
for the whole range of airspeeds reasonable for the landing con guration ( 55 85 m=s).
In the model no altitude e ects are taken into account. The lift-drag polar is modelled
using the well known parabolic approximation:
CD = CD0 + k (CL CL0 )2 (3.40)
where CD0 is the zero-lift drag coecient and k is a constant for the induced drag term.
Only at higher angles of attack this relation no longer holds, see equation 2.28. Although
the modeling is simple, an analytical performance analysis becomes much more tractable.
Version: 1
45 Date: April 4, 1997
GARTEUR/TP-088-27

For the RCAM model the parameters of the polar are (table 2.2):
CD0 = 0:13
CL0 = 0:45 (3.41)
k = 0:07
The total drag can be calculated from:
D = CD0 12 Vair 2 S + k (C 1 2S=D +D
L CL0 )2 2 Vair 0 i (3.42)
D0 is the zero lift drag and Di is the induced drag of the aircraft. All minima for parameters
to be calculated are valid for steady state. A very important constraint is therefore vertical
equilibrium: L  W . Substitution gives:
CL = W (3.43)
1 V 2 S
2 air
The minima for parameters can now be calculated analytically.
The maximum angle of attack for the RCAM is 18 deg. Using the simple parabola ap-
proximation that was in the original model, this corresponds to a maximum lift coecient
of CLmax = 3, which is quite high. For this reason the lift-drag relation was modi ed, so
that the polar becomes at for higher angles of attack. The original and modi ed polars
are compared in gure 3.9. As the maximum lift coecient a value of CLmax = 2:75 has
been chosen.

original
3.5

2.5

modi ed
CL

1.5

0.5

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

CD
Fig.3.9 Modi cation of the lift-drag polar

The stall speed at a mass of m = 120000 kg now follows from:


s
:7 104 2 1 = 52 m=s
Vstall = 117260 (3.44)
1:225 2:75
Version: 1
Date: April 4, 1997 46
GARTEUR/TP-088-27

The speed for minimum drag corresponds to the condition dD=dVA = 0. Di erentiating
equation 3.42 with respect to VA gives (with substitution of eq. 3.43):
v
u s
u W 2 k
VDmin =t S  CD0 + kCL2 0 (3.45)

Substitution in equation 3.42 gives:


q
Dmin = 2W ( k(CD0 + kCL2 0 ) kCL0 ) (3.46)
This expression could of course also be found from:
Dmin = (C =CW ) (3.47)
L D max
The minimum drag speed for the RCAM can be calculated by substituting the following
values:  = 0 = 1:225 kg=m3 , CD0 = 0:13, CL0 = 0:45, k = 0:07, S = 260 m2 . For the
weight is taken: W = 120000  9:81 = 117:7 104 N
v
u s
u : 4 0:07
VDmin = t 117 7 10 2
260 1:225 0:13 + 0:07 0:452 = 71:2 m=s (3.48)
This is a typical speed for the nal approach of a transport aircraft. For higher aircraft
weight the minimum drag speed will also be higher. In that case the aircraft will y on
the backside of the power curve, as described in section 3.1.2.
The minimum drag is:
q
Dmin = 2 117:7 104 ( 0:07 (0:13 + 0:07 0:452 ) 0:07 0:45 = 16:2 104 N (3.49)
Substitution of eq. 3.40 and eq. 3.43 in the expression for the required power (eq. 3.39)
gives:
Pr = CD 21  Vair
3 S = C 1 V 3 S + k (C
D0 2 air
1 3S
L CL0 )2 2 Vair (3.50)
Di erentiating with respect to VA gives:
v
u pq 2
u
u 2 W 1 k 4kCL0 + 3CD0
VPrmin =t S3 2
CD0 + kCL0 (3.51)

For the RCAM this speed can be calculated:


s p p
VPrmin = 1:225 2 117:7 104 1 0:07 4 0:07 0:452 + 3 0:13 = 55 m=s (3.52)
260 3 0:13 + 0:07 0:452
This is very close to the stall speed of the aircraft. In practice the parabolic approximation
of the polar is no longer valid in this area. For the modi ed RCAM polar however, the
curve has still a parabolic shape up to lift coecients of 2.5, see gure 3.9. At the speed
of 55 m=s the lift coecient is:
4
CL = 1 117:7 102 = 2:4 (3.53)
2 1:225 55 260
Version: 1
47 Date: April 4, 1997
GARTEUR/TP-088-27

For analysis of the RCAM model the parabolic estimation is thus still good enough.
The minimum required power can be calculated by substitutions:
CD = 0:13 + 0:07(2:4 0:45)2 = 0:40 (3.54)
Prmin = 0:40 21 1:225 553 260 = 10:6 103 kW

5
x 10
4.5
Max. Thrust
4

3.5 xcg = 0.23c


h=0m
Tmax , D (N)

150000 kg
2.5

2 120000 kg

1.5
100000 kg
1
40 50 60 70 80 90 100 110 120 130
Vtrim (m/s)
Fig.3.10 Performance diagrams for di erent masses (forces)

3.2.3 Performance diagrams


In the previous subsections some important parameters for performance analysis have
been introduced. These parameters can directly be related to the lift-drag polar. Using
the parabolic estimation analytical performance analysis becomes very easy. The goal of
this chapter is not only to understand and to analytically calculate quasi-stationary aircraft
behaviour, but also to obtain insight in the actual performance level of the aircraft and
to produce some gures from which a quick impression of the aircraft characteristics and
performance can be obtained. In this subsection performance diagrams are introduced
that have been created using actual data from trimming the aircraft model at di erent
speeds and masses using the trim routine supplied with the software.
In gures 3.11 and 3.10 two types of performance diagrams have been drawn for the RCAM
aircraft. These diagrams were obtained by trimming the nonlinear model in horizontal
ight at di erent speeds and at three di erent weights.
The lowest airspeeds where the curves begin are the stall speeds. The range of airspeeds is
not very realistic, since an airspeed of 130 m=s is high for an aircraft with fully extended
aps. Theoretically the range can be made larger, so that the point where the Pa and Pr
curves intersect will be visible. At this point the maximum airspeed is achieved: ying at
Version: 1
Date: April 4, 1997 48
GARTEUR/TP-088-27

4
x 10
5.5
xcg = 0.23c
5 h = 0m
100000 kg ||- Pa
4.5 120000 kg - - - -
150000 kg .......
4
VRCmax
Pr
Pa , Pr (kW)

3.5

3
VDmin
V
VstallPrmin
2.5

1.5

0.5
40 50 60 70 80 90 100 110 120 130
Vtrim (m/s)
Fig.3.11 RCAM performance diagrams (power)

higher speeds requires more power than available from the power plant.
For the rate of climb holds (equation 3.36):
RC = Pa W Pr (3.55)
For constant weight the rate of climb will be maximum where Pa Pr is maximum, thus
where the Pa and Pr lines in gure 3.11 are parallel. From this gure can be seen that
for a weight of 117:7 104 N (120000 kg) the RCmax occurs at a speed of  89 m=s. The
power surplus is at that point approximately 1:9 104 kW , so that:
1:9 107 = 16 m=s
RCmax = 120000 (3.56)
9:81
The maximum ightpath angle max can be deduced from gure 3.10. From equation 3.35
can be seen that:
T D
= arcsin W (3.57)
The term between the brackets is often referred to as the speci c excess power. For a
given aircraft weight the maximum ightpath angle occurs when T D is maximum. In
the case of the RCAM maximum thrust is assumed to be invariant to the airspeed. The
maximum ightpath angle is therefore at the minimum-drag speed. This is illustrated in
gure 3.12. In equation 3.49 the minimum drag appeared to be 162 103 N . The maximum
available thrust is 411 103 N , so that for a weight of 117:7 104 N the speci c excess power
equals:
!
T D = arcsin 411 103 162 103 = 0:21 (3.58)
W 117:7 104
Version: 1
49 Date: April 4, 1997
GARTEUR/TP-088-27

CL CD
CL

CD
CL

γ gl
min

CR
xa VA CD

RA
W

γ max

T max
zv za

Fig.3.12 Korhammer Diagram; maximum ightpath angle


The maximum ightpath angle is:
= arcsin (0:21) = 12 deg
In gure 3.13 the maximum ightpath angles for di erent trim speeds at di erent aircraft
weights can be found. The calculated maximum ightpath angle is explicitly indicated.
The data for the gure have been obtained from trimming the nonlinear model, where
the in uence of the negative incidence angle of the stabilizer has been taken into account.
For this reason the maximum ight path angle in the gure is somewhat lower than the
calculated angle.
There is another important quantity, namely the smallest achievable ightpath angle at
constant airspeed. For the RCAM aircraft no speed brake devices have been modelled.
These will not be available for the controller to be designed. The only way to decrease or
dissipate power, is to decrease thrust. Only a minimum value of THmin = 0:5 deg throttle
setting per engine is allowed. This means a minimum thrust of 2  m120 gTHmin =57:3 =
2  120000 9:81 0:5=57:3 = 20:54 kN . In gure 3.14 the minimum possible ightpath
angle for di erent trim speeds at di erent aircraft weights can be found. These have been
calculated using:
 Tmin D 
= arcsin W (3.59)
From gure 3.14 appears that at the minimum drag speed the minimum ightpath angle
is 7:4 deg. The evaluation trajectory for the controller evaluation (see [2]) is own at an
airspeed of 80 m=s, while in the nal phase a glide path of 6 deg has to be own, while
keeping speed constant. For the mass of 120000 kg the maximum descent angle appears
to be = 7:7 deg. If the transition from horizontal ight to the glide path is smooth
Version: 1
Date: April 4, 1997 50
GARTEUR/TP-088-27

13

12

11

10

9
(deg)

5 xcg = 0.23c
h = 0m
4 100000 kg ||-
120000 kg - - - - V max
3 150000 kg .......
40 50 60 70 80 90 100 110 120 130
Vtrim (m/s)
Fig.3.13 Maximum ightpath angle vs speed

−6

−7

−8

−9
(deg)

−10

−11

−12

−13
xcg = 0.23c
−14 h = 0m
100000 kg ||-
−15 120000 kg - - - - V min
150000 kg .......
−16
40 50 60 70 80 90 100 110 120 130
Vtrim (m/s)
Fig.3.14 Minimum ightpath angle vs speed
Version: 1
51 Date: April 4, 1997
GARTEUR/TP-088-27

enough (no severe transients), this part of the trajectory can be own without saturating
the engines at their minimum throttle setting.

In this section only static performance has been analysed. However, the thrust excess that
in equation 3.35 is used for a ightpath angle, can of course also be used for accelerating
the aircraft. This does not seem to change the analysis basically, because it is in fact only
a matter of distribution of excess power. However, speed changes imply a change in the
aircraft state. For example, the equilibrium point on the lift drag polar is going to move
if the aircraft is accelerating. Therefore the analysis should be carried out more carefully
and for limited speed changes.
3.2.4 Analytical expressions for weight variations
In the gures of the previous section the curves were drawn for varying aircraft masses.
From the expressions in section 3.2.1 it is very easy to derive the in uence of weight on
parameters like the minimum required power with the corresponding airspeeds. From
equation 3.39 the following ratios can be deduced for a condition (1) compared to a
condition (2) with both the same equilibrium point on the lift-drag polar.
s
VA2 = W2 (3.60)
VA1 W1
D2 = W2 (3.61)
D1 W1
s
Pr2 =  W2 3 (3.62)
Pr1 W1
For a given point on the lift-drag polar (given angle of attack) the ratio of drag for the
two conditions can be expressed as a function of the corresponding trim speeds:
D2 =  VA2 2 (3.63)
D1 VA1
The same function can easily be derived for the required power:
Pr2 =  VA2 3 (3.64)
Pr1 VA1
For a given weight and altitude the three parameters only depend on the equilibrium
point on the lift drag polar. Therefore the minima for these quantities, which are also
only determined by the equilibrium point, vary in the same way as in equation 3.62.
If for example an aircraft mass of m = 150000 kg is compared to the nominal weight of
m = 120000 kg, the ratio of stall speeds becomes:
r
VA2 = 150000 = 1:1
VA1 120000 (3.65)
The ratio of minimum drags can be found from:
Dmin2 = 150000 = 1:25 (3.66)
Dmin1 120000
Version: 1
Date: April 4, 1997 52
GARTEUR/TP-088-27

The corresponding ratio of speeds is of course also 1.1. The ratio of minimum required
power levels:
s
Pr2 =  150000 3 = 1:4 (3.67)
Pr1 120000
These ratios can easily be veri ed in gures 3.10 and 3.11.
Version: 1
53 Date: April 4, 1997
GARTEUR/TP-088-27

3.3 Turning performance


In this section some aspects of turning performance will be reviewed. Since the aircraft in
the approach to land is ying in the relatively crowded airspace near the airport, a certain
level of manoeuvrability (in this case turning performance) is thus required. However, the
aircraft is also operating in a lower speed range, thus safety limits like the maximum lift
coecient become important.
In this section it is assumed that the aircraft is performing a turn at a constant speed and
with a constant turning radius, rf . The atmosphere is assumed to be in rest.
The following very simple relation holds:
V = rf _ (3.68)
3.3.1 Equilibrium in co-ordinated turns
The equations for force equilibrium in a co-ordinated turn can easily be deduced from
gure 3.15. The equations in the aerodynamic axes can be written as:

.
W
g Vχ
W sin µ

µ
y
b

zb
W

Fig.3.15 Aircraft in co-ordinated turn

L = W cos  + Wg V _ sin  (3.69)


where  is the bank angle and _ the turning rate. For lateral equilibrium can be written:
W sin  = Wg V _ cos  or: tan  = Vg_ (3.70)
Substituting the latter equation into equation 3.69 gives:
L = W= cos  (3.71)
Version: 1
Date: April 4, 1997 54
GARTEUR/TP-088-27

and thus the load factor along the zA -axis becomes:


nz = W L = 1 (3.72)
cos 
For equilibrium along the xA -axis still holds T D = 0.

3.3.2 Parameters for turning performance

From the simple derivation of the force equilibrium in the previous section immediately
follow three important bounds on turning performance:
 The maximum thrust, Tmax
 The maximum load factor allowed, jnz jmax
 The maximum lift coecient, CLmax
In the discussion of turning performance these will give the bounds that determine achiev-
able bank angles, turning radii and turning rates.
If there is vertical equilibrium, the expressions from the previous sections for VA , D and
Pr can be easily modi ed by substitution of equation 3.71 and 3.72:
s
VA = nz W 2 1 (3.73)
S  CL
D = nz W CCD (3.74)
sL 2
Pr = nz W nz SW 2 CCD3 (3.75)
L
For a given weight (W ), bank angle (), wing area (S ) and altitude (), the minimum
values of these performance parameters are only determined by a certain equilibrium
point on the lift drag polar. For the stall speed that point is CLmax , for the minimum drag
(CL =CD )max and for the minimum required power the point corresponds to (CL3 =CD2 )max .
For this reason the optima for di erent bank angles are simply related by the load factors.
Comparison of two turns (1) and (2) gives:
s
Vstall 2 = nz 2 = r cos 1 (3.76)
Vstall 1 nz 1 cos 2
Dmin 2 = nz 2 = cos 1 (3.77)
Dmin 1 nz 1 cos 2
and:
v
u !3 s 
Prmin 2 = u
t z 2 = cos 1 3
n (3.78)
Prmin 1 nz 1 cos 2
Version: 1
55 Date: April 4, 1997
GARTEUR/TP-088-27

Using the ratio (see equation 3.75):


VA 2 = r cos 1 (3.79)
VA 1 cos 2
the ratio of speeds where the minima occur for di erent bank angles, can be calculated
from:
s
VDmin 2 = Dmin 2 (3.80)
VDmin 1
s Dmin 1
VPrmin 2 = 3 Prmin 2 (3.81)
VPrmin 1 Prmin 1
In gure 3.17 and 3.16 performance diagrams have been drawn for a number of bank
angles.
5
x 10
5

4.5
Max. Thrust
4
 = 60 deg
3.5
Tmax , D (N)

3
 = 45 deg

 = 25 deg
2.5

2 h=0m
xcg = 0.23c
 = 0 deg
1.5
m = 120000 kg
1
50 60 70 80 90 100 110 120
Vtrim (m/s)
Fig.3.16 RCAM performance diagrams for di erent bank angles (force)

From equation 3.70:


V _ = tan 
g
Furthermore, from equation 3.68 follows:
V 2 = r tan  (3.82)
g f
The expression for the load factor now becomes:
v
u !
1 q u
t V 2 2
nz = cos  = 1 + tan  =
2 1 + gr (3.83)
f
Version: 1
Date: April 4, 1997 56
GARTEUR/TP-088-27

4
x 10
6

h = 0m
5
xcg = 0.23c
m = 120000 kg
4
Pa
Pa , Pr (kW)

3
 = 60 deg

2  = 45 deg
 = 25 deg
1
 = 0 deg

0
50 60 70 80 90 100 110 120
Vtrim (m/s)
Fig.3.17 RCAM performance diagrams for di erent bank angles (power)

3.3.3 The minimum turning radius

As noted in the previous subsection, the turning performance of an aircraft is mainly


determined by three limitations: the load factor, the maximum lift coecient and the
available thrust.
The minimum achievable turning radius for the rst two parameters can be studied ana-
litically. From equation 3.83 the following expression can be derived (no wind):
2
rf = q V2A (3.84)
g nz 1
For a given airspeed the radius decreases monotoneously with increasing load factor. The
minimum achievable radius limited by the factor is:
2
(rfmin )n = q 2VA (3.85)
g nz max 1
On the other hand, the turn radius can also be written as a function of the lift coecient:
2
rf = q  VA2 2 4 (3.86)
g ( 2 S=W ) CLVA 1
and thus the minimum achievable value for the maximum lift coecient becomes:
VA 2
(rfmin )CL = q  (3.87)
g ( 2 S=W )2 CL2 max VA4 1
Version: 1
57 Date: April 4, 1997
GARTEUR/TP-088-27

both functions 3.85 and 3.87 can be plotted in a diagram, see gure 3.18. In this gure the
area of achievable turn radii at given airspeeds is visible for a maximum lift coecient of
CLmax = 2:75 and a maximum load factor of nz = 1:15. The latter is smaller than the
maximum structural load factor to allow for a margin, since the ap setting is maximum.
Furthermore, the factor is limited for passenger comfort. The point where both curves

2500

2000 Achievable radii

1500
CLmax -limit
rf (m)

nz -limit
1000

(rfmin )abs
500

0
0 20 40 60 80 100 120

VA (m/s)
Fig.3.18 Limitations on the turning radius

intersect indicates the absolute minimum turning radius achievable:


W
(rfmin )abs = g  SC nz max (3.88)
n 2
2 Lmax z max 1
The airspeed at this point is:
s
VA = C nz max (3.89)
Lmax 2 S=W
Substitution of nz = 1:15 and CLmax = 2:75 gives:

(rfmin )abs = 9:81 1:117 :7 104 1:15 = 555 m (3.90)


225=2 260 2:75 1:152 1
and for the airspeed:
s
VA = 1:15 117:7 104 = 55:6 m=s (3.91)
2:75 1:225=2 260
For the thrust it is more dicult to give analytical bounds. For a derivation the reader is
referred to [7], an excellent reference on analytical performance analysis.
Version: 1
Date: April 4, 1997 58
GARTEUR/TP-088-27

3.3.4 The maximum turning rate


The other turning performance parameter is the maximum achievable turning rate. This
rate is also limited by available thrust, the maximum lift coecient and the maximum
load factor. The turning rate is the most important parameter, because in practice turns
are mostly characterized by their rate. This turning rate is namely the easiest measurable
turning parameter.
The turning rate is characterized by:
_ = rV (3.92)
f
By substituting equation 3.84 is obtained:
q
g n2z 1
_ = V (3.93)
For a given speed the rate increases monotoneously with increasing load factor, thus the
maximum rate is achieved at the load factor limit:
q
g n2z max 1
(_ max )n = V (3.94)
In the same way the parameter can be characterized as a function of the lift coecient:
(no wind; V = VA )
g r 
_ = V CL2 ( 2 S=W )2 VA4 1 (3.95)
A
The maximum turning rate is achieved at the maximum lift coecient:
r
(_ max )CL = V CL2 max ( 2 S=W )2 VA4 1
g (3.96)
A
Both functions have been drawn in gure 3.19. Like for the turning radius in this gure
an area of achievable rates can be found. The absolute maximum achievable turning rate
is again at the point where both curves intersect:
s 2
nz max 1 r 
(_ max )abs = g nz max CLmax 2 S=W (3.97)
For CLmax = 2:75 and nz = 1:15 the rate becomes:
s r
(_ max )abs = 9:81 1:151:15 1 2:75 1:225
2 260
2 117:7 104  57:3 = 5:7 deg=s (3.98)
This occurs at the same airspeed as for the minimum radius:
s s
VA =  SC n z W = 11::15 117:7 104 = 55:6 m=s (3.99)
225 260 2:75
2 Lmax 2
The maximum allowed turning rate is just below that of a rate-II turn (6 deg =s). In the
evaluation trajectory (see chapter 4 in [2]) there is a rate-I turn (3 deg =s) at a speed of
80 m=s. This can be achieved easily according to the gure.
Version: 1
59 Date: April 4, 1997
GARTEUR/TP-088-27

10

7 CLmax -limit
6 (_ max )abs
_ (deg/s)
5
nz -limit
4

2 Achievable rates
1

0
0 20 40 60 80 100 120

VA (m/s)
Fig.3.19 Limitations on the turning rate
3.3.5 Climb performance in turns
In section 3.1.1 the following equation has been derived: (the acceleration is set to zero,
cos  1 and wind is not considered here)
sin = WT + CD n  T + n (3.100)
CL z W gl z
During a turn in the rst place the glide angle ( gl = CD =CL ) and the load factor
(nz = 1= cos ) change. If the turn is performed in the horizontal plane, thrust has to
compensate for extra drag to maintain zero ightpath angle.
As long as the throttles do not saturate, extra thrust is available for climbing. If the aircraft
goes from horizontal ight into a turn, the available power for climbing is decreased by
the increased load factor and the changed glide angle. This can in principle be shown in
a Korhammer diagram, although a little cheating is necessary.
In a Korhammer diagram the aerodynamic forces are directly related to the lift-drag polar.
In horizontal ight the aerodynamic, thrust and gravity forces are in the same vertical
plane, so that the symmetric force geometry can be simply completed in the diagram,
including other ightpath parameters ( and V_ ). The problem is that the aerodynamic
axis system is rotated with respect to the ightpath axes about the xF -axes over the bank
angle . The lift, drag and the thrust are (approximately) in the rotated xA -zA plane,
while gravity and the ightpath terms remain in the vertical xF -zF plane.
The best way to modify the diagram is by rotating the vertical xF -zF plane into the xA-
zA plane. Equilibrium along the xF -axis can now be constructed directly (for stationary
ight):
T D = W sin (3.101)
Version: 1
Date: April 4, 1997 60
GARTEUR/TP-088-27

Vertical equilibrium (L = nz W ) can be constructed by stretching the weight vector


with a factor nz . Vertical and horizontal equilibrium are related by projecting the total
aerodynamic force RA on an axis parallel to the xA -axis through the end point of the
weight vector. An example is given in gure 3.20. It should be noted that the diagram in
CL
α

γgl Lift Drag polar

CR CD

= turn
= hor. flight

RA W
γgl
γ

xf

T -n z γW
projection

z a z f’

Fig.3.20 Modi ed Korhammer diagram for a steady turn

this way has become somewhat complex and therefore loses most of its functionality. From
equation 3.100 the e ect of the glide number on the climb angle is visible. If the aircraft
is ying on the right side of the power curve, the necessary increased lift coecient will
result in a smaller glide number (in absolute sense). This slightly compensates decreased
ightpath performance due to the increased load factor. This is for example the case in
the diagram of gure 3.20. If the aircraft on the other hand is operating on the backside of
the powercurve, the glide number will increase. This is an extra performance decreasing
term.

Using equation 3.100 and the performance diagram 3.16 maximum ight path angles have
been calculated for four bank angles as a function of the airspeed, see gure 3.21. Of
course the climb performance degrades considerably at higher bank angles.

3.3.6 Control de ections in co-ordinated turns

In a steady co-ordinated turn expressions for rudder and aileron de ections required for
equilibrium can be derived as functions of the yaw rate. The equilibrium equations for
Version: 1
61 Date: April 4, 1997
GARTEUR/TP-088-27

15

 = 0 deg
10
 = 25 deg

 = 45 deg
(deg)

 = 60 deg
0
h = 0m
xcg = 0.23c
120000 kg
−5
50 60 70 80 90 100 110 120 130
Vtrim (m/s)
Fig.3.21 Maximum ightpath angle vs speed in turn
the co-ordinated turn can approximately be written as follows:
CY = Cy + CyR R + Cyr 2rbVA = 0
rb
Cn = Cn + CnR R + CnA A + Cnr 2VA = 0 (3.102)
rb
Cl = Cl + ClR R + ClA A + Clr 2VA = 0
In the RCAM model Cyr and CnA are zero. Solving for R and di erentiation gives:
2 3
 Rrbss  = CCnr 664
d CY
CYR
77
5 (3.103)
d 2VA nR CY CnR Cn
Using the values that can be found in table 2.2 can be calculated:
" #
d 0
 ss  = 0:092
R : 495 1 : 6 = 5:38  1:24 = 6:7 (3.104)
d 2rbVA 1:6 00::24092 0:118
Cn substituted here also depends on the angle of attack. For a for the RCAM common
value of 3 deg is chosen:
Cn = Cn0 + Cn == 0:147 0:5603=57:3 = 0:118
The expression between brackets in equation 3.103 is usually close to one, so that the
rudder de ection is mainly used to overcome the yaw rate damping Cnr , which acts in
opposition to the established yaw rate. It is clear that the de ection is inversely propor-
tional to the rudder control power, CnR . A positive yaw rate demands negative rudder,
so that the pilot will have to apply a rudder pedal de ection toward the direction of the
turn.
Version: 1
Date: April 4, 1997 62
GARTEUR/TP-088-27

If the term between brackets in equation 3.103 is abbreviated to kY the steady state
rudder de ection becomes:
Rss = CCnr kY 2rb
VA (3.105)
nR
Substitution in the yawing moment equation gives (CnA = 0):
Cn CnR CCnr kY 2rb
V + Cnr 2rb
VA = 0 (3.106)
nR A
Thus the sideslip angle is:
C (1 k )
= nr C Y 2rb VA (3.107)
n
And with substitution of RCAM aerodynamic derivatives and geometry (see table 2.2
and 2.1):
= 0:495(1 0:118
1:24) rb = 1:0 rb
2VA 2VA (3.108)
Thus for a rate I turn (r = 3 deg =s) at a speed of VA = 80 m=s:
Rss = 6:7  23 80
45 = 5:7 deg (3.109)
and:
= 1:0 23 80
45 = 0:8 deg (3.110)
In the problem formulation for the RCAM ight control benchmark ([2]) no lateral ac-
celeration measurement is available. Since is small in a co-ordinated turn, one way to
obtain approximate co-ordination is to drive the sideslip angle to zero using an integrator.
This integration is necessary anyway in order to obtain the required zero sideslip in an
engine failure case. However, for the RCAM there is no solution for equation 3.102. From
the rst equation follows that the rudder de ection is zero, since CYr = 0. On the other
hand, for the RCAM also CnA is zero, so that the yawing moment equilibrium cannot be
ful lled. A solution can only be found if the bank angle is changed, so that the turn is no
longer own co-ordinated. Therefore, a better way to obtain turn co-ordination is driving
the di erence tan  V =g _ to zero.
Finally, the required aileron de ection can be calculated from the roll moment equilibrium:
C + ClR R + Clr 2rbVA
A = l C (3.111)
lA
Thus:
0:205 0:8 + 0:032 5:7 + 0:215 23 80
45
A = 0:088 = 1:9 deg (3.112)
Version: 1
63 Date: April 4, 1997
GARTEUR/TP-088-27

3.4 The e ect of steady wind


In this section the e ects of di erent sorts of wind disturbances are discussed using point-
mass approximations and steady state analysis.
Gust elds induce, compared to the actual airspeed, only small amplitude speed changes
and therefore do not bring the aircraft into new operating points. The analysis of the e ect
of gusts can be carried out based on linear system analysis. However, this does no longer
hold for low frequency large amplitude wind elds like windshears and other atmospheric
fronts. These wind elds have a large impact on the energy state of the aircraft. The
consequence is that large amplitude deviations from equilibrium occur which can only be
compensated by major control actions. To quantify these e ects it is therefore necessary
to use the nonlinear equations of motion of the system.
It is allowed to assume that the in uence of large amplitude wind elds with small wind
gradients mainly a ect the force balance rather than the moment balance of the aircraft.
Therefore it is sucient to consider only the point-mass nonlinear force equations to study
their e ect on aircraft motion.
This again means for the description of the steady state system behaviour, that moment
equilibrium after disturbances is restored rapidly, so that the angle of attack returns to its
initial value. The consequence is that the aerodynamic trim point with gl and airspeed
VA remains almost unchanged. Only the vertical component of steady wind directly a ects
the force equilibrium via w . This can easily be veri ed from the Korhammer diagram
( gure 3.3) and from:
_ + sin = T D L w = T +  CD +  n
ax = V (3.113)
g g W W CL w z
From the remaining components there is only an indirect e ect through:
VA = V VW (3.114)

3.4.1 Tail wind

In the case of tailwind, which corresponds to a positive increment of Vw , with xed gl


and stationary VA , the inertial speed V increases and in second instance w is also slightly
a ected via sin . The reaction of the inertial speed V to a stepwise change in Vw is
delayed, so that the airspeed VA will react stepwisely, thus:
Vat=0 = Vw (3.115)
This means that CR and have to increase in rst instance. Since the system is supposed
to be statically stable, the angle of attack and the airspeed have to return to their equi-
librium point. This means that the aircraft has to accelerate at the cost of its available
potential energy; Vss = Vw . The disturbance is therefore completely transferred into
Version: 1
Date: April 4, 1997 64
GARTEUR/TP-088-27

VA at t = 0 and has been exchanged into V in steady state. Meanwhile the aircraft ac-
celeration leads to a loss of altitude at a constant energy level while the ightpath angle
returns to its equilibrium point, because the force equilibrium returns to its initial state.
3.4.2 Vertical wind
The angle of attack due to wind is given by: w = WVZE A . The attitude angle  is given by:
 = + + w . Since  and can only respond to a stepwise change in WZE with a lag
in time, the initial response is only characterized by a change in angle of attack given by:
 t=0 = w = WVZE (3.116)
A
The angle of attack, the airspeed VA and the glide angle return to their equilibrium values
and hence also the total aerodynamic force RA .
If the thrust is constant, the force polygone of gure 3.4 will return to its initial position.
This is only possible if the ightpath speed vector V is rotated in opposite direction (in
earth co-ordinates) with respect to w . In steady state this results in:
 ss = w = WVZE (3.117)
A
so that the total aerodynamic ightpath angle a remains unaltered. Since a and both
remain unchanged, also the attitude angle  does not change:
 = + + w (3.118)
As can be seen from equation 3.117, the ightpath angle decreases due to the vertical
wind, while a does not change (steady state). If the aircraft has to be in equilibrium on
a pre-speci ed inertial ightpath angle ( ), trimming must be done for a = + w . This
is described in section B.2.
3.4.3 Cross wind
A step in cross wind will result in a sideslip angle:
w = WVY E (3.119)
A
For the track angle holds:
 = + + w (3.120)
Like in the case of vertical wind, the angles  and can only respond with a time delay,
so that the sideslip angle will respond stepwisely:
t=0 = w (3.121)
Due to wind vane stability (Cn ) the aircraft course will change until has become zero,
thus f = w . During the transient motions the aircraft will roll due to Cl , so that the
lift vector is rotated. This will result in some decrease in altitude.
Version: 1
65 Date: April 4, 1997
GARTEUR/TP-088-27

In steady state the track angle has become:


 = + f
where has returned to its initial value. To compensate for lateral wind in order to stay
on a demanded ground track, the heading has to be decreased with f . This is done in
the RCAM trim routine, see section B.2.
3.4.4 Time responses
For the three di erent wind directions simulations have been carried out with the nonlinear
and a linearized RCAM model. The gures can be found in appendix F. The comments
in this section are limited to the nonlinear responses.
From gure F.5 can be seen that VA changes directly, but returns to its initial value, while
the inertial speed V increases with 1 m=s (see equation 3.115). All other outputs return
to their initial values, except the altitude.
In gure F.7 the responses to a vertical windstep can be found. The angle of attack
responds immediately, as predicted in equation 3.116. All variables return to their initial
values, except ; the ightpath angle has a negative steady state value, as indicated
in 3.117.
Finally, in gure F.6 the responses to cross wind can be found. responds immediately
(equation 3.121) and returns to zero after some time. Only the ightpath track angle ()
has a steady state deviation of 1=70  57:3 = 0:82 deg.
Version: 1
Date: April 4, 1997 66
GARTEUR/TP-088-27

3.5 The e ect of windshear


Windshears W_ XE = uwx V , W_ ZE = wwx V have strong in uence on the ightpath and
stability of the aircraft motion. Even though these gradients usually occur over limited
distances and therefore have only e ect over a short period of time, it is important to
know the quasi-stationary reaction of the aircraft. In this section only longitudinal e ects
will be considered.

3.5.1 The F -factor

The e ect of windshear on the aircraft performance can be expressed by the F -factor.
This factor is derived from the e ect of wind on the total energy of the aircraft relative to
the airmass, see for example [17] and [18].
In gure 3.22 some of the parameters used in this section are visible. For symmetric ight
xB

θ WX E VA
WZ E α
Vw
T
V γa
xV γ

zB zV

Fig.3.22 De nition of wind vectors

the force equation along the xA -axis can be written as:


V_ A = T cos D sin W_ XE cos + W_ ZE sin (3.122)
g W W a g a g a

The vertical speed can be calculated from:

h_ = VA sin a WZE (3.123)

with a =  , the elevation angle of the airspeed vector relative to the inertial axes.
Version: 1
67 Date: April 4, 1997
GARTEUR/TP-088-27

In equation 3.15 the expression for the aero-kinetic energy can be found. Dividing this
equation by the weight W , the speci c aero-kinetic energy is found:
2
Eas = V2Ag + h (3.124)
Di erentiating this expression:
E_ as = V_A VA + h_
g (3.125)
Substituting equations 3.122 and 3.123 gives:
" #
_Es = VA T cos D W_ XE cos a + W_ ZE sin a WZE (3.126)
W g g V A
The rst term between the brackets is the speci c excess power, as derived in section 3.2.1
(with cos  1). If there is no wind and all excess power is used to increase the potential
energy, thus V_ A = 0, equation 3.122 can be written as:
sin = sin = T cos D
a W (3.127)
The three wind terms in equation 3.126 describe the impact of the windshear on the
aircraft energy state; these may be combined into a single term, the F -factor:
_ _
F = WgXE cos a WgZE sin a + WVZE (3.128)
A
The e ect of windshear on the aircraft performance can thus be expressed as an e ective
reduction in speci c excess power due to horizontal and vertical shears and downdrafts.
Regions where F is negative are performance increasing shears and regions where F is
positive are performance decreasing. For the RCAM the climb performance at an airspeed
of VA = 70 m=s and a mass of m = 120000 kg with maximum thrust setting is completely
cancelled by an F -factor of 0.21, as appeared in section 3.2.3, equation 3.58.
The F -factor along a reference ightpath for the windshear model implemented in the
RCAM simulation environment (see section 2.3.3) is given in gure 3.23. The ightpath
starts at h = 700 m and goes straight through the core of the wind eld with a ightpath
angle of 3 deg. The encountered windspeeds can be found in gure 2.14.
As can be seen from this gure, the contribution of the downdraft is very important.
Since it is positive, it has a performance decreasing e ect. The e ect of the horizontal
shear is also signi cant, mainly when the head wind suddenly becomes tail wind. Over a
short distance the the F -factor is larger than 0.21, thus if the aircraft is own along this
ightpath, it will not be able to maintain both airspeed and ightpath angle. If priority
is given to airspeed, the aircraft will deviate from its ightpath and vice versa.

The F -factor gives an indication of how much thrust is required to y safely through the
wind eld. For the required thrust to compensate for the downdraft can be written:
T = WZE (3.129)
W VA
Version: 1
Date: April 4, 1997 68
GARTEUR/TP-088-27

0.4
F−factor
0.35
WXEdot−contrib
WZEdot−contrib
0.3
WXE−contrib

0.25

0.2
F−factor

0.15

0.1

0.05

−0.05

−0.1
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
XE (m)

Fig.3.23 Winds shear F -factor


Unfortunately, wind is not measurable and therefore can only be derived using a combi-
nation of airdata and accelerometer outputs.
Compensation for the W_ XE term must be done carefully. In most situations this term is
intially negative, thus for compensation thrust must be decreased. However, ying through
the core of a windshear eld the shear term may rapidly become positive, so that thrust
has to be increased quickly in order to maintain airspeed. However, engines usually have
a large time constant, so that not enough time may be left.

In [12] another relation between the aerodynamic ightpath angle and the wind gradients
is derived.
Starting with equation 3.122 and assuming that the aircraft ies at a constant airspeed in
a rectilinear ight through a wind shear eld and assuming that the angles , and a
are small, this expression can be written as follows:
_ _
0 = T W D a WgXE + WgZE a (3.130)
Since the airspeed is constant, the speci c excess power in constant wind conditions would
be used for an aerodynamic ightpath angle a0 . The in uence of the windshear can now
be characterized through:
a0 W_ XE
a = g (3.131)
1 W_ ZE
g
This equation shows how the aerodynamic ightpath angle depends on the windgradients if
the airspeed does not vary. a is constant when the gradients are constant. The ightpath
angle is given by:
= a w (3.132)
Version: 1
69 Date: April 4, 1997
GARTEUR/TP-088-27

For a vertical windshear with V_ A = 0 is obtained:


a = a0 (3.133)
1 W_ gZE
it can also be shown that (for WXE = 0):
_
_ = _ w = WgZE (3.134)
Note that a constant airspeed implies:
V_ = W_ XE (3.135)
and
_ = _ + _ + _ w (3.136)
The speed on the ightpath is increasing with the windgradient and may grow rapidly.
3.5.2 Energy altitude reserve
It is easy to estimate the e ect of windshear with the energy approach. If an aircraft
enters a wind shear area in a trimmed condition and thrust is held constant, then the
speci c power error in terms of thrust, is proportional to the the F -factor:
T V  W + VA W_  V F = h_ (3.137)
W A ZE g XE A Eaerr

(the e ect of W_ ZE is assumed to be less important compared to the other terms, see also
gure 3.23). Like in equation 3.11 a vertical energy altitude rate, h_ Eaerr , is introduced;
this time with respect to the airmass. This term can be integrated to a speci c energy
altitude error:
Z t V A

hEaerr = WZE + WXE d_ (3.138)
t0 g
If it is assumed that all quantities remain constant during the shear then:
hEaerr = WZE  t + VgA W_ XE  t = WZE  t + VgA WXE (3.139)
thus the energy error is proportional to the duration of the vertical wind velocity and
amplitude of the horizontal windshear component. For a typical approach speed of VA =
70 m=s, a vertical wind WZE = 5 m=s and a windshear gradient of W_ XE = 4 m=s2 during
only 3 s, the loss in energy height will be HEA = 5  3 m + 70=9:81  4  3 m  100 m.
The main contribution in this loss is due to the gradient W_ XE .
Now it is necessary to evaluate how much energy reserve the aircraft has for windshear
compensation. A typical approach speed lies at 1:3 Vstall , while the critical airspeed is at
1:05Vstall . For a stall speed of 54 m=s the approach speed is then VA = 70 m=s. The speed
reserve will be equal to: VA = 70 1:05  54 = 13:3 m=s.
Version: 1
Date: April 4, 1997 70
GARTEUR/TP-088-27

The energy height is:


hEa + hEa = h + h + 21g (VA + VA )2 (3.140)
where h is the deviation from the nominal altitude h. The energy height reserve is then:
2
hEa = h + VgA VA + 2VgA (3.141)
which is 104 m for a speed reserve of 13:3 m=s and H = 0 m. A bu er zone under the
glide path is often not sucient, so that it will be necessary to use speed reserve to hold
the aircraft on the ightpath until the are initiation.
3.5.3 Time responses
In this section some aircraft responses to the windshear implemented in the RCAM-
simulation model will be presented, see gure F.10 in appendix F. The windshear consists
of a simultaneous increasing and decreasing headwind and increasing and decreasing tail-
wind in a time period of only 120 s (a). The e ect is severe by a downdraft at the same
time (a). This downdraft has a sharp peak that causes a sudden drop in the load factor
of the aircraft (b). The altitude response (f) is dramatic. The aircraft starts climbing
due to the increasing headwind and therefore increasing airspeed (e), the sudden change
in the sign of the horizontal windshear in combination with the downdraft that decreases
the angle of attack (d), causes the altitude to decrease with 350 m in only 20 s. After the
windshear the lightly damped phugoid mode remains. The windshear period is approxi-
mately equal to the phugoid period of the aircraft at a trimspeed of 80 m=s and therefore
exciting this mode maximally. A nice view of the response is given via an animation in
gure 3.24.
Version: 1
71 Date: April 4, 1997
GARTEUR/TP-088-27

800

700

600

500
altitude

400

300

200

100

0 −1000
0
5000 0
10000
1000
y distance
x distance

Fig.3.24 Aircraft response to windshear


3.6 Aircraft control during engine failure
One of the most important design requirements for the RCAM autopilot is the ability to
cope with an engine failure. For this reason an open loop aircraft analysis will be carried
out.

3.6.1 Control strategies

The requirement on an engine failure is that the aircraft remains controllable in steady
ights at low speeds, where this malfunction is most critical (e.g. during take o ). Fur-
thermore, the aircraft should be able to maintain a desired heading with zero sideslip
angle. Based on these requirements it is necessary to study lateral aircraft dynamics in
order to obtain information like the minimum control airspeed Vmc and the desired roll
angle to maintain zero sideslip heading. Furthermore, it is important to have an indication
of the remaining climb performance in an engine failure situation.
An engine failure, as shown in gure 3.25, is characterized by a yawing moment NT =
T  YCGE (YCGE in FM ) resulting from asymmetric thrust development T at a distance
YCGE from the center of gravity. NT is positive if the right engine fails. It is important
to know the factors in uencing the magnitude of NT since the degree of diculty of
operating with an engine failure is directly related to this parameter. It is assumed that
the inoperative engine does not produce any thrust or drag.
Version: 1
Date: April 4, 1997 72
GARTEUR/TP-088-27

The nondimensional relation for the yawing moment is:


CNT = qNSbT = TYqSb
CGE (3.142)
If steady level ight is considered the vertical force equilibrium is given by:
W = L (3.143)
= CL qS
Elimination of the dynamic pressure q = 21 VA2 gives:
T C yCGE
CNT = W (3.144)
L b
Equation 3.144 shows that the lateral yawing moment coecient due to thrust is propor-
tional to the operative engine thrust T , the distance of the engine from aircraft center
of gravity yCGE and the lift coecient CL (or equivalently, the angle of attack ). The
yawing moment coecient is inversely proportional to the airspeed VA .
Next, the equilibrium equations for the side force Y , yawing moment N and rolling moment
L are considered:
CY = Cy + CyR R + CL  = 0
T y
Cn = W CL b + Cn + CnR R = 0
CGE (3.145)
Cl = Cl + ClA A + ClR R = 0
For simplicity the cross term CnA is assumed to be zero. For the directional control
problem (constant heading) the expression for the required rudder de ection to maintain
an equilibrium ight condition is derived rst. Since there are theoretically three equations
and four unknowns, there is in fact an in nite number of solutions. However, in practice
a pilot will use only one of the following three limiting solutions:
 no sideslip angle, this solution is aerodynamically the most e ective one, since drag
remains along the xB -axis;
 no bank angle, this solution needs a somewhat larger rudder solution;
 no rudder de ection, where equilibrium can be achieved only by increasing the
sideslip and roll angle.
A ight controller will probably nd a solution in between, unless one of the parameters is
explicitly driven to zero in a special engine failure mode or via an integrator in the loop.
The rudder de ection for the zero sideslip condition can be derived from the expression of
the yawing moment:
T yCGE
Rss = W CCL b (3.146)
nR
from equation 3.146 it can be deduced that the required rudder de ection for equilibrium
increases with increasing asymmetric thrust T , with increasing lift coecient CL (increas-
ing angle of attack, decreasing airspeed). The required rudder de ection R is inversely
Version: 1
73 Date: April 4, 1997
GARTEUR/TP-088-27

proportional to the rudder control e ectivity CnR and aircraft mass m. To achieve zero
sideslip some bank angle is needed to provide side force equilibrium. This situation is
sketched in gure (3.25). The bank angle required for zero sideslip angle can be obtained

β=0

y APT

NT T

Inoperative
Engine

W sin Φ

D = Drag in line of flight


β = 0 Minimum Drag

Yδ Nδ
R R

δR


R
δR

W sin Φ

W Equilibrium
W sin Φ = Yδ
R

Fig.3.25 Asymmetric equilibrium for zero sideslip

from equation 3.145:


CyR R
= CL (3.147)

The bank angle  has to remain in most of the situations less than 5 degrees. This can be
calculated for the RCAM. Consider the situation that the right engine thrust is reduced
to its minimum value of 0:10 105 N (idle), while ying at a speed of VA = 80 m=s. The
nominal aircraft mass is 120 000 kg. The required thrust for steady horizontal ight for this
mass and airspeed appears to be (from trimming the model) T = 1:79 105 N , so that the
operative engine at least will have to deliver a thrust of (1:79 0:10) 105 = 1:69 105 N . The
di erence in delivered thrust between both engines is then: (1:69 0:10) 105 = 1:59 105 N ,
causing the asymmetrical thrust moment.
Version: 1
Date: April 4, 1997 74
GARTEUR/TP-088-27

The lift coecient will approximately be equal to ( = 0 = 1:225 kg=m3 and S = 260 m2 ):

CL = 1=2W = 120000 9:81 = 1:16


VA S 1=2 1:225 802 260
2

Furthermore, from tables 2.1 and 2.2 in chapter 2 can be obtained: yCGE = yAPT 1 =
7:94 m, b = 45 m and CnR = 0:092. The required rudder de ection can now be
calculated:
5
Rss = 1:59 10 =(120000 90::81)092
1:16 ( 7:94=45)  57:3 = 17:2 deg (3.148)
The roll angle will be equal to: (with CYR = 0:24)
 = 0:24 ( 17:2) = 3:6 deg (3.149)
1:16
This less than the accepted 5 degrees in practice. Also the rudder de ection is less than
the maximum of 30 degrees. Of course these values calculated here are in steady state,
however, during the transient just after the engine failure these values may be larger.
The required aileron de ection can be calculated from the roll equilibrium:
C +C 
A = l C lR R (3.150)
lA
Thus, with Cl = 0:205, ClA = 0:088 and ClR = 0:032:

A = 0:205 0 + 0:03217:2 = 6:3 deg (3.151)


0:088

The second trim solution can be obtained by setting the bank angle to zero. This situation
is sketched in gure (3.26). To achieve steady heading ight the required rudder for
equilibrium can be obtained via the solution of the lateral force and the yawing moment
equations in 3.145:
T yCGE
Rss = W CL b Cy (3.152)
CnR Cy CyR Cn
this can be also written as:
T yCGE
Rss =  CL Cyb Cn 
W (3.153)
CnR 1 CnR Cy
R
From the last equation it can be seen that the required rudder angle for trim is larger than
in the case of zero sideslip angle.
The required sideslip angle for the zero bank angle can be obtained from the side force
equation in 3.145:
CYR R
= Cy (3.154)
Version: 1
75 Date: April 4, 1997
GARTEUR/TP-088-27

T y APT
NT
Inoperative
Engine

sin γ E
m a xf= W

D = Drag in line of flight


Yδ Nδ
R R

δR

δR Yδ
R

T sin β

Looking into relative wind

Fig.3.26 Asymmetric equilibrium for zero bank angle


The values can be calculated for the RCAM, with CY = 1:6, = 1:6 deg and Cn =
Cn? + Cn = 0:147 + 0:560 1:6=57:3 = 0:131
5
Rss = 1:59 10 =(120000 9:81) 1:16 ( 7:94=45) ( 1:6)  57:3 = 21:9 deg
0:092 ( 1:6) 0:24 0:131 (3.155)
The required sideslip angle is then:
= 0:24121 :6
:9 = 3:3 deg (3.156)
The required aileron de ection follows from equation 3.150:
A = 0:205 3:30+ 0:032 21:9 = 0:3 deg
:088 (3.157)

Finally, equilibrium can be obtained via the zero rudder option. The situation is sketched
in gure (3.27). From the yawing moment equation 3.145 it is possible to derive the
required sideslip angle for the steady heading equilibrium condition:
T yCGE
ss = W CL b (3.158)
Cn
For a positive yawing moment, the sideslip requirement is negative. The sideslip angle is
generally very large, particularly when ying at low airspeeds, when the operative engine
thrust is high or if the static lateral directional stability (Cn ) is poor. The bank angle
Version: 1
Date: April 4, 1997 76
GARTEUR/TP-088-27

T
y APT
NT

Inoperative
Engine

W sin Φ

D = Drag in line of flight


Φ

W sin Φ

T sin β

Looking into relative wind

Fig.3.27 Asymmetric equilibrium for zero rudder de ection


Version: 1
77 Date: April 4, 1997
GARTEUR/TP-088-27

required to balance the sideforces for the zero rudder de ection condition is obtained from
the side force equation:
CY
 = CL (3.159)
For a positive yawing moment, the required bank angle for the zero rudder de ection
condition must be negative. This bank angle is relatively large at low airspeeds. At rst
glance this ight condition might be desirable since the pilot would not have to perform
lateral directional control via his rudder. However, the drag induced by the large sideslip
angle is high. Furthermore, this situation is dynamically dangerous since the vertical
tail is near stall at these sideslip angles. Finally, ying with a large roll angle is highly
uncomfortable for passengers due to the large lateral acceleration. In the case of automatic
control, the lateral directional controller will try to minimize the sideslip and the bank
angle, somewhere in between the zero sideslip and zero bank angle situation. For the
RCAM the equilibrium will be found for:
5
ss = 1:59 10 =(120000 09:131
:81) 1:16 ( 7:94=45)  57:3 = 12:1 deg (3.160)
and:
 = ( 1:6) ( 12:1) = 16:7 deg
1:16 (3.161)
The aileron de ection is:
A = 0:205 ( 12 :1) + 0:032 0 = 28 deg
0:088 (3.162)
This equilibrium condition requires more aileron de ection than allowed: the ailerons sat-
urate at 25 deg. Therefore, at this trim condition the zero-rudder option is not applicable.
It is clear that there is a trade o in lateral directional control during an engine failure.
The control of an airplane may be limited either by rudder or aileron control. The rudder
de ection is minimised if the aircraft is banked into the side of the operative engine, but a
high bank angle may be very uncomfortable for the passengers. Furthermore, to keep the
vertical velocity constant with increasing bank angle, CL must be increased. This results
in an increase in induced drag and an increase in stall speed. A higher sideslip angle will
also result in higher drag. These performance and control considerations determine in fact
the optimum equilibrium ight condition.
3.6.2 The minimum control speed
Having now discussed the di erent trim solutions in the case of an engine failure, the
minimum control speed, a key parameter for safe operation, will be introduced.
The most critical condition for the airplane due to engine failure occurs at high thrust
settings and low airspeeds, as is during take-o . Under these conditions, the lateral
directional control cannot be regained and maintained after a sudden engine failure if
Version: 1
Date: April 4, 1997 78
GARTEUR/TP-088-27

the velocity is below a certain airspeed, the minimum control airspeed (Vmc ). It may be
de ned for a static or a dynamic case. For certi cation purposes the minimum control
speed for the static case is determined by failing the engine prior to approaching the
minimum condition. It may also be considered as a sudden dynamic condition in which
the critical engine is failed at various airspeeds approaching the minimum condition. For
both cases, Vmc may be limited by the lateral or directional control de ections available
to counteract the rolling or yawing moments and/or control forces involved. At any rate,
there will be a di erent static and dynamic minimum control airspeed for each power
setting, con guration and bank angle. Vmc for maximum single engine thrust can be
derived from equation 3.145 with = 0 (the compensating yawing moment may only
come from R ), Timax and Rmax :
Timax yCGE = C  (3.163)
qSb nR Rmax

where the dynamic pressure (q = 21 Vmc


2 ) is the critical variable.
s
Vmc = CTimax yCGESb 2 (3.164)
nR Rmax
For the RCAM Vmc can be calculated by substituting Rmax = 30 deg and Timax =
THimax m120 g = 5710:3  120000  9:81 = 205:4 kN , CnR = 0:092, S = 260 m2 ,
yCGE = 7:94 m, b = 45 m and  = 0 = 1:225 kg=m3 :
s
Vmc = 205:4  103  ( 7:94) 2 = 68:6m=s (3.165)
0:092  45 5730:3 260 1:225
The minimum trim airspeed is the minimum airspeed at which the heading can be main-
tained without pilot control force inputs with the critical engine inoperative. The minimum
trim airspeed depends on the con guration, power setting and bank angle. The minimum
required engine trim is most appropriately applied to an engine out in cruise condition
with power for level ight or climb. These conditions relate to the problems of operating
for relatively long periods during cruise or climb with an engine inoperative. It is desirable
to be able to climb, hands o at maximum range engine out climb speed, and to cruise,
hands o at maximum range engine out cruise speed. Therefore the minimum trim speed
should be smaller than the maximum range engine out speed.

3.6.3 Climb performance with engine failure

Due to an engine failure, the available power decreases, while the drag, and therefore
the required power increases. In the RCAM model drag components due to sideslip and
control de ections have not been modelled. The only increase in drag is caused by the roll
angle. Due to this roll angle the lift vector is rotated, so that total lift has to be increased
to compensate for the aircraft weight. This is thus the same situation as in a turn. Also
for this case a performance diagram can be drawn. From this diagram the maximum
Version: 1
79 Date: April 4, 1997
GARTEUR/TP-088-27

5
x 10
2.5
h=0m
2.4
xcg = 0.23c
2.3 m = 120000 kg
2.2

Max. Thrust
Tmax , D (N)
2.1

1.9  = 20 deg
1.8  = 10 deg
1.7  = 0 deg
1.6
Vmc
1.5
50 55 60 65 70 75 80 85 90 95 100
Vtrim (m/s)
Fig.3.28 Performance diagram for engine failure; bank angles
climb angles can be drawn as a function of the airspeed for di erent roll angles. The
performance diagram can be found in gure 3.28. From this diagram the maximum climb
angle as a function of the airspeed can be obtained. In gure 3.29 the curve for normal
operation has also been drawn for comparison. Also curves for nonzero bank angles have
been added. It is clear that the total performance degradation due to an engine failure is
dramatic. During take-o the aircraft will hardly or even not be able to climb out with
one engine operative and fully extended aps, since the climb performance will degrade
even more due to a considerably higher weight at take-o .

3.6.4 Time responses

To illustrate the e ect of an engine failure, a nonlinear simulation has been carried out
with the RCAM. The aircraft response is plotted in gure 3.30. The right engine fails at
t = 5 sec and thrust is restored at t = 15 s. This gure shows the severe e ect of the
large yawing moment when it is not compensated quickly enough by rudder and aileron
de ections.
The gures F.8 and F.9 in appendix F show time responses of longitudinal and lateral
outputs. The roll angle  reaches a value of more than 50 deg in less than 15 seconds. In
the same time period the heading changes more than 40 deg, while the lateral acceleration
shows peaks of up to 0:15g. In stationary ight the thrust of the engines contribute to
both the forces in xB -direction and the pitching moment equilibrium. At rst instance
the failure of an engine decelerates the aircraft, causing a negative horizontal load factor
and a decreasing speed. The engine failure also introduces a sudden pitching moment,
giving rise to a negative pitch rate (q). Only a few seconds later the coupling between
Version: 1
Date: April 4, 1997 80
GARTEUR/TP-088-27

20

h = 0m
xcg = 0.23c
15
120000 kg
Normal operation
10
(deg)

 = 0 deg Engine failure


0

 = 20 deg
Vmc
−5
50 55 60 65 70 75 80 85 90 95 100
Vtrim (m/s)
Fig.3.29 Maximum ightpath angle vs speed, engine failure
longitudinal and lateral motions takes over. Due to the roll angle the lift vector is rotated
up to 50 deg, so that the component that compensates the weight becomes smaller with a
factor of  1 sin  = 0:36. The altitude decreases approximately 450 m in 20 seconds.
The loss of potential energy is partly compensated by an increase in kinetic energy due to
the increasing speed. After the left engine thrust has been restored, the lightly damped
phugoid remains.
Version: 1
81 Date: April 4, 1997
GARTEUR/TP-088-27

1200

1100

1000

900
altitude

800

700

600

500

400
0
2000
2000 1000
4000 0
−1000
y distance x distance

Fig.3.30 Aircraft response to right engine failure


Version: 1
Date: April 4, 1997 82
GARTEUR/TP-088-27

4 Open loop analysis of the dynamic aircraft model


In chapter 3 aircraft performance is analysed from a quasi-stationary point of view, mainly
using point-mass considerations. In this chapter the dynamic characteristics of the aircraft
around its center of gravity are discussed by looking at linearized state-space models,
transfer functions, time responses and bodeplots.

4.1 The linearized aircraft models


The aircraft model is based on a low speed landing con guration of a large civil transport
aircraft. The nominal mass has previously been de ned as m = 120:000 kg and the
nominal center of gravity location is [0:23 0 0]T c in FM , see table 2.1. The velocity at
the nominal analysis point is chosen at V = 70 m=s and the ightpath angle ( ) is zero.
The trim settings of the controls are given in table C.1.
Linear models are obtained using the numerical trim and linearization routine supplied
with the RCAM-software (section B.2). Except at the nominal analysis point, the model
is linearized with a maximum forward and maximum aft center of gravity location, a
maximum vertical center of gravity position, the maximum and minimum aircraft mass
(see table C.3) and at three di erent airspeeds (60, 80 and 90 m=s). It is important to
have insight in how severe variations of parameters in uence the model characteristics and
which variations should be paid extra attention to in controllers design, in order to make
a sensible trade-o between performance and robustness to these variations.
The mass and center of gravity changes are the most important robustness issues, since
these are usually only approximately known. Large speed variations can be compensated
for via a gain-schedule. However, as a robustness issue, speed variations are interesting
as well, since a controller that performs well over a considerable speed range, reduces the
need for gain-scheduling. In the RCAM design benchmark ([2]) single controller solutions
with constant gains are preferred.

From the linearization a 9-state state space model is obtained. The numerical values of
the state-space matrices of the nominal model can be found in table C.3. In these tables
the states have been rearranged, so that the decoupling between longitudinal and lateral
motions is visible. The interaction term in the A-matrix between  and w_ B is of course
due to the smaller gravity component along the zB -axis when the aircraft has a roll angle.
However, only perturbations around symmetrical ight will be considered, so that  will
be small, or even zero if only longitudinal dynamics are excited. Usually separate lateral
and longitudinal controllers are designed. Because of the interactions, the longitudinal
controller needs additional compensation when the aircraft is turning.
As can be seen from the B -matrix, each of the engines a ects the lateral motions (p_ and r_ ).
Of course the signs of both engines are opposite, so that the interaction is only important
in the case of asymmetrical thrust (e.g. engine failure). The engines are usually not used
Version: 1
83 Date: April 4, 1997
GARTEUR/TP-088-27

to control lateral dynamics and therefore both engines deliver the same thrust during
normal operation.
The eigenvalues of A-matrix sub-blocks for the longitudinal and lateral dynamics are
summarized in table 4.1.

mode real im. !0 (rad/s)   (s)


longitudinal
phugoid -0.9674 10 2 +/-0.1438 0.1441 0.0671 {
short period -0.7330 +/-0.9711 1.2167 0.6025 {
lateral
Dutch roll -0.1632 +/-0.4801 0.5071 0.3218 {
r ! :  1=s 0.0 { { { {
Spiral -0.2859 { { { 3.50
aper. roll -1.1003 { { { 0.91
Table 4.1 Characteristics of the longitudunal and the lateral modes

The characteristics of the eigenmotions are illustrated by time simulations in appendix F.


In gures F.1 to F.7 the results of linear and nonlinear simulations are compared. In
general, the lateral responses (to aileron, rudder, lateral wind) match very well. The
altitude responses however, di er considerably. Apparently, the coupling between the
longitudinal and lateral motions is not captured very well by the linearization.
In general for longitudinal responses holds that the short period matches very well, but
there are small di erences in the damping and the eigenfrequency of the phugoid. This
is mainly due to the dependence of the phugoid frequency on the airspeed, while for the
linear model this frequency remains constant.

The step responses to tailplane and throttle input ( gures F.2 and F.4) are very common
for transport aircraft. The short period is well damped, as can be seen from the pitch
rate and load factor responses and table 4.1. The phugoid mode is clearly dominating and
badly damped. Both modes have di erent physical interpretations. This will be discussed
in the following sections.
The oscillation in the responses due to the rudder and aileron input ( gures F.1 and F.3)
comes from the Dutch roll mode, which is lightly damped (table 4.1). When this mode
is excited, the aircraft is slipping, rolling and yawing, where the roll motion has a certain
phase lag. The characteristics of this mode strongly depend on the aerodynamic derivatives
and moments of inertia, and therefore on the aircraft con guration.
The aperiodic roll mode is dominated by the roll rate. Due to both rudder and aileron
de ections the aircraft rolls with approximately rst order dynamics to a steady state roll
Version: 1
Date: April 4, 1997 84
GARTEUR/TP-088-27

angle. This motion is continued in the spiral mode, which is slightly unstable for many
aircraft. However, not for the RCAM: the spiral mode is stable and has a remarkably low
time constant of 3:5 s (see table 4.1). This is very likely caused by the unconventional
shape of the aircraft fuselage and the extra vertical tail surfaces. For the RCAM spiral
mode the roll angle reaches a steady state value, while the aircraft is slipping somewhat to
the outer side of the turn since and  have the same sign. The altitude is continuously
decreasing, so that the ightpath looks like a spiral.

4.2 Transfer functions and block diagrams


Before carrying on with the analysis of the openloop dynamics of the RCAM, a very
convenient tool for dynamical system analysis is reviewed: the block diagram in relation
to transfer functions or state space representations. Using this diagram the meaning of
some of the terminology used in this chapter may be easier to interprete.
As very simple elementary system is considered as an example:
! ! ! !
y_ =
0 1 y + 0 u (4.1)
v_ a0 a1 v b0
This system describes a mass spring damper system, where the output y is the position
of the mass, the input u is a force, a0 = mk , a1 = mc and b0 = m1 , where k is the spring
constant, c is the damping number and m is the mass.

y v +
1 1 u
s s
b0
- -

a1

a0

Fig.4.1 Block diagram of example system

In gure 4.1 the system of eq. 4.1 is represented graphically in a block diagram. Note that
in section B.1 the actuator and engine models have been implemented in Simulink in the
same way; this allowed easy implementation of (rate) limits and logic.
It is very easy to derive a state-space representation or a transfer function from a given
block diagram. Drawing this diagram from right to the left facilitates this even more: the
variables and system blocks appear in both the diagram and the equation in the same
order. This is safer in the case of multivariable systems, where the order of multiplication
of matrices is important.
From the diagram in gure 4.1 can be seen that there are two integrations with two
internal feedback loops in the relationship between u and y. If both y and v are chosen as
Version: 1
85 Date: April 4, 1997
GARTEUR/TP-088-27

the outputs of this system the transfer functions can be calculated from the state-space
representation using:
H (s) = C (sI A) 1 B (4.2)
Substitution of eq. 4.1 gives:
y(s) = a0 +ab10s+s2 u v(s) = a0 +ab01ss+s2 u (4.3)
Using these transfer functions the initial and steady state response of y and v to a step
input of u can be calculated with the initial and nal value theorems for functions in the
Laplace domain:
lim y ( t ) = lim sy ( s ) = lim sb0 1 = b0 (4.4)
t!1 s!0 s!0 a + a s + s s a
0 1 2 0
If the relation between y and v is only an integration (a0 = 0), the response of y to a step
input goes to in nity, while its derivative, v reaches a steady state value:
lim y_ (t) = tlim
t!1 !1 v(t) = slim
!0
s(sy(s)) = ab0 (4.5)
1
The initial response is an acceleration:
lim y(t) = slim 2
t!0 !1 s(s y(s)) = b0 (4.6)
which does not depend on the internal feedback loops a0 and a1 :
lim v_ (t) = slim
t!0 !1 s(sv(s)) = b0 (4.7)
The same characteristics can directly be read from the block diagram. Because there are
integrations between u and v and y, the states can not directly change due to a step input,
therefore at t = 0s: y = v = 0 and v_ = b0 u. Between u and v there is one integration
so that the initial rate of v is nite. Between u and y there are two integrations, and
therefore the initial response is an acceleration.
If d = n m is the di erence between the number of poles (n) and the number of zeros
(m), then the dth derivative of the output due to a step input is nonzero. d equals the
number of integrations between the input and the output variable and is therefore called
the relative order. The total gain over the connection lines between the input and the
output in the block diagram equals the gain of the initial system response.
The steady state gain can indirectly be obtained from the block diagram. Provided that
the system is stable, the state derivatives will be zero at t ! 1. Therefore the inputs to
all integrators in the diagram are zero. This means that v = 0 (input to second integrator)
and b0 u a0 y = 0 (input to rst integrator), so that in steady state y = ab00 u.
In many cases a block diagram can make it easier to understand input-output relationships,
interactions in a system, controllability and observability of states, (time-scale) separation
of modes etc.
For controller design block diagrams may give insight in which feedback loops via which
control inputs will be e ective or not, since the input-output relationships are more tran-
parant.
Version: 1
Date: April 4, 1997 86
GARTEUR/TP-088-27

4.3 Approximations for the longitudinal modes


Although the equations of motion have been simpli ed considerably by linearization, sep-
aration of longitudinal and lateral dynamics and other assumptions, even more simpli ed
models for the longitudinal modes will be introduced in this section. This is done to obtain
a better physical insight in the modes and their interactions.
4.3.1 The linearized longitudinal model
The longitudinal equations will be considered rst. In stead of the original states q, ,
uB and wB , used in the nonlinear model, another set of states, q, f , V and will be
used. In this chapter mainly zero wind conditions are considered, so that this set can also
be written as: q, , VA and The use of these states is more convenient in the physical
interpretation of the longitudinal modes. The transformation can be written as follows:
x_ = Ax + Bu (4.8)
z = Tx
where z contains the desired set of states and x is the original state vector. With x = T 1 z ,
x_ = T 1 z_ = AT 1 z + Bu (4.9)
thus:
z_ = TAT 1 z + TBu (4.10)
The following state-space realization is obtained:
0 1 0 10 1
B q_ C BB Mq M Mu 0 C B q C
B
B _ C
C BB Zq Z Zu Z C CC BBB CCC +
B
B C
C = B@ Xq Xu X C
(4.11)
@ V_ A X g A B@ V CA
_ 1 Zq Z Zu Z
0 1
BB ME MTH1 MTH 2 C 0 1

BB ZE ZTH1 ZTH 2 C E
CC BB TH 1 CC
B@ XE XTH1 XTH2 C @ A
A TH 2
ZE ZTH 1 ZTH 2
The numerical values can be found in table C.4, while Bode-plots are given in appendix E.
Note that adding the second and the fourth row gives:
_ + _ = _ = q (4.12)
By making a number of additional assumptions, some elements of the analytically derived
models in [6] have been rounded o . For compatibility with these models, some of these
simpli cations will also be applied to the numerically derived linearizations used in this
section. These simpli cations are:
Version: 1
87 Date: April 4, 1997
GARTEUR/TP-088-27

 Zq  1 (in table C.4 Zq = 0:97). This means that there is only a kinematic relation
between and q: the vertical force due to the pitch rate is neglected (in RCAM CZ
has a q-component via equation 2.30);
 Z  0 (in table C.4 Z = 0:007). This means that the vertical component of the
weight vector remains approximately constant when the ightpath angle changes;
 Xq  0 (in table C.4 Xq = 0:05). This is a reasonable assumption for conventional
civil aircraft. Not only is Xq small, but also the pitch rate q is small compared to
V , and ;
 X  g (in table C.4 X = 9:785). For the in uence of the ightpath angle only
gravity is thus taken into account.
The inputs of both throttles will be taken together, so that ZTH = 2 ZTH 1 = 2 ZTH 2 .
The linearized longitudinal equations now become:
0 1 0 10 1 0 1
B q_ C BB M q M M u 0 q
CC BB CC BB E M  M  TH C !
B
B _ C
C B 1 Z Z u 0 C B C B Z  Z  C
C  E
B
B V_ CC = BB 0 X g X g
CC BB V CC + BB X E X TH CC  (4.13)
@ A @ u A @ A @ E TH A TH
_ 0 Z Zu 0 ZE ZTH
A block diagram of this equation can be found in gure 4.2.
4.3.2 Approximations for the short period
For an approximation of the short period, the in uence of VA and to q_ and _ are
neglected (see gure 4.2). In this way the following reduced model is obtained:
! ! ! !
q_ = Mq M q + ME  (4.14)
_ 1 Z ZE E
As can be seen from table 4.1, the eigenfrequency and damping of the short period mode are
considerably higher than those of the phugoid. Therefore the short period oscillations have
almost vanished when the phugoid takes over. There is thus a clear time scale separation
between the modes, so that both can be considered separately for approximations.
The elements of the matrices are mainly determined by the center of gravity location via
the moment derivative Cm . This is discussed in more detail in appendix A.
The characteristic polynomial of the reduced A matrix is:
s2 (Z + Mq )s + (Z Mq M ) = 0 (4.15)
or:
s2 2!0 s + !02 = 0 (4.16)
The eigen frequency therefore equals:
q
!0 = Z Mq M (4.17)
Version: 1
Date: April 4, 1997 88
GARTEUR/TP-088-27

δ TH
αw
Vw

δE

Period
Short
E
Μδ
TH

Mu
Μδ

Z δE


Mq
1
s
q

Zu
Ζ δ TH

αf
.
1
s


-
α
TH

Xα -g
X δE

1
s

Xu


-
VA
-Z u

- Zα
-

Phugoid
1
s
γ

Fig.4.2 Block diagram of the longitudinal state equations


Version: 1
89 Date: April 4, 1997
GARTEUR/TP-088-27

Since the product Z Mq is somewhat smaller, the eigenfrequency is mainly determined


by the static longitudinal stability term M . By substituting the data from table C.4 the
eigenfrequency for the RCAM model becomes:
q
!0 = 0:5916( 0:8598) ( 0:9840) = 1:22 rad=s (4.18)
This is equal to the value found in table 4.1. The terms Cm and Cmq depend on the loca-
tion of the center of gravity (appendix A). If the center of gravity is moved in the direction
of the neutral point, the absolute values of these coecients decrease, and therefore also
the short period frequency decreases. This is con rmed by table E.3 in appendix E. This
appendix contains tables with trim data and bode plots of linearized models.
All coecients are inversely proportional to the aircraft mass. Therefore the short period
frequency will decrease with increasing mass. This is illustrated in table E.5.

The damping equals:


 = 21 pZZ M+ MqM (4.19)
q
In appendix D elements of the state space matices in terms of the aerodynamic coecients
can be found. These were taken from [6]. With a given eigen frequency the damping mainly
depends on Mq and Z and thus on Cmq , Cm _ (equal to zero in RCAM model) and CZ .
The damping for the RCAM models can be checked:
 = 21 0:59161:21 0:8598 = 0:60 (4.20)
This value is also equal to the damping number found in table 4.1. The transfer function
HqE can be calculated from eq. 4.14:
HqE = s2 M(ZE s ++ (M ZE mE Z ) (4.21)
Mq )s + (Z Mq M )
The initial response of q to an elevator step can easily be read from the block diagram in
gure 4.2:
lim q(t) = ME E
t!0
(4.22)
The steady state values can also be found using this gure, but that is somewhat more
complicated, since there are more interactions than in the simple mass-spring-damper
example.
lim q ( t) = lim sq ( s ) 1  = (M ZE mE Z )  (4.23)
t!1 s!0 sE (Z Mq M ) E
The short period, where mainly pitching moment dynamics are involved, is to be con-
trolled by a primarily moment generating device: in this case the elevator or the tailpane.
Although it may have considerable e ect on the moment equilibrium, thrust is primarily
a power source used to control the energy state of the aircraft ( ightpath, speed). During
a pitch motion relatively few energy is dissipated.
Version: 1
Date: April 4, 1997 90
GARTEUR/TP-088-27

4.3.3 Approximations for the phugoid


A rst approximation for the phugoid is obtained by simply taking the lower right sub-
block of the A matrix in equation 4.13:
! ! ! !
V_ A = Xu X VA + XTH  (4.24)
_ Zu 0 ZTH TH
In this equation left and right throttle have again been taken together. The coecients
have been obtained from a numerical linearization routine. In order to get some insight in
the physical meaning of the matrix coecients, this model will also be derived from the
point-mass force equations with respect to the ight path axes, described in equation 3.1.
If there is no wind, thus w = 0 and VA = V , these equations can be written as:
V_ ! D
!
1
!
sin
!
m = + T+ mg (4.25)
V
_ L ( f + iT ) cos
The force equation in xF -direction can be expressed by:
V_ = m1 CD S 12 V 2 + mT g (4.26)
And for small speed and perturbations around the analysis point (0):
2
V_ + @ V_ = m1 CD S( V20 + V0 @V ) + m1 (T + @T ) ( + @ )g (4.27)
Note that is assumed to be constant here, since in the approximation in eq. 4.24 _ = 0.
In the analysis point V_ = 0, CD S V22 + g = T . The equations can be rewritten, omitting
@ -symbols:
V_ = ( 1 CD SV0 )V g + 1 T
m m (4.28)
The same linearization can be carried out for the vertical force equation:
_ = mV 1 C S  V 2 + ( + i ) T 1
L 2 f T mV V g (4.29)
1 C S( V02 + V @V ) + ( + i ) T0 + @T 1 g
_ + @ _ = mV (4.30)
L 2 0 f T mV V0
0 0
With CL S 2 V02 = mg, neglecting the vertical thrust and drag components and omitting
the @ -symbols:
_ = CL S
mV (4.31)
From this linerization is obtained: (with T = 2  m120 gTH )
Xu = CD S m V0
X = g
XTH = 2 m120 m
g (4.32)
Zu = CL S m
ZTH = 0
Version: 1
91 Date: April 4, 1997
GARTEUR/TP-088-27

These coecients can easily be calculated. In the trim condition _ = 0 deg =s, CD and
CL can be read from gure 2.8: CL = 1:5 and CD = 0:22. Furthermore,  = 1:225 kg=m3 ,
g = 9:81 m=s2 , V0 = 70:0 m=s, S = 260 m2 and m = m120 = 120000 kg.
Xu 0:22 260 1:225
= 120000 70 = 0:041

X = g = 9:81

XTH = 2 120000 9:81


120000 = 19:6 (4.33)

Zu = 1:5 260 1:225


120000 = 0:0040

ZTH = 0
The calculated values are indeed almost equal to the numerically calculated coecients,
compared to table C.4. The derivatives in the -equation that turned out to be zero (Zu ,
ZTH ), are not equal to zero in table C.4, but these parameters are less important for a
physical interpretation.

The characteristic equation of the phugoid approximation is:


s2 Xu s Zu X = 0 (4.34)
Thus (with L  W ):
q s p
!0 = Zu X = g S 2g
m CL = V0 (4.35)
This equation shows that the phugoid eigenfrequency (in the approximation) does not
depend on aerodynamic parameters, but only on the velocity. The eigenfrequency for the
RCAM can be checked by substitution of eq. 4.33:
p
!0 = 2  9:81=70 = 0:2 rad=s (4.36)
This is slightly higher than the value found in table 4.1 (where !0 = 0:14 rad=s).
For a higher airspeed the frequency will decrease. This is con rmed by table E.1.
The damping can be calculated from:
 = 21 p XZu X = p1 CCD (4.37)
u 2 L
The term CD =CL is approximately equal to gl , which is proportional to the energy dis-
sipation by the aerodynamic forces. Because the phugoid is basically a periodical exchange
between potential and kinetic energy (this is illustrated later on), damping will thus occur
by this energy dissipation. If the aircraft is ying on the right side of the powercurve,
the ratio CD =CL  gl , and therefore also the phuguid damping, will increase when V0
increases. This is con rmed by table E.1.
Version: 1
Date: April 4, 1997 92
GARTEUR/TP-088-27

For the RCAM nominal model the damping will approximately be equal to:
 = p1 01::22 = 0:1 (4.38)
2 5
Also in this case the approximation is somewhat higher than the value in table 4.1: 0.1 vs
0.07. When deriving the phugoid approximation, only derivatives with respect to V and
have been taken into account. CD and CL are constant in that case, because _ is set to
zero. However, assuming the angle of attack to be constant is not realistic. Therefore the
approximation will be expanded by taking into account. This is done by including the
moment equation with the assumption that the short period oscillations vanish quickly
after perturbations. q and q_ and can then be taken zero:
M + MuV + MTH TH + ME E = 0 (4.39)
Thus:
= M1 (Mu V + MTH TH + ME E ) (4.40)

This includes in the model the fact that the phugoid mode is also controllable by the
elevator via the angle of attack and thus via the short period dynamics, as illustrated in
gure 4.2. Including the in uence of the angle of attack in equation 4.24 gives:
0 1
! !B CC ! !
V_ X g Xu X B XTH XE TH
= V A+ (4.41)
_ Z Zu 0 @ ZTH ZE E

Substitution of equation 4.40 gives:
! 0 
X g X
1 !
V_ = @ 
X u M u M  A V + (4.42)
_ Zu Mu MZ 0
0 X g
 X g
1 !
@ X  TH M  TH M  ME M Z  A TH
Z
MTH M ZE ME M E

The phugoid mode can be interpreted from an energetical point of view. In section 3.1.2
was derived for the speci c energy rate:
E_ s = E_ = V_ + sin = sin (4.43)
V WV g E

In controlling this energy rate (total energy management) there are apparently two degrees
of freedom: the total (speci c) energy rate and the distribution between its components:
sin and Vg_ . This distribution is usually characterized by the speci c energy distribution
rate, the di erence between the acceleration term and the ightpath term:
_
D_ = sin Vg (4.44)
Version: 1
93 Date: April 4, 1997
GARTEUR/TP-088-27

D_ quanti es the change in the distribution between potential energy (altitude) and kinetic
energy (velocity) per unit time.
If there is no external input of energy, for example when the phugoid is initiated by an
elevator input (the short period is assumed to be energy conservative), only potential and
kinetic energy can be exchanged: E_ s = 0 and thus:
V_ = sin (4.45)
g
And for D_ :
_
D_ = 2 Vg = 2  sin (4.46)
If D_ = 0, no energy is transferred between kinetic and potential energy.
The energy household during a phugoid oscillation can be illustrated using Korhammer di-
agrams. In chapter 3 these diagrams have been used for quasi-stationary analysis, however,
they can also be used for studying dynamic behaviour. For that reason a MatLAB-routine
has been written (KORHAMMER.M) that plots the diagram at every time step of a pre-
viously recorded nonlinear simulation. Running this on a PC or Workstation gives a nice
animation. In this section however, these diagrams will be drawn for only nine points on
one sine wave, as illustrated in gure 4.3. In this gure a time registration of the phugoid
excited by a tailplane input is drawn. Figure 4.4 shows an animation with the aircraft at
approximately the same time points.

15

10 h_
3
4
V , h_ (m/s)

5
2
V

0 1 5 9

6
8
−5 7

−10
0 50 100 150

time (s)
Fig.4.3 Phugoid simulation

A registration of the speci c energy rate and the speci c energy distribution rate can
be found in gure 4.5. The reason why E_ s is not equal to zero, is mainly the slightly
Version: 1
Date: April 4, 1997 94
GARTEUR/TP-088-27

1100

1080

1060

1040

1020

altitude
1000

980

960

940

920

900
0 1000 2000 3000 4000 5000 6000 7000
x distance

Fig.4.4 Animation of phugoid


varying drag. When the aircraft is climbing, the speed decreases and therefore also the
drag decreases (provided that the aircraft is ying on the right side of the power curve).
If the thrust is held constant, there is a power excess, resulting in a positive energy rate.
When the aircraft is descending, the opposite occurs.
In the same gure also the V_ =g and sin have been plotted. The energy exchange is
visible from the fact that both terms indeed have opposite signs and are approximately
equal. The korhammer diagrams are presented without labels, therefore in gure 4.6
a characteristic con guration has been drawn with labels added to the elements. The
horizontal components are somewhat exaggerated with respect to the vertical components.
It should be noted that in this gure not speci c but the energy terms (divided by V ) are
given: W sin in stead of sin and W Vg_ = m V_ in stead of V_ =g. The connection point
between the thrust vector and the acceleration term is marked with a dot or a star (?). For
the energetical interpretation one should always look at the projection of the acceleration
and weight vector on the xF -axis.
In the rst gure (corresponding to point 1 in gure 4.3) the airspeed is at its initial
value and the climb rate is minimum. In that situation the magnitudes of the kinetic
and potential energy rates are equal and at their maximum value, while their signs are
opposite. This can easily be seen from the projections of the acceleration term and the
ightpath term on the xF -axis: they overlap, while the interconnection between the thrust
and the acceleration vector is on the zF -axis.
From the de nition follows that potential energy in this case is transferred to kinetic en-
ergy at maximum rate. Therefore the distribution term achieves its minimum value (the
negative sign follows from the de nition of D_ ), as can be seen from the right bar in the
sub gure. In the gure can also be seen that the vertical acceleration is zero, because
Version: 1
95 Date: April 4, 1997
GARTEUR/TP-088-27

0.3

0.2 1 2 3 4 5 6 7 8 9

V_
g
_ E_ s =V; V_ =g; sin

0.1

0
E_ s
V

sin
−0.1
D;

−0.2
D_
−0.3
0 50 100 150

time (s)

Fig.4.5 Speci c energy rate and distribution rate

CL

CD

CL
γ gl

CR
xa VA CD

γ
RA
xv

W
. . begin point γ
thrust vector
E D Xf

dV
. dt
T -m V

W sin γ
za zv

Fig.4.6 Korhammer diagram example with labelled elements


Version: 1
Date: April 4, 1997 96
GARTEUR/TP-088-27

1.25 1.25

0.625 0.625

0.3 0 0.3 0
0 0.125 0.25 0 0.125 0.25

0.2 0.2

0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2

0 1 2 0 1 2

1.25 1.25

0.625 0.625

0.3 0 0.3 0
0 0.125 0.25 0 0.125 0.25

0.2 0.2

0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2

0 1 2 0 1 2

1.25 1.25

0.625 0.625

0.3 0 0.3 0
0 0.125 0.25 0 0.125 0.25

0.2 0.2

0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2

0 1 2 0 1 2

Fig.4.7 Korhammer diagrams at di erent time points during phugoid period (A)
Version: 1
97 Date: April 4, 1997
GARTEUR/TP-088-27

1.25 1.25

0.625 0.625

0.3 0 0.3 0
0 0.125 0.25 0 0.125 0.25

0.2 0.2

0.1 0.1

0 0

−0.1 −0.1

−0.2 −0.2

0 1 2 0 1 2

1.25

0.625

0.3 0
0 0.125 0.25

0.2

0.1

−0.1

−0.2

0 1 2

Fig.4.8 Korhammer diagrams at di erent time points during phugoid period (B)
Version: 1
Date: April 4, 1997 98
GARTEUR/TP-088-27

the projections of the weight vector and the aerodynamic force vector on the zF -axis are
equal. This corresponds with the fact that the ightpath angle has reached its maximum
value. If the thrust vector is below the xF -axis, zF -component of the aerodynamic force
is larger than the zF -component of the weight vector, and thus there is an upward accel-
eration (remember that the weight vector always ends on the xF -axis). This occurs in the
second gure. If the thrust vector is above the xF -axis the opposite holds. The vertical
acceleration in the second gure is is caused by the airspeed that is still increasing due to
the positive sign of the mV_ projection on the xF -axis. Therefore the total aerodynamic
force increases, even though the angle of attack is less than in the rst gure (compare the
points on the lift drag polars in the rst and the second gure). The angle of glide gl does
not change much, so that the larger total aerodynamic force also means a higher drag.
Due to this extra drag some energy is dissipated, which can be seen from the non zero
E_ s-bar. The situation in this gure is half way to the situation in the third gure. There
the xF -projection of the acceleration vector has become zero. If the damping would have
been zero, the ightpath angle would also have been zero, but this is obviously not the
case. While the velocity is going to decrease, the ightpath angle is still slightly negative,
resulting in a negative E_ -term. The distribution rate is almost zero and changes sign:
from now on kinetic energy is ushed back to potential energy again.
The situation sketched in the fourth gure shows the picture half way to the situation in
the fth gure. There the distribution rate is maximum, the acceleration and ightpath
terms again achieve maximum values in absolute sense and have opposite signs. In this
case however, the ightpath angle is positive and the acceleration is negative.
In the seventh gure the airspeed has reached a minimum, while the ightpath angle is
still slightly positive. Due to the lower airspeed, the total aerodynamic force is smaller,
resulting in a downward acceleration. Via the situation in the eighth gure, the picture of
the rst gure returns. However, due to the damping the absolute values of the acceleration
and ightpath angle are smaller than in the rst gure.

Overlooking all pictures, it can be seen that during a phugoid period the acceleration vector
is rotating about the end point of the weight vector, while its length varies. Unfortunately
the variation of the work point along the lift-drag polar is dicult to see, because the
angle of attack does not change much.
Although many remarks made above are quite trivial, this is an example of how a Ko-
rhammer animation visualizes the changing force pattern during longitudinal manoeuvres.
Korhammer diagrams can be used to analyse the performance of the longitudinal controller
in the sense of throttle/elevator co-ordination.

The total energy principles outlined above, are the key to decoupled co-ordination of
elevator and thrust in energy terms, see for example [13]. From equation 4.25 the force
Version: 1
99 Date: April 4, 1997
GARTEUR/TP-088-27

equation in the xF -direction can be taken:


V_ = 1 D + T sin g
m m (4.47)
Rewriting gives:
T = mV_ + mg sin + D (4.48)
or:
!
T =W V_ + sin + D (4.49)
g
From this equation can be seen that the power supplied by the engines is absorbed by
the drag and the manoeuvring terms Vg_ + sin . If the aircraft is ying stationary along a
horizontal ightpath the following will hold:
Treq = D (4.50)
Furthermore, aircraft are usually own near their minimum drag speed, where the power
curve is relatively at, see gure 3.10. Therefore, if the thrust setting is changed with a
term T , most of the extra power will be available for the manoeuvring terms. This can
also easily be seen from a Korhammer diagram and equation 3.14. If there is no wind and
with CD =CL  gl the equation can be written as:
T = n + sin + V_ (4.51)
W gl z g
During manoeuvres the factor nz will normally be approximately equal to -1, thus drag
will mainly change with gl . If the aircraft is ying near its minimum angle of glide (in
absolute sense), this angle will not vary much when the angle of attack changes. Extra
thrust will thus be in favour of the ight angle and acceleration.
From the approximations of the short period and the phugoid it appeared that the elevator
mainly controls the faster innerloop states and via these states the phugoid, while the total
energy of the aircraft remains approximately constant. If the phugoid is excited by for
example an elevator input, the following must hold (with sin  ):
_
 =  Vg (4.52)
The elevator input is essentially energy conservative, because the forces it generates in the
direction of the ightpath can be neglected. Therefore this equation can be written as
follows:
 d  d ( _ =g) !
V
dE = dE (4.53)
T =const T =const
for the design of a longitudinal controller with optimal co-ordination between elevator and
throttle, the considerations above should be exploited. When for example the velocity has
to increase at constant altitude, power from the engine is needed to accelerate the aircraft,
Version: 1
Date: April 4, 1997 100
GARTEUR/TP-088-27

while the elevator should distribute this power input such that indeed all the energy is
used to increase kinetic energy. The same holds when a ightpath angle is demanded at
constant speed. Of course injection and distribution of power have to be performed at
the same speed, which means that the altitude and the speed control loops must approx-
imately have the same bandwidth. An example of ideal elevator/throttle co-ordination in
energy and distribution rate terms is given in gure 4.9.
This control concept is based on point-mass dynamics. Therefore, due to for example en-

γc .
Vc

. . .
Dε = E ε = γε Eε

.
Vε /g γε

. .
Dε = - Vε /g

Fig.4.9 Energy rate and distribution rate errors for optimal elevator-throttle co-ordination

ergy dissipation in the innerloops, always more power is necessary so that the co-ordination
will not be ideal. Furthermore, aircraft do not always y near the minimum drag speed,
so that drag variations may play an important role in the energy management. However,
the total energy concept is a good concept for both analysis and design. Using modern
synthesis methods the co-ordination of the control inputs is preferably left to the opti-
mization algorithms. In this case the concept can be used to analyse the extent of the
achieved decoupling between altitude and speed control. As a way of thinking for design,
the total energy considerations have proven their value in research projects of Boeing and
NASA, see for example [13, 14].
Version: 1
101 Date: April 4, 1997
GARTEUR/TP-088-27

4.4 Approximations for the lateral modes


For the lateral motions it is more dicult to nd suitable approximations, because the
parameters in the equations strongly depend on the aircraft con guration. Furthermore,
the lateral motions occur in two planes simultaneously: the yF -zF and the xF -yF plane,
while there are strong interactions due to asymmetry of the aircraft. In the longitudinal
case only motions in the xF -zF have to be studied, due to aircraft symmetry with respect
to this plane.
4.4.1 The linearized lateral model
Generally speaking, the characteristic equation for the lateral motions contains four roots:
a complex pair due to the Dutch roll, a fast real pole of the aperiodic roll motion and a
relatively slow real pole due to the spiral mode. The linearized lateral equations of motion
can0be written
1 0 as follows: 10 1 0 1
B r_ CC BB N r N N p 0 r N
CC BB CC BB A R CC N  !
B
B _ C B 1 Y Y p g=V 0 C B C B Y  Y  C  A
B
B C=B CB C +B A R C (4.54)
@ p_ CA B@ Lr L Lp 0 CA B@ p CA B@ LA LR CA R
_ 0 0 1 0  0 0
The term 0 comes from the kinematic relation in equation 2.23 and is usually neglected.
Also the term YA can in most cases be neglected. The numerically determined A and B
matrix for the lateral motions can be found in table C.5. The original states are (p r  vb )T .
Only vb is transformed to f by dividing by V0 . Since no wind disturbances are taken into
account, can be used for f .
For the input-output relations between the inputs and the states adopted here, bode plots
have been calculated for di erent masses, center of gravity locations and airspeeds; these
can be found in appendix E.
The roll angle contribution to _ in equation 4.54, g=V0 can easily be checked in the
numerically determined A-matrix: g=V0 = 9:81=70 = 0:14, which indeed complies with
the value in table C.5. In this table also the state (yaw angle) has been added.
An illustration of the interactions in the lateral equations of motion can be found in
gure 4.10. In the following subsections approximations for the lateral modes will be
discussed by separating the planes in which the motions occur: the roll motion in the
yF -zF plane and the yaw motion in the xF -yF plane.
Initially, the roll motion (initiated by an aileron input) will a ect the yaw motion less than
the yaw motion (initiated by a rudder input) will a ect the roll motion, see gure 4.10.
The gains Np and g=V0 in the r_ -equation are less important than the gains L , LR and
Lr in the p_-equation, see table C.5. For this reason it may be expected that the roll
motion can be reasonably approximated with a rst order model. On the other hand, an
approximation of the Dutch roll by only considering the motion in the horizontal plane
will be inaccurate.
Version: 1
Date: April 4, 1997 102
GARTEUR/TP-088-27

δR
βw

δA

yaw motion
R

A

Np
Nr
1
s
r

Yp
-
R

βf
.

g/V0
1
s

β

φ
p
δR

δA

r
A



R

Lr

Lp
1
s
p
Roll motion

θ0
1
s
φ

Fig.4.10 Block diagram of the lateral state equations


Version: 1
103 Date: April 4, 1997
GARTEUR/TP-088-27

In the following sections some approximations for the di erent modes will be discussed.
Since the time scales of the modes are relatively close these approximations are only
suitable to investigate initial behaviour and to obtain approximations for characteristic
numbers (!0 ,  ,  ) in relation to the aerodynamic derivatives.

4.4.2 Approximation of the aperiodic roll motion


For a approximation of the aperiodic roll motion, simply the states p and  are taken from
the linear model:
! ! !
p_ = Lp 0 p (4.55)
_ 1 0 
The transfer function of this mode can thus be approximated by:
HpA = s LAL = lLA 1 +1T s (4.56)
p p R
with the time constant:
 = L1 = =2V SI(b= x
2)2 Clp (4.57)
p 0
For the linear model Lp = 1:1084, so that the time constant is:  = 1=1:1084 = 0:90. As
expected, this complies very well with the value found in table 4.1: 0:91.
The gain of the initial response to an aileron step input is equal to:
lim p_ = slim
t!0 !1 s HpA = LA (4.58)
Due to the short time constant of the spiral mode, the roll rate becomes small only shortly
after the initiation of the roll mode. For the RCAM there is no real time scale separation
between the modes; calculation of a quasi-steady state gain for p does not make sense.
4.4.3 Approximation of the Dutch roll

A very rough approximation can be made by separating roll and yaw motion, see the block
diagram in gure 4.10. The Dutch roll mode can then be characterized by the states r
and :
! ! ! !
r_ = Nr N r + NR (4.59)
_ 1 Y 0
The characteristic polynomial is:
s2 s(Nr + Y ) + (N + Nr Y ) = 0 (4.60)
The eigen frequency is therefore equal to:
q
!0dr = N + Nr Y (4.61)
Version: 1
Date: April 4, 1997 104
GARTEUR/TP-088-27

And the damping:


 = 21 pNNr ++ NY Y (4.62)
r
The eigenfrequency is mainly determined by the static stability of the lateral motion: Cn
in the N term. The damping on the other hand, is determined by the coecients CY
and Cnr . The latter is for the RCAM quite high; therefore the Dutch roll mode has a
high damping number, compared to other aircraft. By substitution of numerical values
the approximated and the actual eigenfrequency of the Dutch roll mode can be compared.
q
!0dr = 0:1791 + ( 0:4556)( 0:1486) = 0:50 rad=s (4.63)
Comparing this to table 4.1, this approximation looks good. Calculation of the damping
gives:
 = 21 0:45560:50 0:1486 = 0:61 (4.64)
This appears to be considerably higher than in table 4.1.
Actually, this approximation only applies well to aircraft with non-swept wings. The
coupling with the roll motion can not be neglected. The amount of coupling between both
motions however, gives important information about the lateral control characteristics of
the aircraft.

4.4.4 Spiral stability

Although it is inaccurate to characterize the spiral mode using a linear model, the lin-
earization can be used to derive a stability condition for this mode.
For studying spiral stability in the rst place an extra constraint is added to the lateral
model: it is assumed that the centrifugal force and the lateral weight component by the
roll angle are in equilibrium (turn co-ordination):
g 0  V0 r (4.65)
This means that the aerodynamic force vector is assumed to be in the symmetry plane of
the aircraft and the component CY is neglected. The homogeneous state equations now
reduce to:
0 1 0 10 1
B r
_ C BB N r N N p 0 CC BB r CC
B
B C
0 C B 1 0 0 g=V0 C B C
B
B C = BB L L L 0 CC BB p CC (4.66)
@ p_ CA @ r p A@ A
_ 0 0 1 0 
The characteristic equation of this this matrix is:
"  !#
s2 Lp Np g  L s + Vg N1 (Nr L N Lr ) = 0 (4.67)
V0 N 0
Version: 1
105 Date: April 4, 1997
GARTEUR/TP-088-27

In this ight condition, this equation has two real poles: from these roots time constants
for the roll mode and the spiral mode follow. For many aircraft these time constants di er
considerably, but this is not the case for the RCAM, as can be seen in table 4.1. This is
mainly due to a high Cl (probably because of the high fuselage) and a high Cnr (possibly
due to extra vertical surfaces on the tailplanes).
It it possible that in some ight conditions the real poles become a complex pair. This
gives rise to a "lateral phugoid". Looking at table E.2 and E.6 the time constants rapidly
get closer to each other when the mass increases and the speed decreases. Trimming the
RCAM at a speed of V0 = 60 m=s and with a mass of 150000 kg gives a conjugate pair of
roots: 0:5801 + 0:0923i: !0 = 0:5874 and  = 0:9876.

The characteristic equation with real roots can be written as:


 1  1  1 1
s+ TRs+ =s +
TS
2 + s + 1 =0
TR TS TR TS (4.68)
Therefore:
1 = g 1 (N L N L ) (4.69)
T T V N r r
R S 0
So that for the time constant of the spiral mode holds:
1 = T g 1 (N L N L ) (4.70)
TS R V0 N r r

The expression between brackets is usually very small. TR is positive, so that for a stable
spiral mode is required:
!
1 (N L N L ) = N L Lr > 0 (4.71)
N r r Nrr N
Nr and L are usually negative and N and Lr are mostly positive (see table C.5), thus
for stability is required:
!
L Lr < 0 (4.72)
N Nr
or:
(Nr L N Lr ) > 0 (4.73)
For the RCAM model can be substituted:
(Nr = 0:4556, L = 1:4900, N = 0:1791 and Lr = 0:4811)
(Nr L N Lr ) = ( 0:4556  1:4900 0:1791  0:4811) = 0:59 > 0 (4.74)
This number is very high. As already noted in section 4.1, the spiral mode is clearly stable
for this trim condition.
Version: 1
Date: April 4, 1997 106
GARTEUR/TP-088-27

4.5 Control capacity analysis


The number of possible feedback con gurations for a ight control system is very large.
At the same time, controller gains are not allowed to be high because of rate limits and
saturation of the controls and because of possible excitation of structural modes. There-
fore, the gains of the input-output relations (in this case transfer functions) to design
controller gains for, should be as high as possible and the the time lag in the responses
(relative order of the transfer function) should be as small as possible. Although this may
sound a little bit like a SISO approach, also for MIMO controller design methods it is
necessary to know which feedback signals will be most e ective. Unfortunately, for the
RCAM ight controller the measurements that may be used are prescribed, see table 2.6.
Looking at initial and steady state gains and block diagrams will give an impression of
the e ectiveness of using these measurements for feedback.
Another very important issue in choosing the feedback con guration is non-minimum
phase behaviour. This aspect will also be studied in this section.
4.5.1 Longitudinal responses
As already seen from Korhammer diagrams, due to internal coupling and secondary e ects
like a thrust moment about the lateral axis, the e ects on initial and steady state aircraft
responses are in many cases completely di erent.

Thrust input
Before calculating the initial and steady state gains, the transfer functions need to be
determined. This can be done easily using the well known Cramer's rule. The transfer
function TH to V is calculated by replacing the u-column of (sI A) with the TH -column
of B and calculating the determinant of the matrix obtained, divided by the characteristic
polynomial of A, which is equal to the determinant of (sI A):

s Mq M MTH 0
s Z ZTH 0
HuTH = jsI 1 Aj 1
0 (X g) XTH g
0 Z ZTH s
3 2
= Xs4TH+sa +s3b+2 sa +s2b+
1 s g (M ZTH MTH Z )
a1 s + g (M Zu Mu Z ) (4.75)
3 2
Some of the terms in this equation have not been expanded, because for the inital and
nal value theorems only the rst and the last terms of the polynomials are needed. Now
the steady state response to a unit step on TH can be calculated:
lim H = M ZTH MTH Z (4.76)
s!0 u TH M Zu Mu Z
If ZTH and MTH would be zero, HuTH is equal to zero, and all power provided by the
engines will be transferred to the potential energy rate. In the case of the RCAM model
Version: 1
107 Date: April 4, 1997
GARTEUR/TP-088-27

however, the MTH -term is positive due to the fact that the engines are situated below
the center of gravity. By the e ect on the moment equilibrium, the energy distribution
changes: some of the engine power is used to accelerate the aircraft.
The transfer function from TH to can be calculated in a similar way as HuTH :

s Mq M Mu MTH
1 s Z Zu ZTH
H TH = jsI 1 Aj 0 (X g) s Xu XTH
=
(4.77)
0 Z Zu ZTH
s3 +c3 s2 +c2 s+M (XTH Zu ZTH Xu )+ x (MTH Zu ZTH Mu )+Z (MTH Xu XTH Mu )
s4 +a3 s3 +a2 s2 +a1 s+g (M Zu Mu Z )
The steady state response to a unit thrust input is:
lims!0 H TH =
M (XTH Zu ZTH Xu ) (X g)(MTH Zu ZTH Mu )+Z (MTH Xu XTH Mu (4.78)
g(M Zu Mu Z )
and if ZTH and MTH are equal to zero:
lim H
s!0 TH
= XgTH (4.79)
The initial response is mainly determined by the relative orders:
 d = 0: direct feed through, TH ! V_
 d = 1: one integration, TH ! V; nz
 d = 2: two integrations, TH ! ; 
Thus:
lim _ = XTH TH = T
V (4.80)
t!0 m
The numerical values for the RCAM model can be found in table 4.2.

Elevator
If only the short period approximation is considered rst, the initial elevator response is
determined by the following relative orders:
 d=0, direct feed through: E ! q_
 d=1, one integration: E ! q (and via ZE )
 d=2, two integrations: E ! ; 
The relative order is of great signi cance for nonlinear inverse controller design. In that
case it is important to know how many integrations there are between control inputs and
the outputs to be controlled in order to nd an invertible relationship between these inputs
Version: 1
Date: April 4, 1997 108
GARTEUR/TP-088-27

Tailplane
outputs: inital gain steady state gain
longitudinal short per.
q_ -1.8667 0.0 0.0
q 0.0 0.0 -0.69
0.0 -2.3 -1.29
 0.0 -2.4 {
V 0.0 330.8 {
0.0 -0.1 {
Thrust
outputs: inital gain steady state gain
V 19.5 172.0 {
0.0 2.1 {

Table 4.2 Initial and steady state gains for step inputs
and derivatives of the outputs. This seems quite straight forward, but secondary e ects
of controls may cause the relative order to be lower than it is intended to be. An example
is the transfer function H E (short period approximation):
  s Mq M ! 1
ME
!
H E = 0 1
1 s Z ZE
= s2 (MZE+(sZ )M
s
q ) + ME
+ (Mq Z M ) (4.81)
q
Due to the ZE factor in the nummerator, the relative order is 1. However, the contribution
of the elevator to the total lift is only a secondary e ect, because the device is primarily
intended to control the pitching moment. Thus designing an inverse controller based on
this unintended single integration relationship doesn't make any sense. For this application
ZE should be set to zero so that the relative order is 2. Otherwise a right half plane zero
is cancelled by the controller, giving rise to internal instability. In a nonlinear sense this
is known as unstable zero-dynamics.

The analysis of steady state behaviour from the short period approximation should in
this case be carried out carefully, because some time after an elevator input the phugoid
component of the response takes over. It is possible to calculate a short time steady state
gain, because there is quite a lot of time scale separation between both modes and the
short period is well damped.
q H = Z M
ME M ZE
E tTsp = slim
!0 qE Z Mq
(4.82)
Version: 1
109 Date: April 4, 1997
GARTEUR/TP-088-27

And for the angle of attack:


= lim H = ME Mq ZE (4.83)

E tTsp s!0  E M Z Mq
The numerical values can be found in table 4.2.
As seen from the phugoid approximation, this mode is also controllable by E via the angle
of attack. Looking at the (approximate) initial behaviour rst:
 d = 0: direct feedthrough, (quasi step response), E ! ; 
 d = 1: one integration, E ! V
The steady state gains can be calculated using the phugoid approximation 4.43:
lim H
s!0 qE
= 0 (4.84)
lim H
s!0 E
= ME Z M
Xu (X + g)Zu (4.85)
Zu Z Mu
lim H = MM Zu Z Mu  ME (4.86)
s!0 E Z ZM
u u M
lim H = M Z Z M  ME Z (4.87)
s!0 V E M Zu Z Mu M Zu
lim H
s!0 E
= ME Zg (XMu Z (X Z Mg)Z)u (4.88)
u u
Again, the numerical values can be found in table 4.2.
The sign of the nominator of the last transfer function may change. This causes nonmin-
imum phase behaviour if it becomes positive, which is directly related to ying on the
backside of the power curve. This will be clari ed in the next section.
4.5.2 Longitudinal non-minimum phase responses
From the initial response V_ = T=m it is clear that the initial response is always positive.
For the steady state gain was found:
lim H = M ZTH MTH Z (4.89)
s!0 u TH M Zu MuZ
These values can be related to the force and moment derivatives using the formulas in
appendix D.
M  =2V02 ISy cCm Z  =2V0 mS CL
Mu  =2V02 ISy cCm0 Zu  =2V0 mS CL0 (4.90)
MTH = zAPT
Iy m120 g ZTH  iT m g
m 120
For moment equilibrium the aerodynamic and the thrust moments need to be in equilib-
rium:
M0A + M0T = 0 or: zAPT T0 + =2V02 ScCm0 = 0 (4.91)
Version: 1
Date: April 4, 1997 110
GARTEUR/TP-088-27

or:
Cm0 = zAPT 1 T
c =2V02 S 0 (4.92)
On the other hand, for steady state condition also the following holds:
T0  =2V02SCD0 (4.93)
and:
Cm0 = CD0 zAPT
c (4.94)
This term is negative if zAPT is positive, like in the case of the RCAM, where the engines
are situated below the center of gravity. From the equations above the following can be
derived:
zAPT =cCL iT Cm
V0S (CL0 Cm + CD0 CL zAPT =c) m120 g (4.95)
The denominator will usually be negative (Cm < 0; zAPT < c), as long as the term
CL0 Cm is dominating. Thus if zAPT and iT are positive, the steady state gain will have
a negative sign, opposite to the sign of the initial response, causing nonminimum phase
behaviour. A consequence of the right half plane zero is that for the design of a SISO
feedback loop V via TH no integrator can be put in the controller, because this new pole
will immediately be attracted by this zero into the right half plane, resulting in instability.

A very common example of nonminimum phase behaviour is the response of the load factor
to an elevator input. This has an easy physical interpretation: for a positive elevator
command the steady state response of will be negative. The change in elevator position
however, will initially increase the total lift of the aircraft, so that the initial response
of the load factor (which is proportional to _ ) will be upward. When the aircraft starts
pitching the load factor will become negative. The order of magnitude of the positive zero
is the same as that of the poles of the short period.
The position of the zero depends on the location of the sensor for the vertical acceleration.
Not only nz has a direct feed through from the elevator, but also q_ due to the pitching
moment that arises from the de ection. When the vertical acceleration is not measured
in the center of gravity, the q_ term will have a contribution of lsensq_, where lsens is the
distance in xM -direction between the sensor and the center of gravity. The total measured
vertical acceleration therefore equals:
zsens = zcg lsensq_ (4.96)
The initial gain from an elevator input to zcg is approximately equal to V0 ZE (see equa-
tion 4.14, with zcg  _ V0 ) and the initial gain of the input to q_ is ME . The initial
response of the vertical acceleration measured at a distance lsens from the center of grav-
ity is therefore:
zsens = (V0 ZE lsens=ME ) E (4.97)
Version: 1
111 Date: April 4, 1997
GARTEUR/TP-088-27

Thus in the case of the RCAM model the sensor needs to be moved from the center of
gravity over: (for the RCAM tailplane derivatives are used)
lsens = V0 ZT =MT = 70( 0:0709= 1:8667) = 2:7 m (4.98)
in order to make the measured acceleration zero. This location is called the percussion
point. If the sensor is positioned more than 2:7 m in front of the center of gravity, the
zero has shifted to the left half plane.
This non-minimum phase behaviour can of course also be avoided by relocating the con-
trols. An example is the use of canards. In the Lockheed 1011 ightpath control is
augmented with direct lift devices on the wing during the approach to land, see [15].

Another, slower nonminimum phase response arises when the aircraft is ying on the
backside of the power curve. This can be seen from the steady state gain of the ightpath
angle:
ME (X Z Z (X g))
ss = gM (4.99)
Z u u
u
Furthermore:
W  =2V02SCL0 (4.100)
Substituting from appendix D:
Zu  S (C + C )
m L0 0 D0
Xu  S
V0 m (CD0 0 CL0 ) (4.101)
Z  =2V0 mS (CL + CD0 + 0 CD )
X  =2V02 mS (CD 0 CL )
gives:
=2V02 mS CmE
gCm CL0 (CD0 CL CL0 CD ) (4.102)
The left ratio is normally negative. When ying on the right side of the power curve, the
right term is positive, so that a positive elevator de ection results in a negative ightpath
angle. This term however changes sign when:
CD0 < CD (4.103)
CL0 CL
see also section 3.1.3. For the linearized RCAM models this phenomenon can be illustrated
by plotting the speci c zero at di erent linearization points, see gure 4.11. At the airspeed
of VA = 70 m=s the zero is very close to the origin. This is also visible in gure E.1, where
at lower frequencies the frequency response of to T shows di erentiator-like behaviour.
Indeed, the ightpath angle has become independent of the elevator position in steady
state.
Version: 1
Date: April 4, 1997 112
GARTEUR/TP-088-27

0.1

0.08

0.06

0.04

90 70 55
imag 0.02

−0.02

−0.04

−0.06

−0.08

−0.1
−0.1 −0.08 −0.06 −0.04 −0.02 0 0.02 0.04 0.06 0.08 0.1

real
Fig.4.11 RHP-zero due to backside operation
This right half plane zero will give troubles when the ightpath angle is held constant while
the speed is not controlled. If in that case the airspeed may decrease due to for example
sudden head wind, the drag increases as can be seen from the power curve (see gure 3.10)
decelerating the aircraft even more. The ight condition will therefore not return to its
initial operating point. When an aircraft is in low speed con guration in the approach
to landing, it might enter the backside of the power curve. In that case a conventional
ight controller designed using a SISO, one loop at a time approach (autopilot: ightpath
via elevator, autothrottle: speed via throttles) will no longer work properly. If one of the
systems is switched o , the other will cause instability of the aircraft. This is usually
solved by interchanging the controls of the ightpath angle and speed loops: speed is then
controlled using the elevator, while is controlled via the throttles.
In fact it is not a real stability problem, since the uncontrolled aircraft will nd equilibrium.
When the control problem is attacked as a multi-variable problem, this speed instability
is completely avioded.

4.5.3 Lateral responses

As already seen in section 4.4 the lateral motions are more dicult to characterize. After
a lateral control input the aircraft will get a roll angle and the spiral mode will dominate
the response. Since the equations have been linearized for zero roll angle, it does not make
sense to calculate steady state responses using the transfer functions, since the motions
can no longer be considered linear. However, some remarks can be made using the block
diagram in gure 4.10.

Aileron input
For the initial response holds (see gure 4.10):
 d = 1, (1 integration): A ! p, (A ! r via NA )
Version: 1
113 Date: April 4, 1997
GARTEUR/TP-088-27

 d = 2, (2 integrations): A ! .
For the initial response of the yaw angular acceleration holds:
lim r_ = NA A
t!0
(4.104)
Positive NA may give rise to non-minimum phase behaviour (adverse yaw), since the
initial yaw acceleration is opposite to the intended yaw rate. In the case of the RCAM,
in the nominal trim condition holds NA = 0:0135 < 0 . It is not unlikely that in other
trim con gurations the number becomes positive.

Rudder input
The e ect of a rudder input can be studied from the complete block diagram, see g-
ure 4.10. The following holds:
 d = 1, (1 integration), R ! r, (R ! p, R ! )
 d = 2, (2 integrations) R ! , (R ! ).
The response of after only one integration is due to the term YR . This is the same kind
of side e ect as ZE : a force is generated by a device that principally controls a moment.
Since LR = 0:2223 > 0, the initial roll acceleration will be in opposite direction compared
to the direction of the steady state roll angle.
Version: 1
Date: April 4, 1997 114
GARTEUR/TP-088-27

4.6 Aircraft responses to disturbance inputs


In this section the aircraft responses to wind inputs will be discussed. This is done for
longitudinal, lateral and vertical wind steps.
4.6.1 Longitudinal wind input
In gure F.5 (appendix F) a nonlinear simulation of responses to a unit step on the WXE -
input can be found. A nonzero steady state response to a wind input indicates that the
output needs disturbance rejection at lower frequencies if it is to be regulated. In many
cases it indicates where a controller needs an integrator in order to obtain good tracking of
a reference signal. On the other hand outputs with nonzero initial response are suitable as
feedback signals for rejection of higher frequency disturbances, because from these signals
immediate information about the disturbance is obtained. In the case of the RCAM model
this will be of little use, because all actuators have relatively low bandwidth. However
when faster devices like spoilers are available, it is possible to achieve rejection of higher
frequency gust inputs.
A block diagram with the WXE input can be found in gure 4.2 (since the track angle is
zero, WXE is felt as a tailwind). From this gure and the responses in gure F.5 can be
seen that nz and VA contain direct feed through terms. On the other hand, there is one
integration between V , and . It is clear that the maximum relative order is only one,
which means that the disturbance in uence is mainly near the outputs. In gures E.7
and E.7 in appendix E bodeplots for the wind inputs can be found.
For VA and V the following holds:
lim VA = WXE
t!0
(4.105)
lim V = Xu WXE
t!0
(4.106)
-symbols will be omitted in the following equations, because in this section only linear
models are considered for small perturbations around an equilibrium point. From sec-
tion 3.4.1 it appeared that a constant wind has no in uence on the steady state response
of and thus the moment equilibrium remains unchanged. Therefore the steady state
response can be analysed using the phugoid approximation. From gure 4.2 follows, with
_ ss = V_ = 0:
ZuVAss = 0 and: XuVAss g ss = 0 (4.107)
Thus:
VAss = 0 and ss = 0 (4.108)
and:
V = VAss + WXEss = WXEss (4.109)
Version: 1
115 Date: April 4, 1997
GARTEUR/TP-088-27

This can also be seen from gure F.5, where indeed VA and the other states return to their
initial values, except V which increases with WXE .
In gure E.7 in appendix E bodeplots for the wind inputs can be found. The observations
made above are con rmed by the rst picture: the relative order is of VA is zero, while all
variables in the gure have zero steady state gain .
4.6.2 Vertical wind input
The e ect can be seen as a disturbance on the angle of attack: w  WZE =V0 , see the
blockdiagram 4.10. From this gure can immediately be seen that , nx and nz respond
with a zero relative order, thus by direct feed through. The nonlinear time simulations
can be found in gure F.6 (section 3.4.4). The steady state responses for all outputs are
zero, except the ightpath angle.
This can also be seen from the block diagram, where for t ! 1 the integrator inputs are
zero. Therefore:
Mq qss + M ss = 0 (4.110)
qss + Z ss = 0
Thus and q need to be zero, which means that the moment equilibrium returns to its
initial situation. For the inputs of the integrators of and V holds:
ZuVAss Z ss = 0 (4.111)
X ss + Xu VAss gss = 0
Thus VAss = V = 0 and  = 0. Furthermore:
ss = fss w ! fss = w (4.112)
= ss ss w ! ss = w
For the initial response the following relations hold:
lim = w
t!0
(4.113)
lim _
t!0 f
= Z w (4.114)
Although the RCAM model has no inputs for wind rates, these are very important in the
case of turbulence. From the time simulations it is clear that the short period mode is
excited via q. As can be seen from the block diagram, there is a feedback loop from to
q_, so that the relative order of a wind input to q is one.
Again these observations can be seen in bode plots as well: see gure E.7.
4.6.3 Lateral wind input
A lateral wind input directly a ects the side slip angle: w = WY E =V0 . For the initial
responses can be written (see the blockdiagram in gure 4.10):
limt!0 = w and: limt!0 _f = Y w (4.115)
Version: 1
Date: April 4, 1997 116
GARTEUR/TP-088-27

From the same diagram also steady state behaviour can be analysed.
pss = 0
Lr rss + L ss + Lppss = 0 (4.116)
Nr rss + N ss + Nppss = 0
Thus pss, rss and ss are zero. Furthermore:
ss = fss wss = 0 ! fss = w (4.117)
Time simulations are discussed in section 3.4.4.
Version: 1
117 Date: April 4, 1997
GARTEUR/TP-088-27

5 Conclusions
In this document the Research Civil Aircraft Model has been discussed. In the second
chapter the equations of motion have been described in a structured way, as used in [6].
On the basis of this model a number of performance and dynamic characteristics have
been calculated. Although many of the formulas can be found in standard text books, it
was very interesting to do the whole range of calculations for one nonlinear aircraft model.
Usually aircraft performance (along trajectories) and dynamics around the center of grav-
ity are considered as two separate disciplines. Of course the second is the most important
for controller design, but it is not possible to design a ight controller without knowing
the performance limits of the aircraft, especially when the controller is designed for a
large part of the ight envelope, and when the controller has to guide the aircraft along
trajectories with turns, climbs, and descents, under severe disturbances.
Using point-mass mechanics and steady state analysis, control strategies have been de-
rived for engine failure, wind elds, windshear, climbing and accelerated ight. At this
point energy considerations and Korhammer diagrams are very interesting concepts to
physically understand (longitudinal) aircraft behaviour and control, including phenomena
like backside operation.
The longitudinal performance and dynamics of the RCAM are representative for a civil
airliner, while the lateral dynamics have strange properties, like a very fast spiral mode.
In fact, in spite of many simpli cations, the most interesting ight situations are covered
by the model, including backside operation and stall.
Version: 1
Date: April 4, 1997 118
GARTEUR/TP-088-27

References
[1] Flight Mechanics Action Group 08. Communication Handbook. GARTEUR FM-
AG08/TP-088-5, second edition, February 1996.
[2] Flight Mechanics Action Group 08. Robust Flight Control Design Challenge Problem
Formulation and Manual: the Research Civil Aircraft Model (RCAM). GARTEUR
FM(AG08)/TP-088-03, second edition, January 1996.
[3] Flight Mechanics Action Group 08. Robust Flight Control Design Challenge Problem
Formulation and Manual: the Research Civil Aircraft Model (RCAM). GARTEUR
FM(AG08)/TP-088-03, third edition, February 1997.
[4] Flight Mechanics Action Group 08. Robust Flight Control Design Challenge Problem
Formulation and Manual: the High Incidence Research Model (HIRM). GARTEUR
FM(AG08)/TP-088-04, third edition, February 1997.
[5] S. Bennani and G.H.N. Looye. RCAM Design Challenge Presentation Document: the
-Synthesis Approach. Technical report, GARTEUR, April 1997. RCAM entry of
Delft University of Technology.
[6] Rudolf Brockhaus. Flugregelung. Springer-Verlag, Berlin Heidelberg, 1994.
[7] G. Bruning, X. Hafer, and G. Sachs. Flugleistungen. Hochschultext. Springer-Verlag,
Berlin Heidelberg, 1986.
[8] F.E. Cellier. Continuous System Modeling. Springer-Verlag, New York, 1991.
[9] H. Elmqvist. Dymola { Dynamic Modeling Language User's Manual. Dynasim AB,
S-22370 Lund, Sweden, 1993.
[10] O.H. Gerlach. Vliegeigenschappen I. Delft University of Technology, Delft, 1981.
[11] X. Hafer and G. Sachs. Flugmechanik (Moderne Flugzeugentwurfs- und
Steuerungskonzepte). Hochschultext. Springer-Verlag, Berlin Heidelberg, 1980.
[12] K.U. Hahn. Beitrage zur Flugleistungsbestimmung beim Startsteig ug unter
Berucksichtigung variabler Windein usse. PhD thesis, TU Braunschweig, 1988.
[13] A A. Lambregts. Operational aspects of the integrated vertical ight path and speed
control system. AIAA Paper 83-1420, 1983.
[14] A A. Lambregts. Vertical ight path and speed control autopilot design using total
energy principles. AIAA Paper 83-2239, 1983.
[15] L.O. Lykken and N. Shah. Direct lift control for improved automatic landing safety
and performance for a large transport aircraft. In Proceedings of the AIAA Guidance,
Control and Flight Mechanics Conference, August 1971.
Version: 1
119 Date: April 4, 1997
GARTEUR/TP-088-27

[16] Donald McLean. Automatic Flight Control Systems. Series in Systems and Control
Engineering. Prentice Hall International (UK) Ltd, Hertfordshire, 1990.
[17] Sandeep S. Mulgund and Robert F. Stengel. Aircraft ight control in wind shear
using partial dynamic inversion. In Proceedings of the American Control Conference,
pages 400{404, June 1993.
[18] Sandeep S. Mulgund and Robert F. Stengel. Target pitch angle for the microburst
escape maneuver. Journal of Aircraft, 30(6):826{832, Nov.{Dec. 1993.
[19] Martin Otter. Objektorientierte Modellierung mechatronischer Systeme am Beispiel
Geregelter Roboter. PhD thesis, Fakultat fur Maschinenbau der Ruhr-Universitat
Bochum, November 1993.
[20] G.J.J. Ruijgrok. Elements of Airplane Performance. Delft University Press, Delft,
1990.
[21] Brian L. Stevens and Frank L. Lewis. Aircraft Control and Simulation. Wiley-
Interscience Publication. John Wiley & Sons, Inc., New York, 1992.
Version: 1
Date: April 4, 1997 120
GARTEUR/TP-088-27

A The e ect of the center of gravity location


In this section aerodynamic trim equilibrium and static stability are discussed for the
longitudinal aircraft motions. The focus will mainly be on the e ects of the (horizontal)
the center of gravity location. For a more thorough discussion the reader is referred to
[10].
Obviously, for equilibrium the total pitching moment has to be equal to zero, while the
term Cm has to be negative for static stability. In equation 2.41 the total aerodynamic
moment has been derived with respect to the aerodynamic center. This equation can be
written with respect to the center of gravity by adding the lift and drag terms from the
wing-body combination (for de nitions of the terms xcg etc. see gure 2.10). The aircraft
is in equilibrium ight, so that the pitch rate is zero:
2
Cm = Cmac + CLwb xcg cxacwb + CLt SSt VVt2 xcg c xact (A.1)
The term CLwb can be replaced by CLwb ( 0 ) (equation 2.28). The derivative with
respect to the angle of attack can thus be written as follows: (with substitution of eq. 2.44)
 d"  St V 2 xcg xac
Cm = CLwb xcg cxacwb + CLt 1 d t
S V 2 c
t (A.2)
Although it is not done here, it is interesting to add the in uence of the drag to the moment
equilibrium, since in the RCAM model vertical center of gravity shifts are allowed, see
gure 2.10.
It is clear that the center of gravity location has a very important e ect on the static
stability. The more the location is moved aft, the more the aircraft is destabilized. At the
neutral point (elevator xed) Cm has become zero. Substitution of lt = xact xacwb and
equation 2.32:
 d"  St V 2
CL = CLwb + CLt 1 d S Vt2
gives:
xnfix xacwb CLt  d"  Stlt Vt2
c = C 1 d S c V 2 (A.3)
L
Substitution in equation A.2 with xcg = xcg xnfix + xnfix gives the following simple
expression:
xnfix xcg
Cm = CL c (A.4)
The right hand term (xnfix xcg )=c is also known as the \stability margin". Substituting
the values from table 2.2 and 2.1:
xnfix xacwb 3:1 64 24:8 1 = 0:35
c = 6:07 (1 0 : 25) 2606:6 (A.5)
The range of allowed values for (xcg xacwb )=c is for the RCAM model 0:15 0:12 = 0:03
to 0:31 0:12 = 0:19 (data obtained from table 2.1) The stability margin therefore lies
Version: 1
121 Date: April 4, 1997
GARTEUR/TP-088-27

between 0:35 0:19 = 0:16 and 0:35 0:03 = 0:32. The locations of the aerodynamic
centers, the neutral point and the center of gravity can also be found in gure 2.10.
In gure 4.2 the in uence of Cm via the M -block on the short period dynamics can
be seen. Additional feedback of an angle of attack measurement via the tailplane will
e ectively change M and therefore Cm . In this way the neutral point can arti cially be
moved to a more aft location (equation A.4), increasing the stability margin. This margin
can thus be interpreted as a gain margin in a control theoretical sense. The augmentation
of the longitudinal stability allows for a lower natural stability margin, which is the idea
behind "relaxed stability" mentioned in section 2.1.4.
The required elevator de ection for trimming can be obtained from:
 = 1 (C + C ( ))
E CmE m0 m 0 (A.6)
If now is substituted:
CL  CL ( 0 ) = 1=2W
VA2 S (A.7)
the elevator position as a function of the trim speed is obtained:
!
E = C 1 C m W
Cm0 + C 1=2V 2 S
(A.8)
mE L A
Di erentiating this expression with respect to the airspeed gives:
dE = 1 Cm 4W (A.9)
dVA CmE CL VA3 S
Since CmE is always negative and CL is always positive, for a negative Cm is required
that dE =dVA is positive. In gure A.1 the required tailplane position of the RCAM
has been plotted as a function of the trim speed. Static stability is con rmed by the
dT > 0. In equation A.8 also the in uence of the aircraft mass is
observation that dV A
visible. In gure A.2 trimcurves have been plotted for three masses.

 The conclusions derived here for elevator trim, also hold for the tailplane de ection
Version: 1
Date: April 4, 1997 122
GARTEUR/TP-088-27

20

18 mass: 120000 kg
h=0m
16 0.15 c ||-
0.23 c - - - -
14 0.31 c ........
T (deg)

12

10

0
40 50 60 70 80 90 100 110 120 130

Vtrim (m/s)
Fig.A.1 Trim curves for di erent CoG locations

20

18 xcg : 0.23c
h=0m
16 100000 kg ||-
120000 kg - - - -
14 150000 kg ........
T (deg)

12

10

0
40 50 60 70 80 90 100 110 120 130

Vtrim (m/s)
Fig.A.2 Trim curves for di erent masses
Version: 1
123 Date: April 4, 1997
GARTEUR/TP-088-27

B Software for the RCAM nonlinear model


In this appendix more details are given about the implementation of the models in the
Matlab/Simulink environment. In the rst section the contents of the sub-blocks in g-
ure 2.15 are discussed. In the second section the trim routine is described.

B.1 Implementation in Simulink-blocks


In this section the contents of the sub-blocks of gure 2.15 are given. For completeness
the main block diagram is repeated in gure B.1.
The AIRCRAFT subblock in the simulink window has been plotted in gure B.2. In this

LOAD FILE EXECUTE


WITH NEW
TRIM DATA TRIM
time
DOUBLE − DOUBLE − Clock time
CLICK ME CLICK ME

traject
reference
signals
Controls
controls
controller control−
outputs inputs

REFERENCES sim
ouputs for
simulation
CONTROLLER

Engine failure
(1/0)
lon
ACTUATORS
longitudinal
0 measurements
Const. windfield
(1/0)

lat
0 lateral
Turbulence measurements
WIND &
(1/0) TURBULENCE
AIRCRAFT
wind
total wind−
inputs

Fig.B.1 Simulink block diagram of the RCAM simulation environment

gure the control and wind inputs and the three groups of outputs are visible. From the
block NONLINEAR AIRCRAFT MODEL the compiled c-code with the nonlinear state
equations is called, see also appendix F in [3]. The meaning of the alpha-numeric variable
names can be found in [1].
The CONTROLLER subblock will contain the control laws that are to be designed. The
contents of this block will be described in detail in [5].
A more detailed view of the ACTUATOR MODELS block can be found in gure B.3.
There is a separate input (ENGINE FAILURE (1/0)) that can be set to 1 if the left
engine is failed. The contents of the TAILPLANE ACTUATOR block can be found in
gure B.4. The RUDDER ACTUATOR and AILERON ACTUATOR blocks are similar.
Version: 1
Date: April 4, 1997 124
GARTEUR/TP-088-27

In this block the limiters and rate limiters have been implemented. The integrator is
limited with the maximum and minimum tailplane de ections.
A detailed picture of the ENGINE MODEL can be found in gure B.5. In this gure
the trim setting and the control input are added and position and rate limits have been
implemented. In the upper block system a switch has been implemented for engine failure.
As soon as the EF (1/0) input is set to 1, the rst order system is opened while the other
with a time constant of 3:3 s is closed, so that the thrust decays with rst order dynamics
of the lter in equation 2.54.
The contents of the WIND MODELS block can be found in gure B.6. The WINDSHEAR
block can be found in gure B.7. In this gure can be seen that two look-up tabels have
been implemented for the vertical and horizontal wind. The MEMORY block is necessary
to prevent an algebraic loop in Simulink.
In the block DRYDEN the forming lters for Dryden turbulence described in section 2.3.2
have been implemented.

Q (rad/s)

NX (−)

NZ (−)
Floating scope
WV (m/s)
Mux 2
DA ( rad) Z (m) lon

VA (m/s)
DT( rad)
V (m/s)
DR ( rad)
1 Demux BETA (rad)
u THROTTLE1 (rad)
P (rad/s)

THROTTLE2 (rad)
R (rad/s)

WXE (m/s) (in FE) PHI (rad)


Mux rcamex Demux
UV (m/s) Mux 3
WYE (m/s) (in FE)
lat
NONLINEAR VV (m/s)
WZE (m/s) (in FE) AIRCRAFT
MODEL
Y (m)
2 Demux WXB (m/s) (in FB)
wind CHI (rad)
WYB (m/s) (in FB)
PSI (rad)

WZB (m/s) (in FB)


THETA (rad)

ALPHA (rad)
Mux
GAMMA (rad) Mux 1
sim
X (m)

NY (−)

Fig.B.2 Simulink block diagram of the nonlinear RCAM model


Version: 1
125 Date: April 4, 1997
GARTEUR/TP-088-27

U0(1:5) Demux
Trim settings
controls

DA_C

Aileron actuator

DT_C

Tailplane actuator

DR_C Mux 1
1 Demux
Control inputs
uc Rudder actuator aircraft model

THROTTLE1_C

THROTTLE2_C

2
engine failure Engine model
(1/0)

Fig.B.3 Simulink block diagram of the RCAM actuator models

1
DT_trim

+
2 + 1
6.7 1/s +
DT_c − DT
Tailplane Rate limiter Limited
actuator Integrator
bandwidth

Fig.B.4 Simulink block diagram of the RCAM elevator model

Engine
failure
ENGINE 1 constant
5 −1/3.3
ef (1/0)
1 + − +
1 1
THROTTLE1_trim + +
+ − 0.6 s
Switch Integrator THROTTLE1
2 + Lim Engine dyn.
THROTTLE bandwidth Rate lim
THROTTLE1_c THROTTLE

0.5/57.3 0.5/57.3
THROTTLE_min THROTTLE_min

ENGINE 2

3 + −
+ +
THROTTLE2_trim 1 2
+ − 0.6 +
s
4 + Engine dyn. Integrator THROTTLE2
Lim
THROTTLE bandwidth Rate lim
THROTTLE2_c THROTTLE

Fig.B.5 Simulink block diagram of the RCAM engine model


Version: 1
Date: April 4, 1997 126
GARTEUR/TP-088-27

CONSTANT WIND
1 +
From traj.
generator
Const. wind input w.r.t. FE
logic and scaling axis system

longitudinal WINDSHEAR
measurements +

2 Mux 1
simulation TOTAL
outputs WIND−
Windshear INPUT

w.r.t FB
Dryden TURBULENCE axis system
Turbulence
*

Product
Dryden

4
Turbulence
(1/0)

Fig.B.6 Simulink block diagram of the RCAM wind models

1 Demu
lon. −1
meas. WINDSHEAR DATA
positive = 0 Mux 1
downward HORIZONTAL
Constant Memory Windshear
−1 (break alge−
braic loop)
positive =
WINDSHEAR DATA downward
DOWNBURST
2 Demu
sim.
outp.

Fig.B.7 Simulink block diagram of the RCAM windshear model


Version: 1
127 Date: April 4, 1997
GARTEUR/TP-088-27

B.2 The routine TRIMRCAM.M


B.2.1 Overview

TRIMRCAM is the trimming and linearization routine supplied with the RCAM software.
When running this program, the following functions are executed:
 initialization: the user may enter parameters to de ne the trim condition and to set
other parameters for time simulations,
 trimming: for this purpose a nine states (uB , vB , wB , p, q, r, , , ) nonlinear
RCAM model is called, for the actual trimming job the Simulink routine TRIM.M
is used,
 linearization: a state-space model is obtained for small perturbations around the
equilibrium point, using the Simulink routine LINMOD.M,
 initialization for simulations in the Simulink environment: the state vector is mod-
i ed to set the speci ed initial position, track angle, crab angle (wind correction)
etc. These modi cations do not a ect the equilibrium.
 simulation: time responses of the linear and the nonlinear model (starting from the
trim point) are compared in plots. This enables the user to check the quality and
the scope of the linearization.
The main parameters that may be set to de ne a trim point are:
 the aircraft mass, m,
 the center of gravity location, [xcg ycg zzg ]T ,
 the airspeed, VA,
 the ightpath angle, ,
 the inertial wind vector, [WXE ; WY E ; WZE ]T
These parameters may be chosen within their limits, these can be found in table 2.1.

B.2.2 Trimming in a wind eld

Although the suplied trim algorithm can only be used for nding longitudinal equilibrium,
it is possible to correct for a constant wind vector. In gure B.8 the vectors of the
airspeed (Va ), the inertial velocity (V) and the windspeed (VW ) are drawn. The airspeed,
the ightpath angle and the track angle have been speci ed for the trim condition. To
compensate for the vertical wind, the ightpath angle w.r.t. to the airmass ( a ) has to
Version: 1
Date: April 4, 1997 128
GARTEUR/TP-088-27

xE
VW
VA VW
VAcos γa VW (projection)
γa V VW
χ
γ βw
Vcos γ
βw y
E
-z V

VWZE
VW

VAsin γ a Vsin γ

Fig.B.8 Corrections for trimming in constant wind eld


be somewhat larger than . Trimming has to be done for a instead of . To correct for
lateral winds w.r.t. the desired ightpath, the aircraft should have a crab angle w .
The quantities a , V and w need to be calculated before the Simulink trim routine
(TRIM.M) is called. The equations can be derived using gure B.8. The vectors have been
projected on the xE yE plane. First, the wind vector is projected on axes perpendicular
and tangent to the to the groundspeed vector:
VW? = WY E cos  WXE sin  (B.1)
VWk = WY E sin  WXE cos 
The following relations can also easily be derived:
VA cos a sin w = VW? (B.2)
VA cos a cos w = V cos VWk
For the vertical components holds:
VA sin a = V sin + WZE (B.3)
VA , ,  and wind components are known. These equations need to be solved for a , V
and w . This can be done analytically by the assumption that cos a  cos . This is only
valid if and WZE are not too large. The solution can now be calculated:
 WY E cos  WXE sin  
w = arcsin VA cos (B.4)
V = WY E cos sin  W cos  + V cos
XE cos A w
 
a = arcsin V sin V+ WZE (B.5)
A
Version: 1
129 Date: April 4, 1997
GARTEUR/TP-088-27

The RCAM is trimmed numerically with the following constraints:


 =  = =  = 0 rad
 VA and a are as speci ed,
 T and TH 1;2 may be varied to nd equilibrium, the remaining controls and wind
inputs are zero.
The track angle  and the remaining three states are the position co-ordinates and do not
in uence the forces and moments. They can be modi ed after equilibrium has been found
numerically.
The position states (x, y, z ) can be set directly. The body rates (p, q, r; these are equal
to zero) and the attitude angles  and  don't need to be modi ed. To set the correct
values for , (=  w ) and , the states uB , vB , wB and need to be modi ed.
In trim routine rst the crab angle, the inertal speed and the ightpath angle are recal-
culated, using the speci ed wind vector and a and VA obtained from the numerical trim.
The track angle is speci ed, so that the heading angle becomes: =  w . This state
can now also be changed in the state vector. The velocity components along the body
axes can be obtained from:
VB = RBV RV F VF (B.6)
T ; RBV and RFV are given in eq. 2.16 and eq. 2.15.
where RV F = RFV

B.2.3 linearization

In the third part of TRIMRCAM a numerical linearization is executed.


From the nonlinear system description:
x_ = f (x; u) (B.7)
y = g(x; u) (B.8)
(x is the state vector, u is the input vector and y is the outputvector)
a state space model description needs to be obtained at the trim condition where u = u0
and x = x0 :
x_ = Ax + B u (B.9)
y = C x + Du (B.10)
The state space matrices of this LTI system are obtained from Jacobian matrices:
@f @f
A = @ x j0 B = @ u j0 (B.11)
@g @g
C = @ x j0 D = @ u j0
Version: 1
Date: April 4, 1997 130
GARTEUR/TP-088-27

Column aj of A is obtained numerically by perturbing the j th element of the state vector


x0
aj = f [(x0 + xj ); u02] xf [(x0 xj ); u0 ] (B.12)
j
with xj = [0:::0; xj ; 0:::]T . The columns of C are obtained in a similar manner from
g. The columns of B and D matrices are calculated from f and g also by perturbing the
elements of x0 and u0 respectively.
B.2.4 Simulations for veri cation
In the last part of TRIMRCAM.M it is possible to run a simulation in order to compare the
behaviour of the linearized and the nonlinear models and to check the trim conditions. The
results of a simulation run can be found in gure B.9. Trimming was done at VA = 70 m=s,
= 3 deg for the nominal aircraft con guration (see table 2.1). Track angle was set to
 = 60 deg in a constant wind eld: [WXE ; WY E ; WZE ]T = [ 10 5 4]T m=s. The applied
control inputs were T = 0:5 deg, TH 1 = TH 2 = 0:4 deg.
−6 −6 8 1
x 10 x 10
1 1
6
0.5 0.5 0.5
4

0 0 2

Q [deg/s]
0

PHI [deg]
ALPHA [deg]

BETA [deg]
linear
−0.5 linear −0.5 0
nonlinear
nonlinear
−2 −0.5
−1 −1 0 20 40 60 0 20 40 60
0 20 40 60 0 20 40 60

−6
x 10 80 8
1 60

60 6
0.5 40
40 4
0 20
20 2

VA [m/s]

R [deg/s]
PSI [deg]
GAMMA [deg]
−0.5 0 0 0

−1 −20 −20 −2
0 20 40 60 0 20 40 60 0 20 40 60 0 20 40 60
131

−6 −6 2 5
x 10 x 10
1 1
0 4
0.5 0.5 −2 3

−4 2
0 0
−6 1

DA [deg]
DR [deg]
in FE

WXE,WXB [m/s]
WYE,WYB [m/s]

−0.5 −0.5 −8 0
in FB
−10 −1
−1 −1 0 20 40 60 0 20 40 60
0 20 40 60 0 20 40 60

2 8 4

0 6 3

−2 4 2

−4 2 1

DT [deg]
WZE,WZB [m/s]

−6 0 0

THROTTLE1,2 [deg]
−8 −2 −1
0 20 40 60 0 20 40 60 0 20 40 60
GARTEUR/TP-088-27
Date: April 4, 1997
Version: 1

Fig. B.9 Results linear/nonlinear simulations


Version: 1
Date: April 4, 1997 132
GARTEUR/TP-088-27

C linearized RCAM models


In this appendix trimdata and state space models of the linearized RCAM model can be
found, for the nominal conditions. (m = 120000 kg, CoG: [0:23c 0 0]T , VA = 70 m=s and
= 0 deg)

control setting unit


A 0.00 deg
E -8.10 deg
R 0.00 deg
TH 1 4.20 deg
TH 2 4.20 deg

Table C.1 Trim settings of the controls

parameter num. value unit


h 1000.0 m
VA 70.0 m/s
V 70.0 m/s
5.0 deg
 5.0 deg
0.0 deg
Table C.2 Flight data for trim condition
state / State variables: Control inputs: Wind inputs:
output: p r  vB q  uB wB A R T TH 1 TH 2 WY E WY B WXE WY E WXB WZB
p_ -1.1084 0.4811 0.0000 0.0000 -0.0213 0.0000 0. 0.0000 0.0000 -0.6433 0.2223 0. 0.0407 -0.0407 0.0213 0.0213 0.0000 0.0000 0.0000 0.0000
r_ 0.0456 -0.4556 0.0000 0.0000 0.0026 0.0000 0. 0.0000 0.0000 -0.0135 -0.2546 0. 0.7795 -0.7795 -0.0026 -0.0026 0.0000 0.0000 0.0000 0.0000
_ 1.0000 0.0875 0.0000 0. 0. 0.0000 0.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
_ 0. 1.0038 0.0000 0. 0. 0.0000 0.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
v_ B 6.1028 -69.7335 9.7464 0.0000 -0.1486 0. 0.0000 0.0000 0.0000 0. 1.5607 0. 0. 0. 0.1486 0.1486 0.0000 0.0000 0.0000 0.0000
q_ -0.0033 0.0033 0. 0. 0.0000 -0.8598 0. 0.0000 -0.0141 0. 0. -1.8667 0.2909 0.2909 0.0000 0.0000 0.0012 0.0141 0.0000 0.0141
_ 0. 0.0000 0.0000 0. 0. 1.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
u_ B 0. 0.0000 0. 0. 0.0000 -5.9036 -9.7037 -0.0260 0.1065 0. 0. 0.4326 9.8000 9.8000 0.0000 0.0000 0.0166 0.1073 0.0259 -0.1054
w_ B 0.0000 0. -0.4877 0. -0.0002 67.4568 -1.3407 -0.2277 -0.5997 0. 0. -4.9432 0. 0. -0.0002 -0.0002 0.2787 0.5772 0.2274 0.5992
q 0. 0. 0. 0. 0. 1.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
nx 0. 0.0000 0. 0. 0.0000 -0.6018 -0.9892 -0.0026 0.0109 0. 0. 0.0441 0.9990 0.9990 0.0000 0.0000 0.0017 -0.0109 0.0026 -0.0107
nz 0.0000 0. -0.0497 0. 0.0000 6.8763 -0.1367 -0.0232 -0.0611 0. 0. -0.5039 0. 0. 0.0000 0.0000 0.0284 0.0588 0.0232 0.0611
wV 0. 0. -0.3037 0.0000 0. 0. -69.8834 -0.0872 0.9962 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
133

z 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
VA 0. 0. 0. 0. 0.0007 0. 0. 0.9962 0.0879 0. 0. 0. 0. 0. 0.0007 0.0007 -1.0000 0.0007 -0.9962 -0.0865
V 0. 0. 0.0000 0.0000 0.0007 0. 0. 0.9962 0.0879 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
0. 0. 0.0000 0.0000 0.0143 0. 0. 0.0000 0.0000 0. 0. 0. 0. 0. -0.0143 -0.0143 0.0000 0.0000 0.0000 0.0000
p 1.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
r 0. 1.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
 0. 0. 1.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
uV 0. 0. -0.0266 -3.4971 0. 0. -3.4974 0.9962 0.0872 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
vV 0. 0. -6.0926 69.8834 1.0000 0. 0.0000 0.0000 0.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
y 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
 0. 0. -0.0870 1.0000 0.0143 0. 0.0000 0.0000 0.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
0. 0. 0. 1.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
 0. 0. 0. 0. 0. 0. 1.0000 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
0. 0. 0. 0. 0. 0. 0. -0.0012 0.0142 0. 0. 0. 0. 0. 0. 0. 0. -0.0143 0.0012 -0.0142
0. 0. 0.0043 0.0000 0.0000 0. 1.0000 0.0012 -0.0142 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
x 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0. 0.
ny 0.6221 -7.1084 0.9935 0.0000 -0.0152 0. 0.0000 0.0000 0.0000 0. 0.1591 0. 0. 0. 0.0152 0.0152 0.0000 0.0000 0.0000 0.0000

Table C.3 RCAM state space model for nominal trim condition
GARTEUR/TP-088-27
Date: April 4, 1997
Version: 1
Version: 1
Date: April 4, 1997 134
GARTEUR/TP-088-27

Longitudinal A-matrix Longitudinal B-matrix


state State variables: Control inputs:
der. q VA T TH 1 TH 2
q_ -0.8598 -0.9840 -0.0012 0. -1.8667 0.2909 0.2909
_ 0.9672 -0.5916 -0.0039 -0.0070 -0.0709 -0.0122 -0.0122
V_ A 0.0481 -5.7777 -0.0411 -9.7847 -0.0035 9.7627 9.7627
_ 0.0328 0.5916 0.0039 0.0070 0.0709 0.0122 0.0122
Table C.4 Longitudinal linear model, with transformed state vector

Lateral A-matrix Lateral B-matrix


state States variables: Control inputs:
der. r p  A R
r_ -0.4556 0.1791 0.0456 0. 0. -0.0135 -0.2546
_ -1.0000 -0.1486 0.0874 0.1394 0. 0. 0.0223
p_ 0.4811 -1.4900 -1.1084 0. 0. -0.6433 0.2223
_ 0.0878 0. 1.0000 0. 0. 0. 0.
_ 1.0038 0. 0. 0. 0. 0. 0.
Table C.5 Lateral linear model, with transformed state vector
Version: 1
135 Date: April 4, 1997
GARTEUR/TP-088-27

D Coecients for linear models


In this appendix the analytically derived elements of the state space matrix elements can
be found. These are derived in [6].

Longitudinal model

Xu = qmS V20 (CD0 0CL0 ) Zu = qmVS 0 V20 (CL0 + 0 CD0 )


X = qmS (CD 0CL CL0 ) Z = qmVS 0 (CL + 0 CD + CD0 )
X _ = qmS Vc0 (CD _ 0 CL _ ) Z _ = qmVS 0 Vc0 (CD _ + 0 CL _ )
Xq = qmS Vc0 (CDq 0 CLq ) Zq = qmVS 0 Vc0 (CLq + 0 CDq )
X = g
XE = qmS (CDE 0 CLE ) ZT = qmVS 0 (CLE + 0 CDE )
XTH = m1 @@TTH @@TTH = 2m120g ZTH = mViT @T
0 @TH @TH = 2m120 g
@T
X f = X + X = X g Z f = Z
Mu = qSIyc V20 Cm0
M = qSIyc Cm ME = qSIyc CmE
M _ = qSIyc Vc0 Cm _ MTH = zITH
y @TH
@T @T
@TH = 2m120 g
Mq? = qSIyc Vc0 Cmq Mq = Mq? + M _

Lateral model
Y = qmS CY L = qSb= 2 (Iz Cl + Ixz Cn )

Y _ = qmS b=V02 CY _ L _ = qSb= 2 b=2 (Iz Cl + Ixz Cn )
 2 b=V02 _ _
Yp = qmS b=V02 CYp Lp = qSb=
 V0 (Iz Clp + Ixz Cnp )
Yr = qmS b=V02 CYr Lr = qSb=

2 b=2 (Iz Cl + Ixz Cn )
2 (VI0z Cl + Ixz Cn )
r r
YA = qmS CYA LA = qSb=
2 A A
YR = qmS CYR LR = qSb=
 (Iz ClR + Ixz CnR )
N = qSb= 2 (Ix Cn + Ixz Cl ) Nr = qSb= 2 b=2

2 b=2 (Ix Cn + Ixz Cl )
 2 V0 (Ix Cnr + Ixz Clr )
N _ = qSb=
 2 b=V02 _ _
Sb=
NA = q  (Ix CnA + Ixz ClA )
Np = qSb= NR = qSb= 2
 V0 ( I x C n p + I xz C l p)  (Iz ClR + Ixz CnR )
q = 2 V02  = Ix Iz 2
Ixz
Version: 1
Date: April 4, 1997 136
GARTEUR/TP-088-27

E Characteristics of linearized models


In this appendix data can be found of the linearized RCAM models in a number of trim
conditions. The eigenvalues, damping, eigenfrequencies and time constants are given in
tabular form. In the second part bode plots for linearized longitudinal and lateral input-
output relations can be found.
The following trim con gurations are considered:
 the mass (100 000, 120 000 and 150 000 kg) at a center of gravity location xcg =
0:23c, zcg = 0c and VA = 70 m=s
 the center of gravity location (0.15, 0.23 and 0.30c) and zcg = 0:21c at a mass of
m = 120000 kg and VA = 70 m=s
 the airspeed (60, 70, 80 and 90 m/s) at a mass of m = 120000 kg and a center of
gravity location of xcg = 0:23c, zcg = 0c.
Heading is  = 0 deg and altitude is h = 1000 m.

trim speed variations, longitudinal


m = 120 000 kg, xcg = 0:23c
short period phugoid
VA  !0 (rad/s)  !0 (rad/s)
60 0.6070 1.0541 0.0497 0.1638
70 0.6025 1.2167 0.0671 0.1441
80 0.6000 1.3837 0.0899 0.1268
90 0.5986 1.5529 0.1184 0.1120
Table E.1 Characteristics of lon. modes, speed variations

trim speed variations, lateral


m = 120 000 kg, xcg = 0:23c
Dutch roll roll motion spiral mode
VA  !0 (rad/s)  (s)  (s)
60 0.1549 0.4090 1.135 2.171
70 0.3218 0.5071 0.909 3.498
80 0.3685 0.6407 0.768 5.456
90 0.3725 0.7661 0.668 7.402
Table E.2 Characteristics of the lateral modes, speed variations
Version: 1
137 Date: April 4, 1997
GARTEUR/TP-088-27

center of gravity shifts, longitudinal


VA = 70 m=s, m = 120000 kg
short period phugoid
xcg  !0 (rad/s)  !0 (rad/s)
0.15 0.5519 1.3459 0.0640 0.1546
0.23 0.6025 1.2167 0.0671 0.1441
0.30 0.6634 1.0906 0.0811 0.1288
zcg
0.21 0.6358 1.1490 0.0753 0.1339
Table E.3 Characteristics of the lon. modes, center of gravity shifts
center of gravity shifts, lateral
VA = 70 m=s, m = 120000 kg
Dutch roll roll motion spiral mode
xcg  !0 (rad/s)  (s)  (s)
0.15 0.3515 0.5599 0.913 4.466
0.23 0.3218 0.5071 0.909 3.498
0.30 0.2733 0.4649 0.905 2.831
zcg
0.21 0.3967 0.6358 0.902 4.7916
Table E.4 Characteristics of the lateral modes, center of gravity shifts
mass variations, longitudinal
VA = 70 m=s, xcg = 0:23c
short period phugoid
3
m (10 )  !0 (rad/s)  !0 (rad/s)
100 0.6370 1.3714 0.0864 0.1403
120 0.6025 1.2167 0.0671 0.1441
150 0.5616 1.0571 0.0513 0.1470
Table E.5 Characteristics of the lon. modes, mass variations
mass variations, lateral
VA = 70 m=s, xcg = 0:23c
Dutch roll roll motion spiral mode
3
m (10 )  !0 (rad/s)  (s)  (s)
100 0.3947 0.5939 0.738 4.338
120 0.3218 0.5071 0.909 3.498
150 0.1997 0.4294 1.209 2.690
Table E.6 Characteristics of the lateral modes, mass variations
Version: 1
Date: April 4, 1997 138
GARTEUR/TP-088-27

1 and q, T input 4 and VA , T input



10 10

V = 60 m=s - - - VA V = 60 m=s - - -
10
0
V = 70 m=s |{ V = 70 m=s |{
V = 80 m=s .....
2
10
V = 80 m=s .....

−1
10
0
10

q
Gain

Gain
−2
10
−2
10
−3
10

−4
−4 10
10

−5 −6
10 −4 −2 0 2 4
10 −4 −2 0 2 4
10 10 10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
10
4 and VA , TH input
V = 60 m=s - - -
VA V = 70 m=s |{
V = 80 m=s .....
2
10

0

10
Gain

−2
10

−4
10

−6
10 −4 −2 0 2 4
10 10 10 10 10

! (rad/s)
Fig.E.1 Bode magnitude plots for varying airspeed
1 and q, T input 10
4 and VA , T input
VA
10

xcg = 0:15 c - - - xcg = 0:15 c - - -


xcg = 0:23 c |{ xcg = 0:23 c |{

0

xcg = 0:30 c .....


10 2

xcg = 0:30 c ..... 10

−1
10


0
10
q
Gain
Gain

−2
10
−2
10
−3
10

−4
−4 10
10

−5 −6
10 −4 −2 0 2 4
10 −4 −2 0 2 4
10 10 10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
10
4 and VA , TH input
VA xcg = 0:15 c - - -
2 xcg = 0:23 c |{
10
xcg = 0:30 c .....


0
10
Gain

−2
10

−4
10

−6
10 −4 −2 0 2 4
10 10 10 10 10

! (rad/s)
Fig.E.2 Bode magnitude plots for varying center of gravity location
Version: 1
139 Date: April 4, 1997
GARTEUR/TP-088-27

1 and q, T input 10
4 and VA , T input
VA
10

10 104 kg - - - 10 104 kg - - -
0
12 104 kg |{ 12 104 kg |{
15 104 kg .....
10 2

15 104 kg .....
10


−1
10
0
10
q

Gain
Gain

−2
10
−2
10
−3
10

−4
−4 10
10

−5 −6
10 −4 −2 0 2 4
10 −4 −2 0 2 4
10 10 10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
10
4 and VA , TH input
VA 10 104 kg - - -
2 12 104 kg |{
10
15 104 kg .....


0
10
Gain

−2
10

−4
10

−6
10 −4 −2 0 2 4
10 10 10 10 10

! (rad/s)
Fig.E.3 Bode magnitude plots for varying aircraft mass
10
1 p and , A input 10
1 p and , R input
10
0
10
0

−1
 −1
10 10

p
Gain

Gain

p
−2 −2
10 10

−3
V = 60 m=s - - - −3
V = 60 m=s - - -
V = 70 m=s |{ V = 70 m=s |{
10 10

−4
V = 80 m=s ..... −4
V = 80 m=s .....
10 10

−5 −5
10 −4 −2 0 2
10 −4 −2 0 2
10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
10
1 r and , A input 10
1 r and , R input
10
0
10
0
−1 −1
10 10

r r
Gain

Gain

−2 −2
10 10

−3
V = 60 m=s - - - −3
V = 60 m=s - - -
V = 70 m=s |{ V = 70 m=s |{
10 10

−4
V = 80 m=s ..... −4
V = 80 m=s .....
10 10

−5 −5
10 −4 −2 0 2
10 −4 −2 0 2
10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
Fig.E.4 Bode magnitude plots for varying airspeed
Version: 1
Date: April 4, 1997 140
GARTEUR/TP-088-27

10
1 p and , A input 10
1 p and , R input
10
0
 10
0

−1 −1
10 10

p
Gain

Gain
10
−2
p 10
−2

−3
xcg = 0:15 c - - - −3
xcg = 0:15 c - - -
xcg = 0:23 c |{ xcg = 0:23 c |{
10 10

−4
xcg = 0:30 c ..... −4
xcg = 0:30 c .....
10 10

−5 −5
10 −4 −2 0 2
10 −4 −2 0 2
10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
10
1 r and , A input 10
1 r and , R input
10
0 10
0
r
−1 −1
10 10

r
Gain

Gain

−2 −2
10 10

−3
xcg = 0:15 c - - - −3
xcg = 0:15 c - - -
xcg = 0:23 c |{ xcg = 0:23 c |{
10 10

−4
xcg = 0:30 c ..... −4
xcg = 0:30 c .....
10 10

−5 −5
10 −4 −2 0 2
10 −4 −2 0 2
10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
Fig.E.5 Bode magnitude plots for varying center of gravity location
10
1 p and , A input 10
1 p and , R input


0 0
10 10

−1 −1
10 10

p
Gain

Gain

10
−2
p 10
−2

−3
m = 10 1044 kg - - - −3
m = 10 1044 kg - - -
m = 12 104 kg |{ m = 12 104 kg |{
10 10

−4
m = 15 10 kg ..... −4
m = 15 10 kg .....
10 10

−5 −5
10 −4 −2 0 2
10 −4 −2 0 2
10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
10
1 r and , A input 10
1 r and , R input
10
0 10
0
−1 −1

r
10 10

r
Gain

Gain

−2 −2
10 10

−3
m = 10 1044 kg - - - −3
m = 10 1044 kg - - -
m = 12 104 kg |{ m = 12 104 kg |{
10 10

−4
m = 15 10 kg ..... −4
m = 15 10 kg .....
10 10

−5 −5
10 −4 −2 0 2
10 −4 −2 0 2
10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
Fig.E.6 Bode magnitude plots for varying aircraft mass
Version: 1
141 Date: April 4, 1997
GARTEUR/TP-088-27

10
2 and q, and VA , WXE input 10
2 , q, and VA , WZE input

10
0
VA 10
0
VA

−2
−2

10 10
Gain

Gain


−4 −4
10 10

10
−6 10
−6
q
q
−8 −8
10 −4 −2 0 2 4
10 −4 −2 0 2 4
10 10 10 10 10 10 10 10 10 10

! (rad/s) ! (rad/s)
10
0 p, r, and , WY E input
−1
10

10
−2

Gain

p
−3
10

r
−4
10

10
−5

−6
10 −4 −2 0 2
10 10 10 10

! (rad/s)
Fig.E.7 Bode magnitude plots for disturbance inputs
Version: 1
Date: April 4, 1997 142
GARTEUR/TP-088-27

F RCAM simulations
In this appendix gures with time reponses of the linear and the nonlinear RCAM model
to control inputs and wind disturbances can be found. The input for each simulation is
given in the rst sub gure. All other inputs or de ections w.r.t. trim values are zero.
All simulations start from the same trim conditions, see table C.1 and table C.2.

In this appendix responses to the following inputs can be found:


 aileron step (page 143)
 tailplane step (page 144)
 rudder step (page 145)
 throttle step (page 146)
 longitudinal wind step (page 147)
 lateral wind step (page 148)
 vertical wind step (page 149)
 engine failure (page 150 and page 151)
 wind shear (page 152)
Version: 1
143 Date: April 4, 1997
GARTEUR/TP-088-27

−3
x 10
2 15

1.5 10
DA [deg]

NY
1 5

0.5 0

0 −5
0 50 100 0 50 100
time [s] time [s]

0 0

R [deg/s]
P [deg/s]

−0.2 −0.1

−0.4 −0.2

0 50 100 0 50 100
time [s] time [s]
0.5 0.2

0
0
BETA [deg]

−0.5
PHI [deg]

−1 −0.2

−1.5
−0.4
−2

−2.5 −0.6
0 50 100 0 50 100
time [s] time [s]

5 1000

0 999.5
CHI, PSI [deg]

−5
999
H [m]

−10
998.5
−15

−20 998

−25 997.5
0 50 100 0 50 100
time [s] time [s]

Fig.F.1 Aircraft response to aileron step; | = linear, -.-.- = nonlinear


Version: 1
Date: April 4, 1997 144
GARTEUR/TP-088-27

−5 0.1

−6
0.05
DT [deg]

−7

NZ
0
−8
−0.05
−9

−10 −0.1
0 50 100 0 50 100
time [s] time [s]

1.5 9

1 ALPHA [deg] 8
Q [deg/s]

0.5
7
0
6
−0.5

−1 5
0 50 100 0 50 100
time [s] time [s]
12 6

10 4
GAMMA [deg]
THETA [deg]

8 2

6 0

4 −2

2 −4
0 50 100 0 50 100
time [s] time [s]

75 1080

1060
70
VA, V [m/s]

1040
H [m]

65
1020
60
1000

55 980
0 50 100 0 50 100
time [s] time [s]

Fig.F.2 Aircraft response to tailplane step; | = linear, -.-.- = nonlinear


Version: 1
145 Date: April 4, 1997
GARTEUR/TP-088-27

2 0.01

0.005
1.5
DR [deg]

NY
1
−0.005
0.5
−0.01

0 −0.015
0 50 100 0 50 100
time [s] time [s]

0.5 0.2

0
0

R [deg/s]
P [deg/s]

−0.2

−0.5
−0.4

−1 −0.6
0 50 100 0 50 100
time [s] time [s]

0
0.5
BETA [deg]

−1
PHI [deg]

−2

−3 0

−4

−5
0 50 100 0 50 100
time [s] time [s]

20 1001

1000
0
CHI, PSI [deg]

999
H [m]

−20 998

997
−40
996

−60 995
0 50 100 0 50 100
time [s] time [s]

Fig.F.3 Aircraft response to rudder step; | = linear, -.-.- = nonlinear


Version: 1
Date: April 4, 1997 146
GARTEUR/TP-088-27

10 0.05
THROTTLE1,2 [deg]

NZ
0
4

0 −0.05
0 50 100 0 50 100
time [s] time [s]

0.4 6.5

0.2
ALPHA [deg]

6
Q [deg/s]

5.5
−0.2

−0.4 5
0 50 100 0 50 100
time [s] time [s]
11 5

10 4
GAMMA [deg]
THETA [deg]

9 3

8 2

7 1

6 0

5 −1
0 50 100 0 50 100
time [s] time [s]

71 1300

70 1250
VA, V [m/s]

69 1200
H [m]

68 1150

67 1100

66 1050

65 1000
0 50 100 0 50 100
time [s] time [s]

Fig.F.4 Aircraft response to throttle step; | = linear, -.-.- = nonlinear


Version: 1
147 Date: April 4, 1997
GARTEUR/TP-088-27

2 0.03

0.02
1.5
WXE [m/s]

0.01

NZ
1
0
0.5
−0.01

0 −0.02
0 50 100 0 50 100
time [s] time [s]

5.3

0.1 5.2

ALPHA [deg]
Q [deg/s]

5.1
0
5

−0.1 4.9

4.8
0 50 100 0 50 100
time [s] time [s]

6 1

5.5 0.5
GAMMA [deg]
THETA [deg]

5 0

4.5 −0.5

4 −1
0 50 100 0 50 100
time [s] time [s]

72 1000

71.5
V
VA, V [m/s]

71 995
H [m]

70.5

70 990

69.5 VA

69 985
0 50 100 0 50 100
time [s] time [s]

Fig.F.5 Aircraft response to lon. wind step; | = linear, -.-.- = nonlinear


Version: 1
Date: April 4, 1997 148
GARTEUR/TP-088-27

2 0.02

0.015
1.5
0.01
WYE [m/s]

NY
1 0.005

0
0.5
−0.005

0 −0.01
0 50 100 0 50 100
time [s] time [s]

1 0.1

0.05
0.5
R [deg/s]
P [deg/s]

0
−0.05

−0.5 −0.1
0 50 100 0 50 100
time [s] time [s]

1.5 0.5

1
BETA [deg]

0
PHI [deg]

0.5

−0.5
0

−0.5 −1
0 50 100 0 50 100
time [s] time [s]

1 1000.3

1000.2
CHI, PSI [deg]

0.5
1000.1
H [m]

1000
0
999.9

−0.5 999.8
0 50 100 0 50 100
time [s] time [s]

Fig.F.6 Aircraft response to lat. wind step; | = linear, -.-.- = nonlinear


Version: 1
149 Date: April 4, 1997
GARTEUR/TP-088-27

2 0.06

1.5 0.04
WZE [m/s]

NZ
1 0.02

0.5 0

0 −0.02
0 50 100 0 50 100
time [s] time [s]

0.4

0.3 5

ALPHA [deg]
Q [deg/s]

0.2

0.1 4.5

−0.1 4
0 50 100 0 50 100
time [s] time [s]

6 0.5

0
GAMMA [deg]
THETA [deg]

5.5

−0.5

5
−1

4.5 −1.5
0 50 100 0 50 100
time [s] time [s]

1000
70.5
980
VA, V [m/s]

960
H [m]

70
940

920
69.5
900
0 50 100 0 50 100
time [s] time [s]

Fig.F.7 Aircraft response to vert. wind step; | = linear, -.-.- = nonlinear


Version: 1
Date: April 4, 1997 150
GARTEUR/TP-088-27

thrust setting
6
0.4
load fact. nx (|), nz (- - -)
5 0.3

0.2
TH 1;2 (deg)

4
0.1

nx ; nz (-)
3 0

−0.1
2
−0.2

−0.3
1
−0.4

0 −0.5
0 20 40 60 80 0 20 40 60 80

time (s) time (s)

4
pitch rate 40
(|),  (- - -) and (...)
3 30

2
20
; ; (deg)

1
q (deg/s)

10
0

−1 0

−2
−10
−3
−20
−4

−5 −30
0 20 40 60 80 0 20 40 60 80

time (s) time (s)

110
VA (|) and V (- - -) size altitude
1000

100 950

900
90
VA ; V (m/s)

850
80
h (m)

800

70 750

700
60
650
50
600

40 550
0 20 40 60 80 0 20 40 60 80

time (s) time (s)

Fig.F.8 Aircraft response to right engine failure


Version: 1
151 Date: April 4, 1997
GARTEUR/TP-088-27

0.2
ny (|), nz (- - -) ang. rates p (|), q (- - -), r (...)
8
0.15
6
0.1

p; q; r (deg/s)
4
ny ; nz (-)

0.05
2
0

0
−0.05

−0.1 −2

−0.15 −4

−0.2 −6
0 20 40 60 80 0 20 40 60 80

time (s) time (s)

(|), (- - -) and  (...)


120
60
roll angle
100
50

80
; ;  (deg)

40

60
 (deg)

30

40
20

20 10

0 0

−20 −10
0 20 40 60 80 0 20 40 60 80

time (s) time (s)


Fig.F.9 Aircraft response to right engine failure
Version: 1
Date: April 4, 1997 152
GARTEUR/TP-088-27

windspeeds: hor (|), vert (- - -) 0.15


load factor
12

10 0.1
8
WXE ; WZE (m/s)

0.05
6

nz (-)
0
2
−0.05
0

−2 −0.1

−4
−0.15
−6

−8 −0.2
0 50 100 150 0 50 100 150

time (s) time (s)

(|),  (- - -) and (...)


2
pitch rate 20

15
1.5

1 10
; ; (deg)

0.5
5
q (deg/s)

0
0
−0.5

−1 −5

−1.5
−10
−2
−15
−2.5

−3 −20
0 50 100 150 0 50 100 150

time (s) time (s)

95
VA (|) and V (- - -) altitude
700

90 650

85 600
VA ; V (m/s)

80 550
h (m)

75 500

70 450

65 400

60 350

55 300

50 250
0 50 100 150 0 50 100 150

time (s) time (s)


Fig.F.10 Aircraft response to windshear
Version: 1
153 Date: April 4, 1997
GARTEUR/TP-088-27

G About the authors


Ir. G. Looye was involved in FM(AG08) as a graduate student at the Faculty of Aerospace
Engineering at the Delft University of Technology. His graduation work involved aircraft
modeling, and robust, nonlinear ight control. This report served as his preliminary thesis
at DUT. The work was supervised by S. Bennani.
Currently, he is a research associate in the Control Design Engineering Group at DLR in
Oberpfa enhofen, Germany.
Ir. S. Bennani is research scientist at the Faculty of Aerospace Engineering at the Delft
University of Technology. His main interests lie in aircraft modeling and robust, nonlinear
ight control; especially gain scheduled controller synthesis.
Version: 1
Date: April 4, 1997 154
GARTEUR/TP-088-27

Das könnte Ihnen auch gefallen