Sei sulla pagina 1di 123

Modelling and Nonlinear Analysis of Aircraft

Ground Manoeuvres

Etienne Coetzee
Department of Engineering Mathematics
University of Bristol

A dissertation submitted to the University of Bristol in


accordance with the requirements of the degree of
Doctor of Philosophy in the Faculty of Engineering.

February 2011
Abstract

Recent studies in the USA and Europe show that passenger numbers are doubling every 15
years, with a consequent increase in traffic and a demand for new airframes. More efficient sur-
face movements will alleviate congestion due to this growth. An understanding of the ground
dynamics of different sized aircraft is therefore essential. The objective of this thesis is to clas-
sify the ground dynamics of different sized aircraft across the entire operational and design
envelope. The nonlinear nature of the problem generally adds to the complexity of such dy-
namics, where small perturbations in velocity, steering angle or brake application may lead to
significant differences in the performance that can be achieved. The use of industrially tested
models of the A320 and A380 are an important aspect of this work. Good agreement is shown
between simulation results and flight test data, underpinning the validity of the models. These
models are constructed in the MSC.Adams and SimMechanics software environments, where
all relevant information in terms of steering angles, clearance distances, and tyre forces are
provided. The computational challenges related to multibody simulations are highlighted, and
consequently alternative analysis methods are explored. The most widely employed analysis
methods that can be used to study aircraft ground manoeuvres consist of geometric, kinematic,
dynamic, and bifurcation methods. To allow for the nonlinear analysis of industrially-tested
models in a user-friendly environment, AUTO has been integrated with Matlab in the form of a
Dynamical Systems Toolbox. The SimMechanics aircraft models are coupled to AUTO within
this new toolbox, where AUTO has direct access to the states, even though the model equations
are a black-box to the user. This is an important capability that allows one to integrate exist-
ing validated models with the bifurcation software, avoiding significant effort in redeveloping
models for bifurcation analysis.
We show that widely used geometric methods for the calculation of turn widths are not applica-
ble to large aircraft such as the A380, due to the asymmetric thrust and braking inputs that are
required for the U-turn manoeuvre. Bifurcation and continuation methods, on the other hand,
are shown to be effective for the analysis of this type of manoeuvre at a fraction of the cost
of simulations. The presence of a fold bifurcation provides new insight into the dynamics of
U-turn manoeuvres, which is not easily observed from simulation data. Kinematic equations
are used to analyse the stability of an aircraft that is being towed, where we conclude that jack-
knifing can be avoided by maintaining a towing radius that is larger than the wheel base. They
also form the basis of the runway exit studies, from which empirical formulas are derived for
steering angle and clearance predictions. The results of the empirical method compare very
well with kinematic studies, as well as detailed dynamic model simulations, as is demonstrated
with a test case example of an A380 model. The empirical formulas can be used to great effect
during the early design phases of an aircraft programme for the prediction of steering angles
and clearance distances, when very little data is available. The greatest advantage of the pro-
posed method is that any aircraft configuration or runway exit can be analysed. The steady-state
force values that are provided from continuation methods can be used to evaluate the FAA 0.5g
high-speed lateral ground loads regulation. A strong correlation exists between the results from
the analysis and the measurements from an operational loads test campaign. We show that the
A380 can only generate a load that is half the value stipulated by the regulation. This is due to
the nonlinear nature of the tyre properties and the overwhelming influence of the aerodynamics
at higher velocities. This analysis provides additional evidence that a lateral load factor of 0.5
cannot be reached for such a large aircraft.
Acknowledgements

I would like to thank my supervisors, Prof. Bernd Krauskopf and Dr. Mark Lowenberg, for their
continued support and encouragement. Without their guidance and expertise this PhD would
not have been possible. I would also like to thank my industrial supervisor at Airbus, Sanjiv
Sharma, who has supported all the nonlinear dynamics activities at Airbus since 2003. He has
been instrumental in advocating their use within an industrial context. I owe Airbus immense
gratitude for allowing me to pursue this PhD, and I hope the results speak for themselves. I also
would like to thank Bob Thompson at Airbus for his valuable inputs, especially with regards
to the explanation of some of the operational usage scenarios. Thanks also to James Rankin
and Phani Thota who helped to lay the foundations for many of the projects that have followed.
I would like to thank my family in South Africa and in the United Kingdom, who have been
right behind me every step of the way. Lastly, I would like to thank my lovely wife Sarah for
her patience, and our eight week old daughter, Elana, for giving me some added incentive to
complete this work before she was born.
“Science is built up with facts, as a house is with stones.
But a collection of facts is no more a science than a heap of stones is a house.”
Henri Poincaré
Author’s Declaration

I declare that the work in this dissertation was carried out in accordance with the regulations of
the University of Bristol. The work is original except where indicated by special reference in
the text and no part of the dissertation has been submitted for any other degree.
Any views expressed in the dissertation are those of the author and in no way represent those
of the University of Bristol.
The dissertation has not been presented to any other University for examination either in the
United Kingdom or overseas.

Signed:

Dated:
Contents

1 Introduction 1
1.1 Research Motivation and Objectives . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Review of Existing Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Thesis Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Models and Hierarchy of Analysis Methods 11


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 Kinematic Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Dynamic Methods - Modelling and Simulation . . . . . . . . . . . . . . . . . 15
2.3.1 Model Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.2 Normal and Towing Operations . . . . . . . . . . . . . . . . . . . . . 19
2.3.3 Model Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3.4 Computational Challenges of Simulations . . . . . . . . . . . . . . . . 20
2.4 Bifurcation Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.4.1 Bifurcation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.4.2 Dynamical Systems Toolbox — AUTO Integration into Matlab . . . . . 22
2.4.3 Application to Ground Manoeuvres . . . . . . . . . . . . . . . . . . . 23

3 Low-Speed: U-turn Manoeuvres 25


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.2 The U-turn Manoeuvre . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3 The Geometric Approach to the U-turn . . . . . . . . . . . . . . . . . . . . . . 28
3.4 U-turn Results from Simulations . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5 U-turn Performance Using the Bifurcation Approach . . . . . . . . . . . . . . 31
3.6 Turn Centre . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4 Low- and Medium-Speed: Towing 37


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.2 Kinematic Towing Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
4.3 Load Factors due to Towing . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

5 Medium-Speed: Runway Exit Manoeuvres 45


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.2 Steering Inputs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.2.1 Steering Angle Variation During Circular Section of Exit . . . . . . . . 48

i
ii CONTENTS

5.2.2 Steering Angle Variation on Horizontal Section of Exit . . . . . . . . . 50


5.2.3 Steering Angle Predictions from Continuation Methods . . . . . . . . . 51
5.2.4 Steering Angle Comparisons for the Different Methods . . . . . . . . . 51
5.3 Clearance Distances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.4 Load Factors During Runway Exit Turns . . . . . . . . . . . . . . . . . . . . . 55
5.4.1 Runway Exit Design Velocities . . . . . . . . . . . . . . . . . . . . . 57
5.4.2 Load Factors for an A320 . . . . . . . . . . . . . . . . . . . . . . . . 58
5.4.3 Load Factors for an A380 . . . . . . . . . . . . . . . . . . . . . . . . 60
5.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

6 High-Speed: Ground Loads Requirements 65


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
6.2 Side Loads Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.2.1 Limitations of the Regulation . . . . . . . . . . . . . . . . . . . . . . 68
6.2.2 FAA Operational Loads Study . . . . . . . . . . . . . . . . . . . . . . 68
6.3 Aircraft Loads from the Static Balance Equations . . . . . . . . . . . . . . . . 69
6.4 Continuation Analysis of the High-Speed Turn . . . . . . . . . . . . . . . . . 73
6.4.1 Load Factors for an A320 . . . . . . . . . . . . . . . . . . . . . . . . 73
6.4.2 Load Factors for an A380 . . . . . . . . . . . . . . . . . . . . . . . . 77
6.4.3 Individual Tyre Loads for an A380 . . . . . . . . . . . . . . . . . . . . 80
6.5 Relating the Continuation Results to the FAA Study . . . . . . . . . . . . . . . 83
6.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

7 Conclusions and Outlook 87


7.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

A Aircraft Definition 97

B State Definition 100

C Kinematic Dimensions 102

D Dynamical Systems Toolbox, Integrating AUTO into Matlab 105


D.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
D.2 Toolbox Development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
D.3 Benchmarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
D.4 Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Nomenclature

ACARE = Advisory Council for Aeronautics Research in Europe


BLG = Body landing gear
C = Position vector of reference point mid way between the main gears, relative to
an inertial coordinate system (-)
c = Geometric mean position of main landing gear bogie pivot points, relative to
an inertial coordinate system (-)
CG = Centre of gravity
co = Oleo damping coefficient (N.s2 /m2 )
COC = Cockpit over centreline
COR = Centre of runway
cx = X-position of reference point mid way between the main gears, normalised to
the wheel base, relative to an inertial coordinate system (-)
cy = Y-position of reference point mid way between the main gears, normalised to
the wheel base, relative to an inertial coordinate system (-)
cz = Tyre vertical damping coefficient (-)
EOR = Edge of runway
F = Position vector of nose gear relative to reference point mid way between the
main gears (-)
FAA = Federal Aviation Administration
Fco = Damping force in oleo due to the orifice (N)
Fko = Spring force in oleo due to gas (N)
Fo = Total force in oleo (N)
Fx = X-component of the force on the tyre in the tyre coordinate system (N)
Fy = Y-component of the force on the tyre in the tyre coordinate system (N)
Fyf = Final steady-state value of tyre lateral force when conducting a turn (N)
Fz = Z-component of the force on the tyre in the tyre coordinate system (N)
FOD = Foreign object damage
hc = Height of the CG above ground
ICAO = International Civil Aviation Organization
JOS = Judgemental oversteer

iii
iv CONTENTS

kz = Tyre vertical stiffness (N/mm)


lcm = Distance from CG to main gears reference position (m)
lcn = distance from CG to nose gear (m)
Lm = track width normalised to wheel base (-)
lm = Track width, distance between left and right main gears, measured from bot-
tom of strut, or from outer wheel-plane (m)
ln = Wheel base, distance from nose gear to main gears reference position (m)
lo = Oleo stroke (m)
lt = Transition distance from straight line to circular movement (m)
ltw = Turn width (m)
M = Position vector of inner main gear reference position, normalised to wheel
base, relative to an inertial coordinate system (-)
m = Mass (kg)
MLW = Maximum landing weight
MRW = Maximum ramp weight
mt = Tyre mass (kg)
mx = X-position of reference point of inner main gears during a turn, normalised to
wheel base, relative to an inertial coordinate system (-)
my = Y-position of reference point of inner main gears during a turn, normalised to
wheel base, relative to an inertial coordinate system (-)
N = position vector of nose gear, normalised to wheel base, relative to an inertial
coordinate system (-)
N = Number of cumulative occurrences of lateral load factor ny (-)
N0 = Number of cumulative occurrences of lateral load factor when ny = 0 (-)
NLG = Nose landing gear
nx = X-position of nose gear, normalised to wheel base, relative to an inertial coor-
dinate system (-)
ny = Y-position of nose gear, normalised to wheel base, relative to an inertial coor-
dinate system (-)
ny = Lateral load factor: lateral force divided by vertical force, at point of interest
(-)
OEM = Original equipment manufacturer
Rn = Radius of turn measured from the turn-centre to the bottom of the nose gear
strut, normalised to the wheel base (-)
rn = Radius of turn measured from the turn-centre to the bottom of the nose gear
strut (m)
Rm = Radius of turn measured from the turn-centre to outer wheel-plane of inner-
most gear during turn, normalised to wheelbase (-)
rm = Radius of turn measured from the turn-centre to outer wheel-plane of inner-
most gear during turn (m)
CONTENTS v

s = Shape parameter for use in probability studies of lateral ground loads. Specific
to each aircraft (-)
t = Time (s)
u = State vector
V = Velocity magnitude (m/s)
Vm = Velocity magnitude at the main gear reference position (m/s)
Vn = Magnitude of velocity at nose gear, measured in SI units (m/s) or normalised
to wheelbase (1/s)
V̇n = Magnitude of acceleration at nose gear (1/s2 )
Vo = Velocity of the oleo (m/s)
Vx = X-component of the velocity of the tyre or aircraft (m/s)
Vy = Y-component of the velocity of the tyre or aircraft (m/s)
Vz = Z-component of the velocity of the tyre or aircraft (m/s)
WLG = Wing landing gear

α = Tyre slip-angle (deg)


αm = Slip-angle at main tyre position (deg)
αn = Slip-angle at nose tyre position (deg)
δ = Steering angle (deg)
δf = Final steady-state steering angle for a circle with a specific radius (deg)
δ90 = Steering angle when nose gear reaches exit point for 90◦ turn (deg)
δ135 = Steering angle when nose gear reaches exit point for 135◦ turn (deg)
δz = Tyre vertical deflection (m)
ζ = Tyre damping ratio (-)
θn = Angular component of polar coordinate of nose gear position during a turn,
measured from negative inertial x-axis (deg)
θ̇n = Angular velocity of polar coordinate of nose gear position during a turn, mea-
sured from negative inertial x-axis (deg/s)
θm = Angular component of polar coordinate of inner main gear reference position
during a turn, measured from negative inertial x-axis (deg)
λ = Control parameter
µR = Coefficient of rolling resistance (-)
ξ = Angle between fuselage longitudinal axis and edge of runway (deg)
ψ = Heading angle (or yaw angle), where north represents reference orientation
(deg)
ωz = Yaw rate around the CG (deg/s)
Chapter 1
Introduction

1.1 Research Motivation and Objectives

The last century has seen huge strides in the progress of aviation, where the development
of breakthrough technologies such as the metal wing, jet engines and fly-by-wire technolo-
gies have given the companies developing these technologies a clear advantage. These game-
changing technologies become even more pertinent when one looks at the General Market
Forecast for aircraft that is published every two years by Airbus [4]. There it is shown that pas-
senger numbers double every 15 years, with a consequent increased demand for new airframes.
It is predicted that 24,000 new airframes will be needed by 2025 [4].
The Strategic Research Agenda [1] of the Advisory Council for Aeronautics Research in Eu-
rope (ACARE) identifies the effects that such an increase in demand will have on the quality
and affordability of aircraft, the effect on the environment, safety, security and the efficiency of
the air transport system. NASA has published a similar document in the form of the National
Plan for Aeronautics Research and Development and Related Infrastructure [46]. Both reports
highlight similar challenges and identify the automation of aircraft movements, on the ground
and in the air, as a means of meeting the objectives set out in these reports. Automation will
increase the throughput of aircraft at airports. It is envisaged that automation will enable the
aircraft performance envelope to be safely enlarged, thereby giving the aircraft operators the
ability to customise their operations based on their market needs. Aircraft manufacturers are
also constantly striving to improve the efficiency of all aspects surrounding the operation of
their aircraft. Obvious fuel savings can be made by decreasing the drag of the aircraft during
the cruise phase. However, less obvious savings can be made by improving the way aircraft
are operated on the ground. Recent studies indicate that efficiencies can be made if surface
movements can be achieved through means other than that of the engines [5]. A large amount
of fuel is consumed when an aircraft’s main engines are used to taxi around airports, where a
forecasted total cost of fuel consumption during taxiing of around $7bn annually, is predicted
by 2012 [5]. It is also predicted that this type of fuel consumption leads to CO2 emissions of
approximately 18m tonnes per year, while Foreign Object Damage (FOD) contributes to a cost
of around $350m per year [5].
New operational procedures are also directed towards the reduction of noise, and taxiing conse-
quently impacts the environment in terms of air and noise pollution. Several schemes for saving
fuel on the ground have been proposed. The Taxibot [5] project is one such scheme, where the

1
2 Chapter 1. Introduction

engines are shut down and the pilot controls the movement of the aircraft on the ground via
a remotely controlled tug. New agile methods are needed to analyse novel architectures re-
sulting from these new procedures, thus enabling the search of new optimal operating regimes
for the proposed architectures. The intended interactions amongst the Aircraft Architectures
and the Operational Environment Architectures are being envisioned and developed in projects
such as NextGen [23] in the USA and CleanSky [20] in Europe. New competition from Japan,
China, Russia and Brazil in the Single Aisle class of aircraft will only increase the need for new
and innovative technologies, where the competitive advantage can be obtained by increasing
and combining the functionality of the sub-systems, whilst lowering the development cost and
improving the operability of these systems. This is one reason why original equipment man-
ufacturers (OEMs) such as Boeing and Airbus are putting greater focus on Architecting and
Integration roles, whilst seeking competitive advantages by utilizing the diversity that emerges
through creating strategic partnerships in design and manufacturing [51].
Not only does automation address the goals of CleanSky, but it also increases the competitive
advantage of the company that masters this skill. Examples of ground automation can already
be seen on aircraft such as the Airbus A380, which has incorporated two functions that cre-
ate greater operational efficiencies for aircraft ground movements. The first is the Brake to
Vacate (BTV) function that reduces the time that the aircraft spends within the active area of
the runway; and the second is the Heading Control Function (HCF), which ensures that the last
demanded heading by the pilot is maintained [71]. In both cases, the crew workload is reduced,
allowing them to concentrate on other more important tasks during the respective flight phase.
Full automation on the ground is yet to be realised; the current sub-systems have only focussed
on straight-line movement. Full automation will only be possible through a clear understand-
ing of all the nonlinear effects that influence the lateral and yaw movement of the aircraft,
especially during the turning phases of the aircraft on the ground.
Aeroplane characteristics manuals from the OEMs usually contain a baseline set of operating
procedures that is used to demonstrate compatability with existing airport infrastructure. The
three most common of these manoeuvres are the U-turn manoeuvre, an exit manoeuvre from
the runway onto a taxiway, and the transition from one taxiway to another. There are several
ways to conduct and interpret these manoeuvres due to the variability that is introduced by the
pilot and the operating procedures of the airlines. A great deal of variance may exist in the way
that any of the ground manoeuvres is conducted, whether around the apron or during taxiing.
This is due to the nonlinear nature of the landing gears [15], as well as the variance of the
inputs from the pilot. A good example that highlights this variance can be found in the A380
Airplane Characteristics Manual [2], which states:

“In the ground operating mode, varying airline practices may demand that more
conservative turning procedures be adopted to avoid excessive tyre wear and re-
duce possible maintenance problems. Airline operating techniques will vary in the
level of performance, over a wide range of operating circumstances throughout the
world. Variations from standard aircraft operating patterns may be necessary to
satisfy physical constraints within the maneuvering area, such as adverse grades,
limited area or high risk of jet blast damage. For these reasons, ground maneu-
vering requirements should be coordinated with the using airlines prior to layout
planning”.
1.1. Research Motivation and Objectives 3

Landing gear engineers observe nonlinear phenomena such as hysteresis, backlash and stiction
on a daily basis, without necessarily appreciating the full meaning behind these observations.
A wheel that locks up during braking is a good example. Many conflicting requirements need
to be considered during the design of a landing gear, where the weight and pavement load-
ing need to be minimised, and the shock absorption maximised. The lateral stability on the
ground is determined by the position of the gears, along with the tyre and oleo (shock damper)
characteristics. Experience has shown that the use of different tyres can mean the difference
between a stable and an unstable aircraft. Landing gears contain highly nonlinear components,
including tyres, brakes and oleos, and therefore traditional analysis is usually done at some
very specific design conditions. There is a perceived need to characterise the behaviour of the
system over a wide variety of parameters, and this is the industrial domain where methods from
nonlinear dynamics can and should be brought to bear.
The main aim of this thesis is to classify the ground dynamics of different sized aircraft
across the entire operational and design envelope. The baseline aircraft used in this study are
the A320 and A380, both from the Airbus family of aircraft. Figure 1.1 depicts these, while Ap-
pendix A contains detailed drawings and information pertaining to these aircraft [3]. The A320
has three landing gears, and is therefore statically determinate, while the five landing gears
of the A380 leads to a statically indeterminate gear configuration. In an attempt to cover all
ground manoeuvres, we propose a scheme where ground manoeuvres are categorised accord-
ing to the operational velocities of the aircraft. To this end, we define a low-speed (0-6 knots),
medium-speed (6-14 knots) and high-speed (> 14 knots) category, where Figure 1.2 indicates
the types of ground manoeuvres that can be conducted within each speed range. We discuss
all the methods that can be used to study ground manoeuvres; from simple geometric methods,
to kinematic and dynamic methods, to relatively new methods from dynamical systems theory.
We discuss how these methods can be employed to gain a complete characterisation of the
aircraft’s dynamic performance on the ground.
Validated models should always be used to serve as a reference point for the current state of the
art. Baseline models were developed within the Landing Gear group of Airbus, where they are
mainly used for ground manoeuvrability studies. The models are of considerable complexity,
especially for the A380, as it is a large aircraft with 5 landing gears and 22 wheels. The
models contain all aspects that are of importance to the dynamics of the aircraft on the ground,
including the tyre properties, gear flexibility, brake logic, engine properties, tow configurations,
and different mass configurations. Such a model provides a versatile means of investigating
different configurations and propulsion schemes for towing and exit manoeuvres [15]. The
different analysis methods are compared for each ground phase, and a judgement is made with
regards to the suitability of each method for each specific phase.
There is a tendency nowadays to immediately resort to detailed simulation models whenever
ground performance studies are conducted, even for tasks that could be performed by more
efficient methods. This is mainly due to a perception that accurate performance predictions
cannot be made unless detailed tyre and landing gear models are available. The art however
lies in the use of the appropriate method, best suited to the type of analysis under consideration.
In this thesis we propose a hierarchy of methods that can be used at various stages of an aircraft
programme, where each method is rated for its suitability at the stage under consideration. The
ultimate aim is to provide tools that can be used by design engineers at all stages of an aircraft
programme, with an increasing level of complexity as the programme progresses. Steering
4 Chapter 1. Introduction

Figure 1.1. Two aircraft from the Airbus family; the A320 on the left and the A380 on the right.

o o
90 Exits 135 Exits
Runway Runway
U-turn High-speed
Exits

o o
90 Exits 135 Exits
Taxiway Taxiway
Forward Take-off
Towing Run

Pushback Fuel Savings Maintenance


Towing Towing Towing
Touch-down
Run

Low-speed Medium-speed High-speed

Figure 1.2. Types of ground manoeuvres.

angles and clearance distances are of importance during the preliminary design phase, while
tyre forces and gear loads become more relevant during the detailed design phase. Figure 1.3
contains typical design phases during an aircraft programme, with the associated ground ma-
noeuvring information that is required at each phase. A summary of the applicability of the
methods discussed in this thesis, to the types of information a design engineer may require,
appears in Table 1.1. The reference to stability refers to an appropriate measure of stability and
its change when a parameter is varied (such as root locus plots), and not merely to observa-
tions from simulations. It is usually not a trivial process to obtain root locus information from
multibody simulations, hence the reason for omitting this capability in Table 1.1.

1.2 Review of Existing Work

The literature relating to ground dynamics is not extensive, and is reviewed here briefly. Chai
and Mason [11] use a geometric method to calculate the steering angle that is needed for run-
way exits; it is equivalent to the steady-state values that are obtained by a kinematic method
1.2. Review of Existing Work 5

Specification Conceptual Preliminary Detail Production Integration & First Flight Operations Entry Into
Design Design Design Start Validation Service
1 2 3 4 5 6 7 8 9
Concept
Definition
Production
- Roll-angle - Roll-angle - Roll-angle
- Tail-clearance - Tail-clearance - Tail-clearance
at take-off at take-off at take-off
- Exit clearance - Exit clearance
- Max steer angles - Max steer angles
- Kinematic stability - Kinematic stability
- Steering torques
- Tyre forces
- Gear loads
- Dynamic stability

Figure 1.3. Design phases with typical ground manoeuvring requirements for each phase.

Table 1.1. Applicability of methods for ground manoeuvre studies

Empirical Kinematic Bifurcation Dynamic


methods simulations methods simulations
Clearance distances X X × X
Steering angles X X X X
Forces × × X X
Kinematic stability × × X ×
Dynamic stability × × X ×
Ease of use X X × ×
Computational efficiency X X X ×

that was developed by Fossum and Lewis [25]. Kinematic methods are used to analyse the sta-
bility of truck-trailer combinations [25], and they form the starting point for many other studies
related to the control of towed vehicles [6, 61, 72]. Traditional dynamic approaches towards
the analysis of vehicle dynamics tend to involve the derivation of nonlinear equations, which
are then linearised for ease of analysis. The stability analysis of nose landing gear wheels is a
good example of this practice [52]. More theoretical approaches pertaining to overall vehicle
dynamics are derived by Gillespie [26] and Wong [73].
Computers have allowed for the wide-spread use of multibody systems methods in the anal-
ysis of complex road vehicles [9, 60], where the full set of nonlinear equations are solved.
Pritchard [52] cites numerous examples from the literature in which commercial and custom
made multibody system software is used for the analysis of vehicle and landing gear dynam-
ics. Multibody systems software is also widely used during the design of new aircraft and for
the analysis of existing aircraft. The Landing Gear group at Airbus uses several commercial
and custom made multibody systems software packages, of which MSC.Adams [45] is the pri-
mary package in use today. Equivalent models are also implemented in SimMechanics [42] for
use on test rigs; they are validated against existing MSC.Adams models and flight test data.
Both MSC.Adams and SimMechanics are software packages that use the multibody-systems
approach to determine the dynamic behaviour of the system. The advantage of working with
multibody systems packages, such as MSC.Adams and SimMechanics, is that the equations are
6 Chapter 1. Introduction

automatically derived; hence, an environment is created where the engineers can focus on the
engineering aspects of the task in hand, and not necessarily on the derivation of the equations
of motion. Such packages provide a complete framework within which models can be built
and simulated. Multibody models are not only used for simulations, but also for bifurcation
analysis, as is demonstrated in this thesis.
Thota et al. [64] showed how nonlinear geometric effects have a significant influence on the
onset of nose gear vibrations. Tyres create the most significant nonlinear effects in traditional
road vehicles [50], and similar effects were found in aircraft tyres at low velocities [13, 54,
55]. Klyde et al. [33] conducted specific ground tests to evaluate aircraft ground handling
characteristics. They showed that the aerodynamic effects are far more significant in aircraft
at high velocities, when compared to cars [35]. The effect of tyre pressure on ground handling
was also investigated in [34], as well as an assessment of the effectiveness of an augmented
steering system [36]. Nonlinear models are also used to a great extent in the area of flight
mechanics, where Thompson and McMillan [63] provide an overview of their use.
Bifurcation analysis has been used successfully to study the longitudinal motion of low-order
road vehicle models with periodic forcing [75] and driver feedback control [39, 40]. Steady-
state behaviour, periodic motions and chaotic dynamics were found in these models. The lateral
dynamics of road vehicles were studied by Nguyen et al. [48, 49, 70], and they showed that the
entry into a spin can be associated with a bifurcation point, indicating a loss of stability.
The first application of bifurcation and continuation methods in aerospace, was in the area of
flight mechanics [43], and it is now used as an effective tool for the study of nonlinear phe-
nomena in aerospace vehicles. Some examples in the field of flight mechanics can be found
in [12, 41], where the aerodynamics creates the dominant nonlinear effects. Bifurcation meth-
ods have also been identified by NASA as a key technology for the analysis of aircraft flight
dynamics in off-nominal conditions [38], in other words, during upset conditions. Bifurca-
tion and continuation techniques have been used to study nose gear vibrations (also known
as wheel shimmy) during straight-line aircraft motion, using low-order mathematical mod-
els [64, 65, 66, 67, 68].
The application of bifurcation and continuation methods, to study an aircraft turning on the
ground, is still quite a new subject. The original research in this area was done by the author
for his Master’s thesis [13], which was the first practical demonstration of the usefulness of
bifurcation methods for the study of aircraft ground manoeuvres. Further studies by James
Rankin identified safe ground operating regions for the A320, with the accompanying modes
that lead to a loss of control [54, 55]. These studies used industrially developed models and a
simplified mathematical model [55, 56]. This thesis expands on the statically determinate gear
arrangement of the A320 by analysing different ground manoeuvres and additional mass cases.
The bifurcation analysis of an aircraft with a statically indeterminate gear arrangement — the
A380 — is new to the field, as is the identification and comparison of the different methods
that can be used to analyse ground manoeuvres. The advantages of these methods are that they
produce a complete picture of the dynamics in all operating regions, and at a fraction of the
cost of simulations.
It is known that the current regulation for high-speed turns is very conservative for large air-
craft [57], and consequently the Federal Aviation Administration (FAA) conducted a measure-
ment campaign of the operational loads that an aircraft may experience during normal opera-
tions. The aim of this campaign was to identify the factors that affect the operational loads, and
1.3. Thesis Overview 7

to assess the existing certification criteria [57, 58, 59]. A further study by the FAA [69] com-
pares the operational loads for a range of different sized aircraft, and showed that the lateral
load factor is reduced when the size of the aircraft is increased. Empirical formulae were de-
rived to calculate the statistical probability of certain load factors. The statistics for the taxi-in
phase showed larger lateral load factors than the taxi-out phase, and it was recognised that cor-
rections were needed to account for the change in the aircraft weight during the taxi-in phase.
Another FAA study [32] was conducted to account for this weight change. Specific ground
tests were also conducted to address concerns that were highlighted in the previous studies,
where Finn et al. [24] performed ground tests to find the maximum lateral loads at individual
landing gears. The use of the A320 and A380 models allows us to assess the validity of these
formulae. It also allows for the identification of the main factors that reduce the lateral load
factor with an increase in size. Rankin et al. [56] conducted simulations to ascertain the max-
imum load factors that could be obtained at runway exits. The results suggest that the limit
imposed in the Federal Airworthiness Regulation (FAR) is conservative for the main landing
gears and possibly not stringent enough for the nose landing gear.

1.3 Thesis Overview

This thesis covers the most widely used operational procedures that are conducted on an aircraft
during ground movements, either under its own power or by means of a tug. The chapters
follow from the types of manoeuvres that are identified in Figure 1.2, where the chapters are
ordered to start from low-speed U-turn manoeuvres, building up to high-speed turns. Figure 1.4
contains an overview of the different chapters. The size of an aircraft has a significant influence
on the dynamics of an aircraft on the ground. The dynamics of an aircraft with three landing
gears, such as the A320, has different dynamics than an aircraft such as the A380, which has
five landing gears. The analysis in each chapter will highlight these differences.
In Chapter 2 we discuss the most widely employed analysis methods that can be used to
study aircraft ground manoeuvres. The first is a kinematic method that was originally devel-
oped for jackknifing studies [25]; it is suitable for clearance and steering angle investigations.
The second method makes use of simulations, and shows how models are constructed in the
MSC.Adams and SimMechanics software environments; all relevant information in terms of
steering angles, clearance distances, and tyre forces are provided. We also discuss how the
models are validated. Good agreement is shown between the simulation results and flight test1
data, underpinning the validity of the models, making them suitable for ground manoeuvre
studies. The computational challenges related to multibody simulations are also highlighted.
The models presented in this chapter have been published in [14, 15]. The final method is
based on bifurcation and continuation methods; it can be used to obtain turn radii, or steady-
state forces on any gear or tyre. To allow for the nonlinear analysis of industrially-tested models
in a user-friendly environment, AUTO [19] has been integrated with Matlab in the form of a
Dynamical Systems Toolbox. The SimMechanics models are coupled to AUTO within this new
toolbox, where AUTO has direct access to the states of the SimMechanics model, even though
the model equations are a black-box to the user. This is an important capability that allows
one to integrate existing validated models with the bifurcation software, avoiding significant
1
Ground tests are also classified as flight tests.
8 Chapter 1. Introduction

Chapter 2:
Models and Hierarchy
of Analysis Methods

Chapter 6: Chapter 3:
High-Speed 0.5W W

Low-Speed
Ground Loads 0.5V
M2
V
0.5VN 0.5V
M1
V U-turn Manoeuvres
M2 M1

Requirements NOSE WHEEL TYPE

Chapter 5: Chapter 4:
Medium-Speed Low- and Medium-Speed
Runway Exit Towing
Manoeuvres

Figure 1.4. Chapter overview.

effort in redeveloping models for a specific application, and allows one to construct bifurcation
diagrams as functions of one or more operational and/or design parameters [13, 54, 64]. A
description of the toolbox and its applications appeared in [16].
Chapter 3 shows how bifurcation methods can be used to obtain turn widths for the A380,
during a U-turn manoeuvre. The main assumption is that the forces are close to their steady-
state values due to the low-speed at which this manoeuvre is conducted. We show that widely
used geometric methods for the calculation of turn radii are not applicable to large aircraft
such as the A380, due to the asymmetric thrust and braking inputs that are required for this
manoeuvre. We show that the feasible region of a ground manoeuvre is defined by an algebraic
constraint, where the desired turn width forms a boundary in parameter space. Bifurcation
methods are used to follow this constraint as parameters, such as the thrust and steering angle,
are varied. From the practical point of view, the size and location of the feasible region give
a clear picture of whether or not a ground manoeuvre can be conducted. We also show how
new dynamics can be discovered by using bifurcation methods. This information was present
in the original simulation results, but was not easily identified, hence the tools from dynamical
systems theory provide us with new insights into the dynamics of U-turn manoeuvres. This
information can be used to inform maintenance and operating procedures. The results presented
in Chapter 3 have been published in [14, 15].
There is a renewed interest in the use of tugs for normal taxi operations [5], hence we study
the effects that towing may have on the aircraft in Chapter 4. Several reports of jackknifing
incidents involving towbarless tugs have been reported recently [30], and consequently we ad-
dress the issue of aircraft towing stability by building on previous work related to the towing
of truck-trailer combinations [25]. The original focus was on straight-line and circular towing
stability, which is also directly applicable to a tug that is towing an aircraft. Airport apron
and taxiway markings consist of a combination of straight and circumferential lines, hence the
1.3. Thesis Overview 9

analysis of straight and circular manoeuvres is adequate. We do not derive the proofs for sta-
bility, as this was done by Fossum and Lewis [25]. We do however show all the different ways
in which jackknifing may occur. An aircraft that is towed in a forward direction is inherently
stable, while an aircraft that is being pushed back is inherently unstable. We also show that
an aircraft will converge to a stable circular movement when the nose gear trajectory is larger
than its wheel base when it being towed, while jackknifing will occur when the towing radius
is smaller than the wheel base. We use the results from a continuation analysis to determine the
effects that towing has when compared to when the aircraft moves under its own power. In this
way, we show that the forces on the aft-axles of the body gears are significantly higher when
the aircraft is being towed; this is offset by lower forces on other parts of the gears.
In Chapter 5 we show that an exit manoeuvre is essentially a transition from a straight line to
a circular trajectory, where the shape of the steering angle curve forms an exponential function
that eventually settles into a steady-state value. This is contrary to previous approaches where
a ramp input is assumed [56]. The kinematic method is used for the initial analysis of runway
exit manoeuvres; it is however still deemed to be a difficult method for everyday use. Therefore
empirical formulae for the steering angle variation are derived. The strength of these formulae
lie in their validity for any aircraft configuration. A comparison is made between the empirical,
kinematic, and dynamic methods, where we conclude that the empirical and kinematic methods
are sufficient to predict the steering angle variation for a towed case, while a minor adjustment
is needed for the self-propelled case. Empirical equations are also derived for the minimum
clearance distances that can be expected for 90◦ and 135◦ exits. A comparison is again made
between the empirical, kinematic, and dynamic methods, where it is shown that the difference
in the clearance predictions from the different methods are negligible. A diagram that indicates
the feasibility of exit manoeuvres at Group V and VI airports is then constructed for any aircraft
configuration; it gives valuable insight into the effects of wheel base and track width on the
clearance distances for specific configurations. This diagram can be used as an effective tool
for design purposes. The last part of the chapter studies the dynamic forces that are generated
during runway exit manoeuvres. Symmetric thrust with no braking is assumed. The results
show that the dynamic force values at the main gears are approximately 10% larger than the
steady-state values for the A320, and that the steady-state values are in fact the maximum
values that can be obtained for the A380. Continuation methods can therefore be used to
obtain the loads at the gears and the tyres for exit manoeuvres. The analysis methods and
results presented in Chapter 5 have been published in [14, 15].
In Chapter 6 we study the loads that can be generated during a high-speed turn. The original
lateral ground loads requirement for an aircraft during a high-speed turn was written in the
middle of the last century, and consequently it is felt that this requirement is conservative
when applied to large modern passenger aircraft [69]; the results from a operational loads
measurement campaign support this statement [69]. We assess the loads that can be generated
by an A320 and an A380, and compare the results to the original requirement. We show that
static balance equations and continuation methods can be used to assess the loads. Comparisons
are made between the two aircraft types, and they show significantly different dynamics in
terms of stability and loads. Symmetric thrust with no braking is once again assumed. Hopf
bifurcations indicate a loss of stability for the A320 [13, 54, 55], while no bifurcations were
detected for the A380. The results for the A380 do however show that the nose gear tyres
operate close to a lateral load saturation point above certain velocities. An increase of the
steering angle has no effect on the turn radius above the value where the saturation occurs, and
10 Chapter 1. Introduction

consequently the radius stays close to a constant value above this saturation value. We show
a strong correlation between the results from continuation analysis and the results from the
measurement campaign, and explain how an A320 can possibly obtain the lateral loads values
that were observed in the test campaign. The A380 can only generate a load that is half the value
stipulated by the requirement. This is due to the nonlinear nature of the tyre properties and
the overwhelming influence of the aerodynamics at high velocities. This provides additional
evidence that a lateral load factor of 0.5 cannot be reached for such a large aircraft.
In Chapter 7 we present a summary of our findings and outline directions for future work.
Chapter 2

Models and Hierarchy of Analysis Methods

2.1 Introduction

Ground manoeuvre predictions are usually made with advanced modelling and simulation tech-
nologies, and they form an invaluable tool within the design process. They are used for detailed
performance predictions to test aircraft manoeuvrability under normal and abnormal condi-
tions, as well as the definition of towing procedures for operators. Detailed models that contain
all the physical characteristics of the landing gears and tyres are used to analyse the dynamics
of an aircraft on the ground. The results from the simulations are used to obtain clearance
distances, steering angles, forces on the tyres and gears, and many other parameters of interest.
Validated models and methods therefore form the basis of the predictions. Figure 2.1 contains
a schematic of typical components within an aircraft ground manoeuvrability model. In this
study we use validated SimMechanics models of the A320 and the A380 [13, 15, 54, 55]. The
A320 consists of three landing gears; one nose gear, and two main gears that are attached to the
wings. Each gear has an axle with two wheels. The A380 has five landing gears; one nose gear,
two main gears attached to the wings, and two main gears attached to the fuselage. Each wing
landing gear (WLG) consists of a bogie with four wheels on each gear, while each body landing
gear (BLG) consists of a bogie with six wheels. The aft axle of each BLG is steered. Six tyres
are therefore present in the case of the A320 model, and 22 tyres in the A380 model. Figure 2.2
contains the numbering conventions of the wheels for both the A320 and A380 models.
Despite all the advantages that simulations bring, there are drawbacks related to the run-times
of such simulations, the skills that are needed by specialist engineers to build satisfactory mod-
els, as well as the availability of data during the early design phases. The costs related to
dynamic simulations make it necessary to evaluate all the different analysis techniques that can
be used for the specific problem at hand. The following approaches can be followed if the aim
is to reduce the analysis time:

• Reduction of computational run-times. This can either be achieved by using more


powerful computational resources, or by focussing on more efficient algorithms. Re-
cent work on the reformulation of the equations of motion by Udwadia and Kalaba [18]
promises to ease the construction of the underlying equations, as well as provide im-
provements in the run-times.

11
12 Chapter 2. Models and Hierarchy of Analysis Methods

Weather & Runway Aircraft Structural Runway


Temperature Surface Dynamics Geometry
Towing Turning
Performance
Landing Gear Gear Mechanical Tyre Aircraft Aircraft
Geometry Characteristics Characteristics Dynamics Loads

Cross Wind Landing Gear


Tiller Steering Effects Loads
Rudder
Pilot Aircraft
Throttles Propulsion Geometry Inputs
Brake
Controls
Pedals Weight &
Braking Balance Systems
Wheel Outputs
Speed Passive Elements

Figure 2.1. Ground manoeuvrability model components.

N1 N2 N1 N2
WLGL W2 W3 WLGR
W1 W4

W6 W7

W1 W4 W5 W8
W2 W3
W9 W12

W10 W11

W13 W16
W14 W15
BLGL W17 W20 BLGR
W18 W19

Figure 2.2. Wheel numbering definition for the A320 and the A380.
2.2. Kinematic Methods 13

• Utilising existing models in a different way. This could refer to the use of simulation
models with bifurcation methods. Existing simulations of the A320 have for instance
been replaced with bifurcation methods for the stability analysis of an aircraft manoeu-
vring on the ground [13, 54, 55].

• Using different analysis methods. Simulations are not always needed, but have be-
come popular due to the large amounts of information that they provide. Geometric and
kinematic methods, instead of detailed simulations, may for instance be used to analyse
clearance distances at runway exits.

In this chapter we focus on three different methods that can be used to analyse ground ma-
noeuvres. Section 2.2 describes how kinematic analysis methods can be used to characterise
ground manoeuvres, as an alternative to dynamic simulations, while Section 2.3 describes how
dynamic models are built and used; the computational difficulties surrounding such dynamic
simulations are also discussed. Finally, Section 2.4 shows how bifurcation methods can be used
to categorise the dynamics. It also discusses the use of simulation models with the Dynamical
Systems Toolbox; a new tool for dynamical systems analysis.

2.2 Kinematic Methods

A kinematic model describes only how points on a body move in relation to one another due to
certain geometric constraints. It can therefore be used to study the motion of a body, while dis-
regarding the forces that cause the motion. Dynamic analysis, on the other hand, incorporates
these forces. A previous towing study by Fossum et al. [25] derived a kinematic model for the
analysis of a truck with a trailer, and it classified the stability of the truck-trailer combination
for straight-line and circular manoeuvres. These same equations can be used for an aircraft that
is being towed, or for an aircraft that conducts a runway exit manoeuvre. In both cases the nose
landing gear is constrained to follow a specific trajectory, where the constraint is enforced by
defining the evolution of the velocity vector at the nose gear position. Figure 2.3(a) shows how
the wheel base and track width are defined. The wheel base ln is measured from the bottom of
the nose gear strut to point c, where point c represents the mean position of all the bogie pivot
points of the main gears. Figure 2.3(b) contains all the relevant position and velocity vectors.
For completeness, we show how the kinematic equations are derived [25]. The wheel base is
normalised, hence |N − C| = 1, and consequently

(N − C) · (N − C) = 1, (2.1)

where the dot represents the dot product of the vectors. It is assumed that the velocity vector
at point c is along the longitudinal axis of the aircraft. This assumption was shown in [25] to
be sufficient, and it correlates with our own experience of the manoeuvres under considera-
tion. The longitudinal velocity is usually an order of magnitude larger than the lateral velocity,
therefore

Ċ = λ(N − C). (2.2)


14 Chapter 2. Models and Hierarchy of Analysis Methods

(a) (b)

wheel base=ln

c F
track=lm N=Vn
m
n δ
C
M N

Figure 2.3. Aircraft representation with, (a) definition of the wheel base and track width, (b) nose
landing gear constraint in the form of a defined velocity vector Vn .

The velocity vector of point c is represented by Ċ, and the normalisation of the wheel base
implies that λ is the magnitude of the normalised velocity. Differentiating Equation (2.1)
yields

2(Ṅ − Ċ) · (N − C) = 0, (2.3)

which leads to

Ṅ · (N − C) = Ċ · (N − C). (2.4)

This equation states that the velocity in the direction of the longitudinal axis is equal at points
n and c. This is true for any rigid body. Taking the dot product of Equation (2.2) with (N − C)
yields,

Ċ · (N − C) = λ(N − C) · (N − C) = λ. (2.5)

From Equations (2.4) and (2.5), λ = Ṅ · (N − C), and Equation (2.2) becomes

Ċ = [Ṅ · (N − C)](N − C) (1/s). (2.6)

Note the normalised units that are used. The velocity can be obtained by multiplying by the
wheel base. When written in the Cartesian coordinates that are normalised to the wheel base,
where N = (nx , ny ) and C = (cx , cy ), Equation (2.6) becomes

c˙x = nx 2 ṅx − 2nx ṅx cx + nx ny ṅy − nx ṅy cy + ṅx cx 2 − ny ṅy cx + ṅy cx cy , (2.7)
2 2
c˙y = nx ṅx ny − ṅx ny cx + ṅy ny − 2ny ṅy cy − nx ṅx cy + ṅx cx cy + ṅy cy . (2.8)
2.3. Dynamic Methods - Modelling and Simulation 15

The angle δ represents the steering angle, or the angle between the longitudinal axis of the
aircraft and the longitudinal axis of the tug, and is given by

!
Ṅ · (N − C)
δ = cos−1 . (2.9)
|Ṅ |

A steering angle of zero indicates that the nose gear axle is perpendicular to the longitudinal
axes of the aircraft. In Chapter 3 we will show how kinematic methods can be used for the
analysis of towing stability, and in Chapter 5 kinematic methods are used for steering angle
and clearance distance predictions.

2.3 Dynamic Methods - Modelling and Simulation

The models that are used at Airbus are built with different test platforms in mind. MSC.Adams
models are used for detailed ground manoeuvrability studies, while SimMechanics models
are used on the test rigs, where the real-time performance of the models is critical. The
MSC.Adams environment is user-friendly and is the preferred model development environ-
ment. MSC.Adams models are then converted to SimMechanics for testing with the avionics
that will be implemented on the aircraft. Figure 2.4 shows a typical MSC.Adams model of the
A380, with a specific focus on the nose landing gear, while Figure 2.5 contains the SimMe-
chanics representation. Similar models exist for the A320 and are used in exactly the same
way. The following sections show how the models are constructed and used. Chapters 3, 5 and
6 show how simulations are used for the analysis of ground manoeuvres.

2.3.1 Model Construction

The first step in the model building process is to describe the rigid parts and the joints connect-
ing the parts [9], where a part is described by its mass, inertia and orientation. A right-handed
coordinate axis system is used. From a pilot’s perspective, the x-axis is in the forward direction
along the fuselage, the y-axis to the right, and the z-axis downward. The same (local) coor-
dinate system is used for the tyres. The calculation of the aerodynamic angles of the aircraft,
and the slip-angles on the tyres are straightforward when these conventions are used for the
local coordinate systems. The nose gear is constrained by a cylindrical joint, which is driven
by an angular motion, as depicted in Figure 2.4. Each type of joint has a number of associated
degrees of freedom. For instance, a prismatic joint represents one degree of freedom (linear
translation). Two states are present, namely a translational displacement and a translational
velocity. A cylindrical joint contains a translational and rotational degree of freedom, hence
four states. Torsional flexibility of the shock absorber (also known as an oleo) is important
for a landing gear with a bogie. Consequently, the oleos of the A380 model contain rotational
joints that are constrained by rotational springs, representing the stiffness of the torque link.
The aft-axle steering inputs are inserted as motions from a control law. Table B.1 in Appendix
B contains a list of the components with the constraints (and states) associated with each com-
ponent for the Airbus A320 and A380 aircraft. The A320 model contains a total of 18 states,
and the A380 model contains 38 states.
16 Chapter 2. Models and Hierarchy of Analysis Methods

Oleo Stiffness
Oleo Damping

Angular Motion
Cylindrical Joint
Tyre Force
Self-aligning
Torque

Tyre Forces
Vertical
Longitudinal
Lateral

Figure 2.4. Detailed definition of nose gear components pertaining to an A380 MSC.Adams model.

0:0

Figure 2.5. Top-level SimMechanics model of an A380.


2.3. Dynamic Methods - Modelling and Simulation 17

The next step in the building process is the addition of internal force elements to represent the
shock absorbers and tyre forces, known as line-of-sight forces, which act between two parts [13,
54, 9]. The oleos consist of nitrogen and oil, where the compression of the nitrogen provides the
cushioning effect, while the flow of the oil through the orifice plates provides the damping. The
characteristics of the oleo are critical in the case of an aircraft with a statically indeterminate
gear arrangement, such as the A380, where small changes in the oleo characteristics could lead
to significant changes in the loads that are transferred to the fuselage. The tyres are modelled
with impact functions that switch on as soon as the distance between the wheel centre and the
tyre becomes less than the wheel radius. External forces such as thrust and aerodynamic forces
are then added, and they are known as action-only forces.
When building the model with SimMechanics, extensive use is made of the new object oriented
features in Matlab, where all geometric aspects are parameterised; from the axle widths, wheel
dimensions, gear positions, to the rake angles on the gears. This means that all joint definitions
and forces are automatically updated when the design variables are changed. This has added to
the ease of use and robustness of the models — an ideal situation for industrial use. The user
can enter one command that will configure the thrust, steering and braking configurations.

Tyre Modelling

Apart from the aerodynamic, propulsive, and gravitational forces, all other loads on the aircraft
are applied at the tyre-ground interface. Tubeless radial tyres are generally used for aircraft due
to better failure characteristics when compared with bias-ply tyres [73]. The forces generated
by the tyres have a dominant effect over the aerodynamic forces at low velocities. The vertical
force component of the tyre can be approximated by a linear spring and damper system [9, 54,
55], where the total vertical force is described as
p
Fz = −kz δz − cz Vz = −kz δz − 2ζ mt kz Vz . (2.10)

Here mt is the tyre mass, Vz is the vertical velocity of the tyre, and δz is the tyre deflection
representing the change in tyre diameter between the loaded and unloaded condition. Stiffness
kz and damping ζ are determined from experiments. They are usually provided by the tyre
manufacturers to the airframe OEMs. Several theories exist for the rolling resistance of a
wheel, of which the following explanation seems the most plausible [9]. Rolling resistance on
hard surfaces is caused by hysteresis in the rubber of the tyre, where the pressure in the leading
half of the contact patch is higher than in the trailing half. A horizontal force in the opposite
direction to the wheel movement is needed to maintain an equilibrium, and it is known as the
rolling resistance [73]. The ratio of the rolling resistance Fx to vertical load Fz on the tyre is
known as the coefficient of rolling resistance µR , where a value of 0.02 is typically used for
aircraft tyres [44]. The models implement an adapted Coulomb friction model that is smoothed
around the stationary point, as given by

Fx = −µR Fz tanh (100Vx ). (2.11)

Lateral motion is generated by directing the tyre at an angle to the direction of motion, leading
to a lateral force. This angle, α, is known as the slip-angle; see Figure 2.6(a). The relation-
ship between the lateral force and the slip-angle is linear for small slip-angles; it is usually
18 Chapter 2. Models and Hierarchy of Analysis Methods

100
200
direction 156
112
of motion 50 68

Fy (kN)
25
Vy
0
Vx
a-slip angle
−50

−100
−180 −90 0 90 180
α (deg)
(a) Slip-angle definition. (b) Lateral tyre forces for different vertical loads. Each curve
represents a vertical load whith units of kN.

Figure 2.6. Relevant tyre quantities.

defined by a cornering stiffness coefficient in the automotive industry [9], where maximum
slip-angles of 5◦ seem to be the norm [9]. Slip-angles on aircraft often go up to 90◦ during nor-
mal manoeuvres, making it necessary to define the tyre properties over all possible slip-angles.
Figure 2.6(b) contains a definition for the lateral force over the entire range of slip-angles for an
aircraft tyre at different vertical loads. The lateral load factor is another useful parameter that
is used to describe how hard the tyre is working; it is defined as the ratio between the lateral
and vertical force.
Braking is implemented by adding a brake force term to the longitudinal force. The available
lateral force that the tyre can generate is reduced when braking is applied; this is taken into
account by using the concept of a traction circle [73]. The resultant force falls within the
traction circle and reaches a maximum at the circle boundary.

Oleo Modelling

The main function of a shock absorber is to dissipate energy during landing and taxiing, so that
the forces that are introduced into the airframe are within operating limits [17]. Large passenger
aircraft tend to have oleo-pneumatic shock absorbers, due to the superior efficiency-to-weight
benefit that these systems provide [17]. The gas in the upper chamber acts as a spring when it is
compressed. A diaphragm or a piston can be used to separate the oil and the gas, otherwise they
are left to mix. Energy dissipation takes place at orifices, which act as the damping element of
the shock absorber.
A level attitude is desired when the aircraft is standing on the runway, and therefore the static
load should be calculated for the maximum aircraft weight, at the fore and aft CG positions.
The extended stroke lengths are calculated from the aircraft geometric considerations, and then
an initial estimate is made of the stroke that is required, based on previous aircraft. Compres-
sion ratios are then chosen based on experience, where a static to extended ratio of 4:1 and
a compressed to static ratio of 3:1 are generally used [17]. The spring force Fko can then be
2.3. Dynamic Methods - Modelling and Simulation 19

1,400 1,400
Compression
1,200 1,200 Rebound
1,000 1,000

co (kN.s2 /m2 )
Fko (kN)

800 800

600 600

400 400

200 200
(a) (b)
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
lo (m) lo (m)

Figure 2.7. Oleo properties. Panel (a) contains the stiffness force as a function of the stroke. Panel (b)
shows the damping coefficient as a function of the stroke.

calculated by multiplying the pressure inside the piston by the piston area. Figure 2.7(a) shows
the spring curve that is used for the nose gear of the A380.
Damping is provided when the oil moves through the orifices within the orifice block and the
recoil rings, where the damping force is dependent on the direction of motion. The damping
force Fco is calculated from

Fco = co (lo )Vo 2 , (2.12)

where the damping coefficient co is a function of the oleo stroke lo . Figure 2.7(b) shows the
damping coefficients for the nose gear of the A380. The combined force in the oleo Fo is then
calculated as

Fo = Fko (lo ) − Fco (lo , Vo ). (2.13)

2.3.2 Normal and Towing Operations

The same model is used for the analysis of normal and towing operations, where the model
is easily reconfigured for different scenarios. The difference between normal operations and
towing lies in where the force is being applied. The tyres are used to generate the required yaw
moment in the cases where the aircraft is self-propelled or where a tug with a towbar is used.
In the towbarless tug scenario the yawing moment is created by the tug. Only the aircraft is
modelled, while a tug is represented by a force at the nose gear position. This is due to the
difficulty of obtaining tug data from the manufacturers, and this is sufficient for landing gear
design purposes, where the loads into the gear are of interest. However, a detailed model of
the tug would be needed if the dynamic stability of an aircraft-tug combination is desired for
varying taxiway conditions.
20 Chapter 2. Models and Hierarchy of Analysis Methods

120
Test Data
Simulation
100
t0
a

80
y (m)

60

40

20
20 40 60 80 100 120 140 160
x (m)

Figure 2.8. Nose landing gear trajectory comparison between test data and simulation results. The
starting point of the comparison is denoted by t0 .

2.3.3 Model Validation

Aircraft models are built in the unloaded position and simulations are typically run with the
initial conditions of the model all set to zero, with the exception of the initial forward velocity
of the aircraft and the height of the aircraft above the ground. A velocity controller is then used
to accelerate the aircraft along a desired velocity profile at the NLG position. The force that
is used to control the velocity can either be provided by the engines or by a tug. The first few
seconds of the simulation are therefore used to “settle” the aircraft into its loaded condition, and
then the velocity controller is used to drive the simulation to the desired target initial conditions.
This is especially important when comparisons are made with test data, where a match between
the initial states of the test data and the simulation is critical. Figure 2.8 shows a comparison of
the nose gear trajectory between a test and a simulation, for a U-turn manoeuvre. The starting
point for the comparisons is denoted by t0 . At this point the model states need to match the
states from the test. Steering, differential braking and engine inputs from the test are used
as inputs into the simulation. Figure 2.8 shows only negligible deviations between the test
data and simulation results, hence we can conclude that the model gives sufficiently accurate
predictions for ground manoeuvrability performance.

2.3.4 Computational Challenges of Simulations

Many software packages use Lagrangian dynamics as a basis for developing computational
algorithms for the dynamic analysis of multibody systems [60], where the final set of equations
consists of differential and algebraic equations of index 3, which are considered to be of high
complexity and, as a result, costly to solve [47]. This is the reason why multibody calculations
tend to be difficult, and why alternative methods are useful. A comparison of the equations of
motion for rigid bodies, when derived from Newtonian and Lagrangian mechanics, leads to the
same set of equations. However, the Lagrangian form does allow for the analysis of flexible
2.4. Bifurcation Methods 21

bodies, leading to simulations that are even more computationally intensive when structural
modes are included into the model. A detailed description of the calculations that are involved
in multibody dynamics is not within the scope of this paper; the interested reader can obtain
more details from references [9] and [60].
Design of experiments (DOE), where the design space is divided into a grid of different combi-
nations of steering angles and velocities, provides a means to determine the effect of parameter
changes on the dynamics of a system. All the dynamic effects are taken into consideration,
and this leads to more reliable predictions. The user is able to test detailed steering and brak-
ing control-logic algorithms, balancing the aircraft performance against the loads on the gear
and tyres. However, it does not necessarily mean that all the dynamics have been categorised,
especially for highly nonlinear systems. A penalty is also incurred due to the difficulty in
automating the testing of such manoeuvres, as well as the high CPU times required for such
simulations. Simulations are conducted at very specific operating conditions during the concept
phase for trade-off studies. Small parts of the envelope are covered. These initial simulations
are also used in support of bifurcation analysis predictions, where bifurcation methods allow
for complete coverage of the envelope. Extensive simulations are conducted in the later stages
of a major aircraft programme, as, and when, data becomes available.

2.4 Bifurcation Methods

The high cost associated with simulations makes numerical continuation techniques attrac-
tive, due to the speed with which a global picture of the dynamics can be constructed. Spe-
cific regions of interest can be identified for further detailed analysis with multibody dynamic
codes [13, 15, 54]. Previous studies of the A320 showed how bifurcation methods can be used
to detect stability margins, showing how specific bifurcations can be attributed to the loss of
grip at specific tyres on the aircraft [13, 54, 55]. The Dynamical Systems Toolbox that was
developed at the University of Bristol allows for the seamless integration of Simulink or Sim-
Mechanics models [15].
It is important to note that the simulation models that are discussed earlier are in fact also
used for the bifurcation analysis. This is a very useful feature, as these models are likely to
be developed in other parts of the company. Hence it is possible to “plug” existing models
into the bifurcation analysis framework provided by the toolbox, avoiding the rebuilding of
models specifically for the purpose of bifurcation analysis. Another benefit of the toolbox is
the additional information that can be obtained from the models. All the tyre data is available,
which allows for the construction of supplementary information that would normally not be
readily available [15]. It is for instance possible to represent the data in new ways that gives
one a much better understanding of how the loads are distributed amongst the tyres. Figures 6.9
and 6.10 in Section 6.4.3 are a good example of new types of graphs that can be used to depict
a global view of the force distribution in the tyres as a result of using numerical continuation
techniques.

2.4.1 Bifurcation Theory

Dynamical systems theory provides a methodology for studying systems of nonlinear ordinary
differential equations. A key method is that of bifurcation analysis, where one identifies differ-
22 Chapter 2. Models and Hierarchy of Analysis Methods

ent ways in which the dynamics of the system can change. In combination with the numerical
technique of continuation, one can perform a nonlinear stability analysis by following solutions
and detecting their stability changes (bifurcations). The bifurcations can then be followed in
more parameters to identify regions in parameter space that correspond to different behaviour
of the system. See, for example [29] and [62] as entry points to the literature.
To summarise some basic ideas consider an ODE model of the form

u̇ = f (u, λ). (2.14)

where u is an n-dimensional state vector, λ a (multidimensional) control parameter, and f a


sufficiently smooth (typically nonlinear) function. In terms of standard equations of motion
for an aircraft on the ground, the state vector u contains the aircraft translational and rotational
states, along with the translational states of the oleos, as described in Section 2.3.1 and Ap-
pendix C. The control parameter consists of the steering angle, thrust, the position of the CG,
and possibly other relevant parameters. Equilibrium solutions of (2.14), also known as trim
conditions, satisfy

f (u0 , λ) = 0. (2.15)

The implicit function (2.15) defines a solution locus of equilibria, which is a one-dimensional
solution curve when a single parameter, such as the steering angle, is varied. The stability of the
equilibria can be determined from the (n × n) Jacobian matrix Df of partial derivatives of the
function f with respect to the state u. Continuation software, such as the package AUTO [19],
or the Dynamical Systems Toolbox used here, is able to follow curves of equilibria while
monitoring their stability. See also [37] for an overview of the different software packages that
are available. Changes of stability, that is, bifurcations, are automatically detected and can then
be followed in additional parameters. Similarly, periodic solutions can be followed and their
stability changes detected. The continuation of suitable solution curves allows one to build up
a comprehensive picture of the overall dynamics in a systematic way.
Typical bifurcations such as saddle-node (fold) and Hopf bifurcations (onset of oscillations)
can be found in engineering systems. Previous work on ground manoeuvring has indeed found
oscillatory behaviour at higher velocity and thrust ranges [13, 54, 55]. Bifurcation analysis is
now a standard and powerful tool that is being used extensively in engineering applications, and
more recently for the analysis of landing gears and aircraft ground dynamics [13, 54, 55, 64].

2.4.2 Dynamical Systems Toolbox — AUTO Integration into Matlab

Bifurcation methods have not been readily adopted by the engineering community because
the methods and tools available have thus far been developed and used mainly within an aca-
demic environment. The development of a Dynamical Systems Toolbox within the Matlab
environment is our attempt to consolidate previous efforts at the University of Bristol to create
a user-friendly environment for engineers [16]. Other efforts around the world to develop dy-
namical systems software in Matlab exist, such as MATCONT [28], but it appears that this has
not been widely adopted by the engineering community. We have thus tried to obtain the best
2.4. Bifurcation Methods 23

of both worlds by integrating the existing Fortran AUTO code into Matlab via mex-functions.
This allows us to use the speed of a lower level language with the user-friendly interface of
Matlab, along with access to the existing algorithms available in AUTO.
Another important aspect of the toolbox is that engineering tools such as Simulink and SimMe-
chanics can be integrated with the dynamical systems software. In this way, industrially tested
models can be used directly in this environment — without the need for converting models to
a format that can be used by the stand-alone version of AUTO. More specifically, AUTO has
direct access to the states of the Simulink/SimMechanics model.
More widespread use of the Dynamical Systems Toolbox will be promoted by providing doc-
umentation and reference material that is easy to use, with concrete examples for the user. We
have combined most of the user manual of AUTO with our own examples, and integrated this
into the Matlab help environment. The Dynamical Systems Toolbox therefore feels like any
other toolbox that has been developed for Matlab, where the user can select the toolbox from
the menu, with the accompanying help and search functionality. We have also started to de-
velop components with the new object oriented programming capability in Matlab, which will
enhance the pace at which new applications will be developed in the future. Appendix D can
be consulted for more details on how the toolbox was developed and how to obtain the source
code.

2.4.3 Application to Ground Manoeuvres

The continuation analysis starts from an equilibrium solution that is obtained from a previous
bifurcation run, or from a simulation. The method used here is in fact the latter, where a velocity
controller is used to control the engine or towing force until a target velocity is reached. At
this stage the steering angle is set to 0◦ , and consequently the aircraft is moving in a straight
line. Similar steps to that of the simulation approach are followed up to this point. The velocity
controller is then switched off so that the engines are operating at a constant thrust.
A continuation run is initiated as soon as the software recognises that an equilibrium condition
has been reached (zero gradients for the continuation states). Instead of feeding the steering in-
put into an ODE solver, the steering input is provided to the Dynamical Systems Toolbox. The
algorithm then follows these equilibrium states as a parameter (e.g. steering angle) is varied,
while their stability is monitored. The occurrence of fold and Hopf bifurcations [37] is moni-
tored, and they indicate a change in the dynamics when detected. All the tyre and gear forces
are provided for each equilibrium state, where these forces represent those needed to obtain
trim conditions. This provides an immense amount of freedom in the design process, and al-
lows one to follow any solution of interest in the relevant parameters. The user can for instance
set a specific condition (algebraic constraint) on the tyre forces, and then follow this condition
directly, without having to do exhaustive simulations, to obtain the envelope for the prescribed
condition. See [13] and [53] for an exact explanation of how the analysis is conducted.
We use continuation algorithms in Chapter 3 to obtain steady-state turn radii as the steering
angle is varied. In Chapter 4 we show how continuation methods can be used to determine
if sustained towing manoeuvres will have a beneficial or detrimental effect on the tyres. In
Chapter 5 we show how the steering angle that is needed for a runway exit can be obtained
directly from continuation analysis, instead of using simulations. In Chapter 6 we show that
24 Chapter 2. Models and Hierarchy of Analysis Methods

steady-state conditions can be used to analyse the lateral ground loads for a high-speed turn,
making continuation methods the perfect tool for such analysis.
Chapter 3

Low-Speed: U-turn Manoeuvres

3.1 Introduction

The Boeing 747 has long been used as the baseline for specifying requirements to which large
international airports must adhere. This means that new civil aircraft designs have to stay
within the manoeuvrability requirements of this aircraft to ensure that no significant investment
is needed for upgrades to existing airport infrastructure. This is also the case for the Airbus
A380, for which only terminal facilities need to be upgraded, while runways and taxiways do
not require any significant alterations.
The most basic and widely used analysis techniques for early design use purely geometric
and static relationships between the gear positions to determine the turn radius of the aircraft.
However, it can be shown that the centre of gravity (CG) location, tyre and brake properties
do in fact play a significant role in the overall performance, which means that the static gear
layout alone does not determine the possible turn radius. In this chapter we specifically focus
on the U-turn manoeuvre, and analyse the way in which this manoeuvre is conducted. Not all
the parameters that define a U-turn manoeuvre are entirely understood due to the demanding
nature of this manoeuvre, especially for large aircraft such as the A380.
One important property of the overall aircraft model lies in the nonlinear nature of landing gear
components, for example, due to geometric effects or large tyre deflections. Therefore, small
perturbations in velocity, steering angle or brake application may lead to significant differences
in the final turn width. In a previous study we showed that it is possible to calculate the radius
of turn for a specific thrust case as the steering angle is varied [13]. In this chapter a comparison
is made between the most widely used methods for the analysis of U-turn manoeuvres, which
consist of geometric and simulation approaches, and a new approach where dynamical systems
theory is used to determine the turn width. The advantage of the latter is that it combines the
necessary accuracy with computational efficiency.
It is shown that the feasible region of a U-turn is defined by an algebraic constraint, where the
desired turn width forms a boundary. Bifurcation methods are used to follow this constraint
as parameters, such as thrust and steering angle, are varied. From the practical point of view,
the size and location of the feasible region give a clear picture of whether or not a ground
manoeuvre can be conducted. The bifurcation diagrams considered in this study encapsulate
all the information that a design engineer would need in terms of turn widths, edge-clearance

25
26 Chapter 3. Low-Speed: U-turn Manoeuvres

distances, operating velocities, and steering angles. Therefore, bifurcation analysis provides
an additional tool that can significantly enhance insight into the parameters that influence the
U-turn performance of the aircraft, and so may contribute to a more mature product when flight
testing commences.

3.2 The U-turn Manoeuvre

A U-turn can be conducted in one of two ways. The first method is called an Edge-of-Runway
(EOR) manoeuvre; it is conducted by placing the aircraft parallel to the side of the runway and
then initiating the turn at any point. The second method is called the Centre-of-Runway (COR)
manoeuvre; it is conducted by starting from the middle of the runway, traversing to the side of
the runway at an angle, and then initiating the turn as soon as the nose gear reaches the edge of
the runway. The COR method tends to allow for larger turn margins due to a shift in the centre
of rotation towards the edge of the runway. Figure 3.1 depicts the two different approaches.
Only the EOR method will be discussed, as the methods are essentially the same, apart from
the initial starting points.
The steps for conducting the U-turn are:

1. To align the aircraft with the edge of the runway. The pilots will leave some space
between the gears and the edge of the runway, but for the purposes of the simulations the
outer plane of the outer wing gear tyres are aligned with the edge of the runway.

2. Set the aircraft in motion by applying thrust to all the engines.

3. Increase the engine thrust on the outboard engine while decreasing the thrust on the
inboard engines. These two actions are done at the same time in the simulations, whereas
they will most likely be done separately in reality.

4. Apply the brake pedal on the side that the aircraft will turn into, meaning that the left-
hand pedal is used for anti-clockwise turns, while the right-hand pedal is used for clock-
wise turns.

5. Set the nose gear steering angle with a ramp input over a time period of approximately 5
seconds, and then the steering angle is held constant.

It is worth noting that the above steps are implemented in this order for the purpose of the
simulation, but these steps can in fact be performed in a different order or can be combined.
It is proposed that the fourth and fifth steps are the most important, and will be similar for the
EOR and COR manoeuvres. Therefore, only the EOR manoeuvre is analysed, from which the
COR turn width follows. Figure 3.2 contains the dimensions of importance for a U-turn, which
feed into the turn width ltw given by

ltw = lm + lt + rm + rn . (3.1)

Here, the outer width between the outermost gears is represented by lm , and lt represents a
transition distance. This transition distance is a function of the nose gear velocity, steering rate
3.2. The U-turn Manoeuvre 27

Figure 3.1. Edge of runway (EOR) and centre of runway (COR) U-turn manoeuvres.

rn lm

ltw
rm
lt
lm

Figure 3.2. U-turn manoeuvre dimensions.

and final steering angle. It is assumed that the aircraft makes a steady turn, hence the radius
does not change after the transition period. Two radii are thus of importance. The first is the
radius of the nose gear rn , while the second is the radius of the inner gear rm . The reference
point for the inner gear is however not located at the bearing point between the strut and the
bogey beam, but a point that is offset by half an axle- and half a wheel-width from the bearing
point, towards the centre of rotation. This definition is explained in Figure 2.3(a). These radii
can be altered by using different thrust and braking combinations.
The turn width from the COR method can be calculated by subtracting a geometric distance
from the EOR turn width solution. The maximum additional distance lcor that can be acquired
by using the COR procedure can be obtained by

lcor = (lm + lt + rm )(1 − cos (ξ)), (3.2)

where ξ represents the angle that is formed between the fuselage longitudinal-axis and the edge
of the runway, as indicated in Figure 3.1. The optimum angle would be the angle that is formed
between the fuselage longitudinal-axis and a line that is drawn between the nose gear and the
outer main gear.
28 Chapter 3. Low-Speed: U-turn Manoeuvres

rn

rm

Figure 3.3. Geometric approach for finding the turn radius.

3.3 The Geometric Approach to the U-turn

The geometric method seems to be the most widely used method for calculating the turn radius
of aircraft [11]. Automotive references go further by recognising that the radius of turn is
dependent on the velocity of the vehicle, and consequently some assumptions are made with
regards to the slip-angles that can be generated [73]. These slip-angles then contribute to the
forces in the tyre and the resulting radius of turn. The main assumption is however that the
slip-angles remain small (< 5◦ ), and consequently a lateral stiffness coefficient can be used.
A formula can then be used to generate the turn radius. This approximation is however not
sufficient for aircraft, as the tyres can operate at very large slip-angles (up to 90◦ ) where linear
approximations cannot be made. The difficulty of obtaining an accurate dynamic formula might
be the reason for the wide-spread use of the geometric method.
An example of the geometric method can be found in the paper on landing gear design by Chai
and Mason [11], where it is proposed that the centre of rotation lies at the intersection between
a line that is drawn perpendicular to the mean distance between the main gear posts, and a
line extended from the nose gear axle. Figure 3.3 depicts this approach, which is generally
used in the initial concept stages of an aircraft programme. The centre of rotation is highly
dependent on different thrust and braking combinations, which are completely ignored within
this geometric approach. Other information such as tyre forces are also not available. The
final turn width can then be calculated with Equation (3.1), while assuming a constant factor
for the transition distance lt . Geometric methods might be sufficient for small aircraft with
tricycle arrangements [17], but will not be sufficient for larger aircraft, especially those with
more landing gears.

3.4 U-turn Results from Simulations

We focus on the low-speed turn width solutions over a range of fixed nose gear velocity magni-
tude levels (1 − 5 m/s). The velocity magnitude of the nose gear and CG will thus be the same
3.4. U-turn Results from Simulations 29

when no steering input is given. In contrast, the CG velocity will drop as the steering angle is
increased, while the nose gear velocity is being maintained. Asymmetric thrust configurations
are of most interest, seeing that such configurations provide the smallest possible turn radii. A
proportional-integral thrust controller is used on the outboard engine to ensure that the nose
gear velocity is maintained, while the other engines are set to idle-thrust.
In reality the pilot will not set an exact fixed velocity, but this is assumed for automation and
comparison purposes. The gains of the controller are set in such a way as to represent the
manner in which the pilot would operate the engine. The nose gear velocity will drop as the
aircraft enters the turn, after which the pilot will increase the thrust to maintain the velocity.
This thrust increase usually leads to an overshoot of the desired velocity. The pilot corrects the
overshoot by decreasing the thrust. The rest of the steps with regards to how the manoeuvres
are conducted are as explained in Section 3.2.
A range of steering angles (20◦ − 80◦ ) and nose gear velocities (1 − 5 m/s) were used for a
medium weight A380 configuration. The selection of test points is a balance between capturing
the most important phenomena and simulation run-times. The steering angle was increased by
5◦ increments, and the velocity by 0.2 m/s. The simulation run-time for this example was
approximately 2.5 hours on an Intel 1.8 GHz processor. Greater fidelity can be obtained by
using more test points, but this would be at a significant cost to the run-times. The nonlinear
nature of tyres does however mean that areas of rapid transition may occur between certain
velocities and steering angles, and consequently the mesh would need to be refined in these
areas.
The results for the turn width simulations in SimMechanics are shown in Figure 3.4(a), where
the maximum steering angle (δ) during the simulation is represented on the x-axis, and the
nose gear velocity magnitude (Vn ) at entry into the turn is represented on the y-axis. The
resulting turn width, or any quantity of interest, is therefore a result of two independent inputs
(δ and Vn ), and can be represented readily by a contour plot. The area to the right of the 51
m contour line indicates the feasible operating region for a U-turn. This value is obtained by
subtracting the required clearance distance (2×4.5) from the 60 m runway width for a Category
VI airport; the type of airport where an A380 can conduct a U-turn manoeuvre. The turn width
increases slightly with an increase in velocity at a specific steering angle, but the simulations
seem to indicate that the turn width is not greatly influenced by the velocity for this specific
case. Figure 3.4(b) contains a contour plot of the turn radius of the nose gear as a function of
different steering angles and nose gear velocities.
The transition distance is shown in Figure 3.4(c) and is highly dependent on the actions of the
pilot. Reaction times were obtained from flight test data, indicating that the steering input took
place over the course of 5 seconds. The biggest transition distance occurs at a steering angle of
70◦ and a velocity of 4 m/s, and could indicate an area where a transition of loads between tyres
takes place. This would have to be verified by studying the tyre forces from the simulations.
Section 3.2 shows that an additional turn margin can be obtained by using the COR instead
of the EOR method. It can be computed by using Equation (3.2), and this distance is shown
in Figure 3.4(d). If it is assumed that the 60 m contour line in Figure 3.4(a) indicates the
boundary of the feasible region, then Figure 3.4(d) indicates that an additional margin of up to
approximately 1.0 m can be obtained by using the COR method.
30 Chapter 3. Low-Speed: U-turn Manoeuvres

5 5
83

38
37
40
81

43
4 δ for feasible U-turn 4
83

31
50
Vn (m/s)

31
82
82

3 3
55
60

39

36
35
34

32
42
41

33
70
65
75

43
80
81
80

2 2

38
37
40
42
81

43
45

31
31
50

(a) 2 (b)
1 - 1
20 30 40 50 60 70 80 20 30 40 50
1.6 60 70 80
5 5
5
0.

-0.5
-0.5 0
1.5
2

0.8
0

4 4
1

1
Vn (m/s)

1.2
1.4
1.6
2

3 3
1.8

0.5
0.5

-1.5
0

0.8

2 0 2
-0.5

-2

(c) (d)
-1

1 1
20 30 40 50 60 70 80 20 30 40 50 60 70 80
δ (deg) δ (deg)

Figure 3.4. U-turn performance from simulations for a medium weight aircraft. The independent
parameters are the maximum final steering angle δ and the velocity magnitude Vn at the nose gear,
while the dependent parameter is indicated by the contour lines. Panel (a) contains the turn width ltw ,
while panel (b) contains the turn radius rn of the nose gear. Panel (c) contains the transition distance lt ,
and panel (d) contains the additional margin lcor for the COR U-turn.
3.5. U-turn Performance Using the Bifurcation Approach 31

rn
Vn (xn,yn)

rn
Vm wz rm
(xm,ym)
rm
(xo,yo) (xo,yo)

Figure 3.5. Turn radius and centre of rotation.

3.5 U-turn Performance Using the Bifurcation Approach

We propose a method that uses bifurcation analysis to determine relevant dynamic quantities
in dependence on key parameters. This then feeds into a geometric model where we make
some reasonable assumptions about the transient dynamics. More specifically, the bifurcation
analysis provides the steady state solutions for the velocity magnitudes Vn and Vm at the nose
and wing gear, as well as the yaw rate ωz . The turn radius for the nose gear (and similarly for
any other point on the aircraft) can then be calculated as

Vn
rn = . (3.3)
ωz

Figure 3.5 depicts the most important dimensions in the calculation. The centre of rotation
(xo , yo ) in the left-hand figure is not readily obtainable from the radii alone, but can be cal-
culated if the directions of the velocity vectors (Vn and Vm ) are known. This does however
lead to unwieldy geometric calculations. An alternative approach is to draw the locus (which
forms a circle) of the possible turn centre solutions for each respective reference point. The
outer intersection point relative to the fuselage centre line is then the solution for the turn cen-
tre. The right-hand figure shows how this method is implemented. The turn width can thus
be calculated by using Equation (3.1), where accurate information is available for all the vari-
ables, apart from lt , the transition distance. It is proposed that upper and lower bounds are
chosen for this distance due to the variability of this value, which will then give the engineer
a clear indication of what a best- and worst-case turn width would be. The transition distance
can be determined by simulations or from test data, but should be representative of the aircraft
response due to normal pilot inputs. A distance of 1 m is assumed in this case, informed by the
simulation results in Figure 3.4(c).
32 Chapter 3. Low-Speed: U-turn Manoeuvres

5 5
83 82

31
83

65

38
37
36
75
70

50

42
41
40
39

35

32
60
55

34
33
81
80

43
43

31
4 A B 4

δ for feasible U-turn


82
81 80
Vn (m/s)

31
3 3

31
82

38
37

32
75
70
65
60
55
50

41
40
39

36
35
34
33
43
80

42
81

43

31
2 2
81

42
(a) L (b) L

31
80

1 1
20 30 40 50 60 70 80 20 30 40 50 60 70 80
δ (deg) δ (deg)

Figure 3.6. U-turn performance from bifurcation analysis for a medium weight aircraft. Panel (a) shows
the turn width ltw , and panel (b) the turn radius rn of the nose gear.

Figure 3.6(a) contains the turn width results that are obtained when Equation (3.1) is used. This
leads to a conservative estimate at velocities below 4 m/s, the region where a pilot is expected
to operate. The turn width initially increases and reaches a maximum at approximately 30◦ ,
after which it drops rapidly with an increase in steering angle. The initial increase can be
attributed to the fact that the controller is trying to maintain a specific velocity at the nose gear,
while the nose gear tyres have not reached their optimum lateral loading condition yet. This
is not a realistic operational scenario, as the nose gear is essentially being dragged sideways
due to the low steering angles. The pilot would in fact operate the nose gear at larger steering
angles. It can be seen that there is good agreement between the turn width results from the
simulation in Figure 3.4(a) and those from the continuation analysis in Figure 3.6(a). The turn
radius of the nose gear is shown in Figure 3.6(b), and it again shows very good agreement with
the results from the simulations in Figure 3.4(b). The run-time for obtaining the bifurcation
diagrams was approximately 15 minutes on an Intel 1.8 GHz processor, which constitutes an
order of magnitude increase in the analysis speed compared to that of simulations.

The thick line in the figures at approximately 69.5◦ steering angle denotes the position of a fold
bifurcation, which indicates that a qualitative change in the dynamics takes place in this region.
Note that this phenomena was not easily observed from simulations, even though the results
were in fact present. The presence of a fold bifurcation allows us to identify a region of interest
where further simulations can be informative. Figure 3.7 shows the physical meaning of the
fold. Point A in Figure 3.6(a) corresponds to a forward movement of the inner wing landing
gear, while point B corresponds to a transition from a forward to a sideways movement of
the inner wing landing gear. This shows that a small change in the steering angle within this
region could have a significant impact on the dynamics of the aircraft, hence the operating
procedures can be updated to avoid this transition. The nose wheel steering angle could for
instance be limited to 68◦ . This would however mean a reduction in clearance distances, which
can be alleviated by adopting the COR method as the preferred method to conduct a U-turn
manoeuvre.
3.6. Turn Centre 33

o
A: δ = 68 Gear
movement forward o
A: δ = 68

o
B: δ =70
o
B: δ = 70
Gear movement
sideways

Figure 3.7. Comparison of aircraft behaviour at point A, which is to the left of the fold (δ = 68◦ ), and
point B, to the right of the fold (δ = 70◦ ). Point A corresponds to a forward movement of the inner
wing landing gear, while point B represents a sideways movement.

3.6 Turn Centre

The position of the turn centre is important as it can play a large part in the types of loads
that can be generated in the gears. This is especially true for larger aircraft, where a pivot-turn
(turn around the inner main gear) could for instance introduce large torsional loads on the inner
main gear. Design engineers also overlay the aircraft and turn centre positions onto airport
drawings, to determine whether the aircraft can manoeuvre around specific corners. However,
this geometric approach ignores the dynamic effects. Simulations are conducted at specific
airports to ensure airport compatibility, but these are often quite complex to set up if many
airports need to be considered. The bifurcation approach provides all the dynamic information
that is needed for accurate calculations of the turn radii and turn centre, and therefore it is
proposed that this method be used to obtain a more accurate estimate of ground performance
(as long as the transient effects are understood). It is assumed that the aircraft will be able to
conduct a safe turn if the calculated turn radius is smaller than that of the turn that needs to be
negotiated. Bifurcation diagrams also provide steering angles and velocities for safe operations
and can be added to the interpretation of the results.
Figure 3.8 compares the geometric and the bifurcation approach. By construction the geometric
turn centre lies on a straight line perpendicular to the main aircraft axis. This longitudinal
position is in good agreement with the calculated actual turn centres. However, the lateral
positions are substantially different for similar steering angles. Figure 3.8 clearly shows that the
turn centre prediction from the geometric methods are inaccurate for relatively small steering
angles. Note that the calculations for the exact values were done here for an asymmetric thrust
34 Chapter 3. Low-Speed: U-turn Manoeuvres

o o
Actual turn centre 2535o 75 o
o o o o o o o o
30 40 45 50 55 60 65 70 80
o o o o o o o o o o o o
25 30 35 40 45 50 55 60 65 70 75 80
Geometric turn centre

Figure 3.8. Position of turn centre for steering angles from 25◦ to 80◦ ; calculated by continuation and
by the geometric approach.

case with asymmetric braking, which cannot be included in the geometric calculations. The
geometric predictions are closer to the correct values for higher steering angles, but they are
still out by several metres. The longitudinal position at high steering angles does seem to be
in good agreement with the position mentioned by Chai and Mason [11]. We can therefore
conclude that it is imperative to calculate the actual turn radius as a function of the steering
angle by other than purely geometric means. Bifurcation analysis emerges as a practical tool
for this approach.

3.7 Discussion

We presented for the first time an in-depth analysis of the U-turn manoeuvre. A comparison
was made between a widely used geometric method, a simulation-based approach, and a bifur-
cation analysis approach. The geometric method uses purely geometric and static relationships
between the gear positions to determine the turn radius of the aircraft. It is a simple method
to use, but engine thrust, tyre and brake inputs are ignored. Hence, the computed turn radii
are generally not reliable as a result of the highly nonlinear nature of landing gear systems.
Namely, small perturbations in velocity, steering angle or brake application may lead to signif-
icant differences in the final turn width for the same basic geometry of wheel settings.
We showed how an industrially tested SimMechanics model is used for simulations of U-turn
manoeuvres. A medium weight case was chosen with asymmetric thrust and braking inputs.
The turn width results from the simulations were presented as contour plots, and showed that
the U-turn performance for the aircraft is well within the requirements for this particular con-
figuration. We then used the same model to demonstrate how bifurcation analysis can be used
to obtain turn width results that are sufficiently close to that of the simulations. The advantage
3.7. Discussion 35

of the bifurcation analysis approach is that it is more efficient (in terms of run-times) and is
also able to find qualitative changes in the dynamics, fold bifurcations in this case, that are
not picked up easily from simulations. The presence of the fold bifurcation signifies a change
in the dynamics, where the inner wing landing gear transitions from a forward to a sideways
movement. Bifurcation diagrams encapsulate all the information that a design engineer would
need in terms of turn widths and edge-clearance distances (provided that an acceptable estimate
for the transition distance is available), operating velocities, and steering angles.
Overall, we conclude that the bifurcation analysis of ground manoeuvres would be suited to
initial and detailed design studies. Initial design studies could be used to define the tyre proper-
ties that are needed to ensure successful U-turn manoeuvres. Detailed design studies could be
used to verify that the tyres that are provided by the tyre manufacturers would in fact ensure the
success of these manoeuvres. Further studies could compare the results of other mass, engine
and braking configurations, and could also investigate the longitudinal movement of the turn
centre to determine the sensitivity of the turn centre to parameters such as the CG position and
tyre pressure.
Chapter 4

Low- and Medium-Speed: Towing

4.1 Introduction

Aircraft ground operations comprise of a combination of tug and self-propelled movements.


Tug operations tend to consist of pushbacks from the gate until the aircraft reaches a favourable
orientation for movement under its own power, or maintenance towing between the hangars
and the gates. Taxi operations, on the other-hand, are presently conducted with the aid of
the aircraft’s engines. One only has to venture to the nearest airport to notice that aircraft
often stand around for extended periods while waiting for take-off clearance, which leads to
unnecessary fuel consumption. Ideas related to the towing of aircraft to a starting grid close to
the runway threshold have been mentioned for the purposes of fuel savings [5]. Recent media
reports state that the use of tugs during taxiing will allow pilots to keep the engines turned
off until the aircraft reaches the runway, where the tug is then separated from the aircraft [5].
The opposite is also true after the aircraft has landed. From the literature available on remote-
controlled tugs, it appears that the pilot would control the tug from the cockpit [5], and that
a tug driver would manoeuvre the tug around the gate, meaning that no extra manpower will
be needed. Another advantage is that no physical modifications to the aircraft are necessary.
A less obvious advantage relates to the turn around time at the gate, which is often prescribed
by a threshold temperature of the brakes: the aircraft is not allowed to dispatch if the brake
temperature is above a certain value. A system that could minimise the use of the aircraft
brakes would therefore not only reduce the energy that is absorbed into the brakes, but will
also reduce brake and tyre wear and allow for shorter intervals at the gate.
This chapter addresses the issue of aircraft towing stability in Section 4.2, by building on
previous work related to the towing of truck-trailer combinations [25]; this also forms the
starting point for many other studies related to the control of towed vehicles [6, 61, 72]. The
focus in these papers was on straight-line and circular towing stability, which is also directly
applicable to a tug that is towing an aircraft. Airport apron and taxiway markings consist of
a combination of straight and circumferential lines, hence the analyses of straight and circular
manoeuvres are adequate. In this study we focus on towbarless tugs with a typical arrangement
depicted in Figure 4.1. The wheels are clamped in a cradle within the tug, and consequently the
relative motion between the aircraft and tug is around the rotation axis of the nose gear strut.
It is not clear if towbarless tugs are the preferred option for moving aircraft around airports,
compared to tugs with towbars, but towbarless tugs do have several distinct advantages:

37
38 Chapter 4. Low- and Medium-Speed: Towing

Figure 4.1. Towbarless tug arrangement.

• Less manpower. Only one person is needed for maintenance towing operations, elimi-
nating the need for a pilot in the cockpit.

• Lighter tow trucks. The weight needed for traction is provided by the aircraft nose gear
that rests on the tug, unlike a tug with a towbar, where weight needs to be added to the
truck.

• Improved health and safety. No man-handling of equipment such as towbars.

• No need for towbar storage. Each aircraft type has its own specific towbar, hence these
towbars need to be stored somwhere, until needed.

The apparent advantages of towbarless tugs have however been offset by several reports of
hazardous incidents involving such tugs [30], including jackknifing, and runway incursions.
Jackknifing occurs when the towing angle exceeds 90◦ . This has renewed interest into the
stability of towbarless tugs and aircraft. Detailed dynamic models of the tug and aircraft are
needed if the effects of low friction surfaces (such as oil on concrete) are to be studied. Unfor-
tunately the tug data that is needed for such analysis is not readily available, and consequently
we only focus on the stability of the kinematics of towing. The kinematics of the solution will
be similar for a towed aircraft and an aircraft that is being propelled by its engines. This is due
to the nature of the kinematic equations, where the application of forces are ignored. The mo-
tion when the aircraft is towed can consequently be described by the equations in Section 2.2.
Chapter 5 shows how these equations can be used for the self-propelled case at runway exits.
4.2. Kinematic Towing Stability 39

Section 4.3 shows that additional analysis is needed to obtain the tyre forces if this is of inter-
est; they can either be obtained from bifurcation methods or full-scale dynamic simulations.
Chapter 3 showed that geometric/kinematic methods were not sufficient for determining the
U-turn performance of very large aircraft [15]. Pushback and towing manoeuvres are usually
done at low speed, hence one of our main assumptions is that the maximum tyre forces at the
main gears do not exceed the steady-state values. In the following chapter we show that this
premise is also true for a large aircraft, such as the A380, when conducting a runway exit ma-
noeuvre. This assumption makes the bifurcation approach very useful due to the computational
efficiency of the method. Direct comparisons are made between a case where the aircraft is be-
ing towed along a circular trajectory (which represents a portion of a taxiway line), and where
the aircraft conducts such a manoeuvre under its own power. The analysis shows that the aft
tyres on the body landing gear experience higher loads when the aircraft is towed, which is
offset by a reduction of loads on other parts of the gears.

4.2 Kinematic Towing Stability

It can be shown that an aircraft that is being towed in a forward direction is inherently stable,
while an aircraft that is being pushed back is unstable [25]. This phenomenon is easily observed
around camp-sites, where one often sees a driver in trouble while reversing a caravan. The
implication of this instability is that a tug driver would be needed for any pushback operations
where a system such as Taxibot [5] is concerned, or that a control system would be required
to avoid jackknifing. Figure 4.3(a) depicts a forward towing manoeuvre with an un-jackknifed
configuration, while Figure 4.3(b) depicts a pushback towing manoeuvre that eventually ends
up in a jackknifed condition. A blue tug indicates an un-jackknifed configuration, while a red
tug depicts a jackknifed configuration. The detailed proof related to the stability aspects of
towing can be found in [25].
An aircraft will eventually settle into a stable circular trajectory if a steering input is applied,
as long as the towing radius rn is larger than the wheel base ln [25]. The relevant dimensions
such as the towing radius can be normalised to the wheel base for convenience [25], hence we
define a normalised turn radius, Rn = rlnn . Therefore the previous statement can be rephrased
as follows. An aircraft that is being towed with a towing angle larger than zero, will eventually
converge to a stable circular solution if Rn > 1. Two equilibrium positions exist for an aircraft
travelling along a circular trajectory, where the stable solution is depicted in Figure 4.2. If
the nose is placed at (−Rn√, 0) in the xy-coordinate frame, and the reference position of the
2 R2 −1
main gears at (− RRn −1
n
, ± Rnn ), the aircraft will start and maintain this trajectory if the
towing angle δ is set to the towing angle at equilibrium δf . The towing angle at equilibrium is
described as

p !
−1 Rn2 − 1)
δf = cos . (4.1)
Rn

Figure 4.3(c) depicts the case where the aircraft starts at a point away from the equilibrium
solution, and eventually settles into a constant radius turn, while Figure 4.3(d) shows the case
where the aircraft starts on the equilibrium solution. An unstable equilibrium position exists,
40 Chapter 4. Low- and Medium-Speed: Towing

Rn

1 Rn − 1
2

 R 2 −1 R n − 1 
2
− n ,−
 Rn Rn 
 

Figure 4.2. Stable equilibrium for circular trajectory.

√ 2 −1
2 Rn
as with a pendulum, if the starting point of the gear reference position is (− RRn −1
n
, Rn ),
while the final towing angle δf at this equilibrium position is set to

p !
Rn2 − 1)
δf = cos−1 + π. (4.2)
Rn

The aircraft can in principle maintain this specific radius, even though it is in an unstable
orientation. This configuration does however mean that the tug is in a jackknifed position. A
small disturbance will cause the aircraft to be pushed back, until a point is reached where the
steering angle is equal to 90◦ . The tug and aircraft eventually reach the stable solution. This
scenario is depicted in Figure 4.3(e). The last case of interest pertains to a scenario where the
tow radius is smaller than the length of the wheel base, that is Rn < 1. Figure 4.3(f) shows
that the tug and aircraft will alternate between an un-jackknifed and jackknifed position, as the
tug travels along the circle.

4.3 Load Factors due to Towing

One of the questions is whether such a radical change in operating procedures — the extensive
use of tugs — may have a significant impact on the gears and tyres. Direct comparisons of
the tyre forces can be made for a case (i) where a tug is used, and (ii) where the aircraft is
being propelled under its own power. We assume that towing operations are done with low
accelerations, and consequently the steady-state force values of the main gear tyres are used
for the comparison. In Chapter 2 we showed how these steady-state values can be obtained
4.3. Load Factors due to Towing 41

(a) Forward stable straight-line without jackknifing (b) Pushback unstable straight-line with jackknifing.

(c) Stable without jackknifing (Rn > 1). (d) Stable equilibrium without jackknifing (Rn > 1).

(e) Transition from jackknifed to un-jackknifed (f) Transition between jackknifed and un-jackknifed
stable solution (Rn > 1). positions (Rn < 1).

Figure 4.3. Towing solutions for straight and circular trajectories. A blue tug indicates an un-jackknifed
configuration, while a red tug indicates that the tug is jackknifed.
42 Chapter 4. Low- and Medium-Speed: Towing

60

40

20
∆Fy (%)

-20

-40

-60
W1 W2 W3 W4 W5 W6 W7 W8 W9 W10 W11 W12 W13 W14 W15 W16 W17 W18 W19 W20
Wheel Number

Figure 4.4. Difference between A380 lateral tyre forces for a towed aircraft and an aircraft using its
engines for propulsion. Positive values indicate that the towing case is more critical, while negative
values indicate that the engine case is more critical.

from continuation methods. A similar approach is used here. The only difference between the
two propulsion modes is the application point of the propulsion force. It can be argued that the
forces on the main gears will be the same for a tug with a towbar and a towbarless tug, due
to a similar constraint force that is needed to maintain the nose gear trajectory on the taxiway
line. The nose gear loads for a tug with a towbar are introduced through the towbar fitting and
the tyres, while the loads for the towbarless tug are introduced through the axle. We therefore
assume that the main gear loads will be equivalent, irrespective of the towing mode.
The maximum ramp weight (MRW) with an aft CG position is used. A towing angle is chosen
to maintain a typical towing radius that would be encountered at a Category VI airport. A radius
of 51m is therefore maintained at a velocity of 4 m/s, at the nose gear position. Figure 4.4
compares the tyre forces that are generated by towing, and the tyre forces that are generated
when the engines are used for propulsion. A positive value means that the force in the tyre
due to towing is larger than the case when engines are used, and the opposite for negative
values. A tug causes significantly higher lateral forces on the aft tyres of the body landing
gear, when compared to the self-propelled case. However, lower forces are experienced on the
forward and middle tyres of the body landing gear. Almost no difference exists for the tyres
on the wing landing gears. An explanation for this increase could lie in the fact that the towing
application point has a larger lever arm to the CG position, when compared to the lever arms
from the engine positions. A larger moment is therefore created around the y-axis (pitch) of
the CG, which is counteracted by the vertical forces on the aft tyres of the BLGs. The increase
in vertical force creates an accompanying increase in the lateral force on the aft tyres. We can
conclude that towing has a significant impact on the lateral tyre forces of the aft axle of the
BLGs, when compared to an aircraft that moves under its own power, and could be a design
case for the aft axles of the BLGs. This force increase is however offset by a reduction of forces
on other parts of the gears.
4.4. Discussion 43

4.4 Discussion

We presented a physical interpretation of towing manoeuvres, starting from the fact that push-
back manoeuvres are inherently unstable. We also concluded that jackknifing can be avoided
by maintaining a towing radius that is larger than the wheel base. Even though no detailed dy-
namic analysis was conducted, useful rules for towing and control purposes are derived from
the kinematics of towing, which builds on previous work by Fossum et al. [25]. These rules
can be used to determine what effect certain landing gear configurations might have when an
aircraft is being towed around specific airports. Detailed dynamic models are needed if the
effects of low friction surfaces (such as oil on concrete) are to be studied.
This was followed by a comparison between the tyre forces that are generated when the nose
landing gear follows a circular trajectory – equivalent to an exit radius – under its own power,
and when the aircraft is towed at an equivalent circular trajectory by a towbarless tug. We
assumed that a comparison of the steady-state forces would be sufficient, and that towbarless
towing would generate the same forces as a towtruck with a towbar. Continuation methods
provide such steady-state information quickly and efficiently. It was shown that the aft tyres of
the body landing gear experience significantly higher loads for the towing case due to a larger
lever arm with respect to the CG position, when compared to the shorter lever arms of the
engine positions. The tyres on the wing landing gear experience lower forces, when compared
to the self-propelled case. We can therefore conclude that towbarless towing could be a design
case for the aft axle of the body landing gear, which is offset by load alleviation on other parts
of the gear.
Chapter 5

Medium-Speed: Runway Exit Manoeuvres

5.1 Introduction

An exit manoeuvre consists of the transition from a straight-line to a circular motion, similar
to the towing studies in the previous chapter. The cornering characteristics of the aircraft
are determined by the gear layout, tyre properties and aerodynamics, and have been studied
in great detail for the A320. Previous cornering studies of the A320 showed that the inner
main gears lose grip above certain velocities and steering angles, which is characterised by a
family of Hopf bifurcations [13, 54, 55]. In this chapter we extend this analysis to the A380.
Different methods are compared for the prediction of steering angle and clearance distances,
and simulations are conducted to assess the lateral load factors on the tyres for the A320 and
A380.
Design engineers would like to obtain an early view of what clearances and steering angles
are needed to manoeuvre around specific airports. Very little design data is available at the
concept phase of a project, yet the design engineer needs to make a decision on what the
wheel base and track width of the aircraft should be. These parameters affect the steering
angles and clearances that will be obtained. A main assumption is that the nose landing gear
follows the taxiway centreline perfectly. This allows us to make a direct comparison of the
methods under consideration. Realistic ground manoeuvres reflect the typical scenarios that a
pilot encounters during normal day-to-day operations. Two methods are used when conducting
an exit manoeuvre. The judgemental oversteer (JOS) method is used where the pilot allows
the nose gear to overshoot the centreline, so that the geometric mean point of the main gears
follows the centreline1 . This method is used on smaller taxiways where clearance distances are
of concern, as is depicted in Figure 5.1(a). The cockpit over centreline (COC) method is where
the pilot steers the aircraft in such a way that the cockpit follows the centreline; it is depicted
in Figure 5.1(b). Symmetric thrust and brake inputs are used for these manoeuvres. Note that
the outer engines are typically set to idle-thrust on very large 4-engined aircraft (such as the
A380), for the avoidance of foreign object damage (FOD).
Figure 5.2(a) depicts relevant parameters for a runway exit, where an exit to the right is standard
convention. The position of the nose gear on the centreline is depicted by the polar coordinates
1
This method is actually an understeer method, but is called an oversteer method by the authors of the aircraft
characteristics manuals.

45
46 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

7 m (23 ft)
60 m
45.7 m (196.9 ft)
(150 ft)
COCKPIT TRACK
PARALLEL TO
GUIDELINE WLG PATH
NLG PATH
9.21 m 4.94 m (16.2 ft) R 25.5 m (83.7 ft) 10.24 m
R 45.7 m (150 ft) (30.2 ft) R 51 m (167.3 ft) (33.6 ft)
TAXIWAY
R 25.9 m (85 ft)
GUIDELINE TAXIWAY
GUIDELINE
WLG PATH

(a) (b)

FAA GROUP V AIRPORT FAA GROUP VI AIRPORT

JUDGEMENTAL OVERSTEER METHOD COCKPIT OVER CENTRELINE METHOD

22.9 m (75 ft)


30 m
(98.4 ft)

Figure 5.1. Extracts from the Airplane Characteristics Manual for the A380, showing the different
methods for exiting the runway, based on the runway category [2]. (a) JOS method, (b) COC method.

Rn and θn , and is measured in a clockwise direction from the negative x-axis. The heading
(or yaw) angle is denoted by ψ. Figure 5.2(b) shows a simulation where the trajectories of
specific points on the aircraft are traced out, of which the inner gear reference position is the
most important for clearance purposes. The dashed circle represents the steady-state radius of
the inner-gear reference position, which is for a scenario where the nose gear follows a circular
trajectory equivalent to the radius of the exit. The aircraft will transition from a straight line
trajectory onto a circular trajectory, eventually settling into a steady-state. It is important to
note that the turn centre is only located at the geometric centre of the circular arc when a
steady-state is reached. This is in contrast with References [11] and [17], where it is assumed
that the turn centre is based at the steady-state position, even during the transition phase of the
turn.
Section 5.2 shows how the steering angle can be calculated from kinematic methods, empiri-
cal formulae that are based on the results from kinematic methods, dynamic simulations, and
bifurcation methods. The different techniques are compared and show good agreement. A sim-
ple empirical formula could therefore provide valuable steering angle information when very
little data is available. Section 5.3 once again compares the results for clearance distances that
are provided from different methods, and shows that simple formulae can be used for such
predictions. These formulae are applicable to any aircraft configuration. Basic landing gear
configurations can be inserted into a novel clearance distance graph, which indicates whether
adequate clearance distances are available for the configuration under consideration.
Section 5.4 compares dynamic and steady-state loads for an A320 and an A380 at a typical
runway exit, and shows that the dynamic loads at the main gears are almost equal to the steady-
state loads that are obtained from continuation methods. The load factor at the CG (for both
aircraft) builds up towards a maximum value, without any overshoot. The maximum dynamic
tyre loads for the A320 are approximately 10% higher than the maximum steady-state values,
5.1. Introduction 47

ψ
δ

Rn
θm y
c
Rm θn
m x
Origin - O

(a)

Steady-state trajectory for


inner-gear reference position

(b)

Figure 5.2. Relevant parameters for a 90◦ exit. Panel (a) shows all the relevant dimensions, while Panel
(b) shows the trajectory of the reference positions.
48 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

while the steady-state tyre loads for most of the tyres of the A380 are in fact the maximum
values. The dynamic values are 10% higher than the steady-state values on the aft axles of the
BLGs. These findings are important for the next chapter, where the regulation for the high-
speed turn assumes steady-state conditions. We therefore propose that continuation methods
can be used to obtain the loads at runway exits, and more significantly, the loads that are gen-
erated during high-speed turns. The use of continuation methods for lateral loads studies has
previously been conducted on an aircraft with a traditional three post landing gear arrange-
ment [56]. No lateral loads studies, using continuation methods, have so far been conducted on
larger aircraft with more than three landing gears and multi-axle bogies.

5.2 Steering Inputs

Previous studies of the steering angle input variation during a turn have assumed a ramp func-
tion [56]. In contrast, the International Civil Aviation Organization (ICAO) has derived exact
solutions for the variation of the steering angle [31] based on elliptic integral equations, where
they show that the steering angle variation is closer to an exponential function. We suggest that
these equations are still too complex for everyday use by design engineers, and consequently
empirical formulae that are easy to use are derived from kinematic simulations. Steering an-
gles are also provided by the dynamic simulations as described in Section 2.3. The following
section compares the steering angle variation that can be obtained from kinematic simulations,
empirical estimates based on the kinematic simulations, bifurcation methods, as well as dy-
namic simulations. The steering angle variation consists of two distinct phases: the phase up
to the exit point of the turn, and the phase where the aircraft straightens out.

5.2.1 Steering Angle Variation During Circular Section of Exit

Equations (2.7) and (2.8) can be used to determine the trajectories by constraining the nose gear
to the runway and taxiway centreline. The velocity vector is tangent to the circular trajectory for
the duration of the turn. The analytical solution for the steering angle is given in Reference [31]
as
√ 2 !
1 − e Rn −1θn
δ = 2 tan−1 √ 2 √ 2 , (5.1)
Rn − Rn2 − 1 − Rn e Rn −1θn − Rn2 − 1e Rn −1θn
p p

where the radius of turn is normalised to the wheel base Rn = rlnn . The nose gear position in
the turn is denoted by θn and is measured in radians, where the nose gear position is θn = 0◦ at
the start of the turn, and θn = 90◦ at the exit of the turn. The final steady-state steering angle
is derived in [25] as
p !
Rn2 − 1
δf = cos−1 . (5.2)
Rn

A simpler formula that captures the steering angle variation δ during the turn, especially the
maximum steering angle δmax , is desirable, and consequently it was decided to obtain an em-
pirical formula based on the results from kinematic simulations. Figure 5.3(a) depicts the
5.2. Steering Inputs 49

40 4
35 δf 3.5
30
3
25
δ (deg)

20 2.5

k
15 2
k = 1.0530Rn − 0.3356
10
1.5
5
0 1
(a) (b)
−5 0.5
−30 0 30 60 90 120 150 180 1 1.5 2 2.5 3 3.5 4
θn (deg) Rn

Figure 5.3. Steering angle input for a 90◦ turn; (a) depicts the steering angle variation during a 90◦
turn obtained from a kinematic simulation, while (b) gives an approximation for the fitted exponent in
Equation (5.4). Data points extracted from the simulations are indicated with dots.

steering angle results for an exit manoeuvre that was obtained from simulating Equations (2.7)
and (2.8); it was concluded that the steering angle could be described by an equation of the
form
 
δ = δf 1 − e−kθn . (5.3)

Simulations were conducted for Rn ∈ [1, 4], from which the steering angle for all the different
combinations were extracted. Note that the normalised track width Lm is not important for
the steering angle predictions. The steering angle builds up in an exponential way, where the
exponent is dependent on the radius of turn rn and the wheel base ln . The radius of turn is
once again normalised to the wheel base. It is therefore possible to find the appropriate expo-
nent k by means of curve fitting methods. Figure 5.3(b) shows dots that represent individual
exponents that were obtained in this way. A linear approximation can then be used to obtain
the relationship between k and Rn and, therefore, the steering input required for a 90◦ runway
exit is found as
 
δ = δf 1 − e−(1.053Rn −0.336)θn . (5.4)

Note that θn denotes the position of the nose gear on the arc of the exit, and is defined in
radians; if θn is converted to degrees, the equation becomes
 
δ = δf 1 − e−(0.018Rn −0.006)θn . (5.5)

Equations (5.4) and (5.5) are applicable to any exit, and consequently the angle that the taxiway
makes with the runway can be inserted. The steering angle at the exit position would converge
50 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

50 Steady-state 50 Steady-state
135◦ Exit 135◦ Exit
40 90◦ Exit 40 90◦ Exit
δ (deg)

δ (deg)
30 30

20 20

10 10
(a) (b)
0 0
10 20 30 40 10 20 30 40
Wheel base ln (m) Wheel base ln (m)

Figure 5.4. Maximum steering angle inputs for, (a) Category V and, (b) Category VI airports.

towards the steady-state steering angle as the exit radius is increased, or as the wheel base is
decreased; the latter can be seen clearly in Figures 5.4(a) and 5.4(b). These graphs can be used
to estimate the steering angle requirements for any aircraft configuration, where Appendix C
contains steering angle calculations for different aircraft types. The position of the nose gear
on the circular trajectory can be converted to the time domain to yield

Vn
 
δ = δf 1 − e−(1.053Rn −0.336) Rn t . (5.6)

The velocity is normalised to the wheel base and is expressed as wheel base lengths per second.
It is assumed that the velocity during the turn stays constant, which is not a realistic scenario
for a constant thrust configuration, as the velocity will decrease as the turn progresses [56]. The
steady-state velocity for a specific steering angle can be obtained from a bifurcation diagram,
and consequently an average velocity can be derived for Equation (5.6). From simple geometry
in Figure 5.2, it can easily be shown that the heading angle ψ is the difference between the
angular position of the nose gear on the circular trajectory, θn , and the steering angle δ. At the
exit from the turn where θn = 90◦ , the heading angle becomes

ψ90 = 90◦ − δ90 (deg). (5.7)

5.2.2 Steering Angle Variation on Horizontal Section of Exit

The change in steering angle when the aircraft nose gear is moving on a horizontal line (x-
direction) forms a tractrix [31]; it can be derived from the y-coordinate of the main gear refer-
ence position cy , which is calculated as [25]

(k − 1) sin(ψ90 + 180◦ )
cy (t) = . (5.8)
ket − e−t
5.2. Steering Inputs 51

The next step is to apply the initial conditions at the exit, thus when θn = 90◦ . Set t = 0, and
insert the heading angle from Equation (5.7) at the exit ψ90 . The factor k is derived as

cos(ψ90 ) − 1
k= . (5.9)
cos(ψ90 ) + 1

The steering angle can then be calculated in the time domain as

δ(t) = sin−1 (cy (t)). (5.10)

Design engineers are more interested in the maximum steering angle and the steering rate,
hence, the need for a simplified equation in the previous section. No attempt will be made to
derive a simpler equation describing the decrease in steering angle from the exit, as this is not
a critical part of the manoeuvre.

5.2.3 Steering Angle Predictions from Continuation Methods

The previous empirical and kinematic methods described above are adequate for calculating the
towing angles during the preliminary and detail design phases. The steering angle predictions
are similar to the towing predictions during the preliminary design phase, but do need to be
refined for the detail design phase. Section 5.2.1 showed that a turn manoeuvre consists of a
transition between a straight line and a circular manoeuvre, and that the steady-state solution
of the steering angle can be used in Equation (5.4). This steady-state steering angle is however
based on geometric methods, which ignore the tyre and aircraft dynamics.
Bifurcation methods, and more specifically, continuation methods, can provide the steady-state
steering angle for different aircraft configurations, meaning that the tyre properties are taken
into consideration [15]. The question is whether continuation methods provide a more accurate
estimate of the steering profile compared to geometric methods. If so, the accuracy of the
steering angle prediction can be improved at a smaller computational cost, compared to full
dynamic simulations. Chapter 3 showed how a SimMechanics model of the A380 can be used
with the Dynamical Systems Toolbox for the prediction of turn widths pertaining to U-turn
manoeuvres. The same method is used here for the prediction of steering angles.
The turn radius can be calculated by dividing the velocity at the nose gear by the yaw rate; both
these states are provided by the continuation analysis. Figure 5.5(a) shows the variation of the
turn radius as the steering angle is increased for a velocity of 3 m/s. The steady-state steering
angle that will provide a turn radius of 51 metres for Group VI airport taxiway turns can be
obtained from the graph. In this case a steering angle of 37.9◦ will provide the required turn
radius.

5.2.4 Steering Angle Comparisons for the Different Methods

The pilot will usually try to maintain momentum (hence velocity) during the turn, by adjust-
ing the thrust. Future automated systems are also likely to maintain a specific velocity profile
during a turn, of which a constant velocity could be one of the candidates. A constant velocity
52 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

100 60
Dynamic
Kinematic
80 Empirical
45
Empirical+Bifurcation

60 δf

δ (deg)
rn (m)

(37.9,51)
a
30
40

15
20

(a) (b)
0 0
0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80
δ (deg) Time (s)

Figure 5.5. Steering angle comparisons; (a) steering angle obtained from continuation methods at a ve-
locity of 3 m/s; (b) comparison of steering angle variation obtained from kinematic, dynamic, empirical
and continuation (bifurcation) methods.

controller, along with a controller to follow the centreline, was incorporated into the detailed
dynamic model, allowing for a direct comparison of the steering angle for the different meth-
ods. Figure 5.5(b) contains a comparison of the input required from (i) a dynamic simulation,
(ii) a kinematic simulation, (iii) an empirical solution (Equations (5.6) and (5.10)), and (iv)
a combination of the steady-state steering angle obtained from bifurcation methods with the
empirical solution.
It can be observed from Figure 5.5(b) that there is very good agreement between the empirical
and the kinematic methods. An aircraft that is being propelled by its engines does however
need a side-force on the nose gear, which is generated by a slip-angle. Typical slip-angle
values for the nose gear during the turn would be between 2 and 3 degrees. The steering angle
that is needed for the self-propelled case can either be provided by the bifurcation method or
by simulations. We adapt the empirical method to include the self-propelled case, by using the
steady-state steering angle that is provided from continuation in Section 5.2.3, where we see
that the steering angle predictions from the continuation method are close to the results from
the simulations.
The empirical method is also an adequate tool for the prediction of towing angles in Chapter 4,
due to the fact that the constraint forces are generated by the tug, and would be sufficient during
the preliminary design phase. Bifurcation methods can be used in conjunction with empirical
methods for the prediction of steering angles of an aircraft that is self-propelled, where a slip-
angle on the nose gear is needed. Bifurcation methods would be adequate during the detailed
design phase of a project.
5.3. Clearance Distances 53

5.3 Clearance Distances

Runway design rules published by the Federal Aviation Administration (FAA) use the steady-
state trajectory to design the fillet sizes for a specific aircraft [23], while ICAO shows that the
trajectory during the turn can be calculated from an elliptic integral [31], which can be evalu-
ated with great effort. An equation for the deviation of the main gear centre position from the
circular arc, which is dependent on the steering angle, is also derived in [31]. This is once again
a difficult equation for design engineers to use, and it does not include the clearance distance of
the inner gear reference position. It would therefore be useful to provide design rules on how
the wheel base and track width influence the performance characteristics. Figures 5.2(a) and
(b) show the dimensions that are relevant to exit manoeuvres, where the dashed circle in Fig-
ure 5.2(b) depicts the steady-state trajectory of the inner gear reference position. The distance
between this circle and the runway edge provides the clearance distance, representing a worst
case scenario; this is equivalent to the method used for predicting clearance distances by the
FAA [22]. Figure 5.2(b) also shows that the aircraft does not reach a steady-state, because the
trajectory of the inner-gear reference position does not converge onto the dashed line. More
accurate estimates are needed to allow the design engineer more leeway with regard to the
positioning of the gears.
Simulations of the kinematic equations are used here to obtain minimum clearance distances
for varying wheel base and track width combinations. The radius of turn and the track width are
normalised to the wheel base, where Rn ∈ [1, 4] and Lm ∈ [0, 0.6]. Here the only restriction is
that the turn radius should not be smaller than the wheel base, which is a reasonable assump-
tion based on the aircraft and airport data that are available. The distance of the inner-gear
reference position m from the centre of the turn is denoted by Rm , and it is monitored during
the simulation. The angular position of point M relative to the negative x-axis is denoted by
θm . The polar coordinates of the minimum clearance position can then be extracted from the
calculations, and are denoted by (Rm90min , θm90min ). The dots in Figure 5.6 are data points
that are obtained from the kinematic simulations; they indicate the angular positions for a 90◦ -
and a 135◦ -exit. The solid curves are fitted to the data points, where the angle of the inner gear
reference position, at minimum clearance, for a 90◦ -exit is represented by

θm90min = −0.602Rn 2 + 7.378Rn + 56.526 (deg), (5.11)

while the angular position of the inner gear reference position, at minimum clearance, for a
135◦ -exit is given by

θm135min = −1.580Rn 2 + 14.964Rn + 85.874 (deg). (5.12)

The next step is to obtain an expression for the minimum clearance distance, for the inner
gear reference position (normalised to the wheel base) for a 90◦ -exit. An expression for the
clearance distance can be obtained in a similar way to the method that was used for the steering
angles. A surface (not shown) can be fitted through the data, which can be described as

Rm90min = −0.024Rn 2 + 1.203Rn − 0.5Lm − 0.553. (5.13)


54 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

130

120

110
135◦ Exit
θm (deg)

100

90

80

70
90◦ Exit
60
1 1.5 2 2.5 3 3.5 4
Rn

Figure 5.6. Angular position where minimum radius occurs for inner gear reference position; Dots
indicate data points obtained from simulations, and the solid curve is fitted.

Table 5.1. Minimum clearance location of inner gear reference position when the COC method is used
for a 90◦ -exit, as depicted in Figure 5.7.

Number X- Y- Distance Distance Distance


coordinate coordinate from origin O rel. to (i) Error
(m) (m) (m) (m) (%)
Dynamic (i) −13.76 33.08 35.83 0.00 0.00
Kinematic (ii) −13.62 32.74 35.46 0.37 1.04
Empirical (iii) −13.67 32.61 35.36 0.48 1.34
Steady-state (iv) −20.10 27.29 33.89 8.58 23.96

An analysis of the results for a 135◦ -exit indicates that the minimum clearance distance for this
turn can be described as

Rm135min = −0.043Rn 2 + 1.323Rn − 0.5Lm − 0.742. (5.14)

Note that the results are applicable to any aircraft; the A380 is used here as a case study to verify
the results. Table 5.1 and Figure 5.7 contain a comparison of the location of the minimum radial
clearance of the inner gear reference position for a 90◦ -exit. They compare the values obtained
from (i) a full dynamic simulation, (ii) a kinematic simulation, (iii) the empirical method that
was derived in Equations (5.11) and (5.13), and (iv) the steady-state position that is used for
airport planning purposes by the FAA [23]. A simple controller was designed for the dynamic
simulation, where any point can be set to follow the centreline, with the nose gear defined as
the reference point for this specific study. The pilot is assumed to act as a closed loop controller
with the aim of keeping the area around the cockpit above the centreline. The errors from the
kinematic and empirical methods are calculated as
5.4. Load Factors During Runway Exit Turns 55

Mimin − Mjmin
e=1− . (5.15)
|Mimin |

Following the convention in Figure 2.3, where M is the position vector of the inner-gear refer-
ence position, Mimin denotes the minimum radius vector position for the dynamic simulation
case. The other positions are denoted by Mjmin where j represents the label of the case under
consideration. The results from the empirical predictions are within 1.5% of the results from
the dynamic simulations. It can be concluded from Figure 5.7 that Equation (5.13) gives a good
estimate of the clearance distance that can be achieved for a 90◦ -exit; in particular, the estimate
is significantly better than the steady-state approximation. The same holds true for the 135◦
case, although the minimum radial distance is close to the steady-state value.
Feasible wheel base and track width combinations can be determined by inserting a minimum
radial distance into Equations (5.13) and (5.14), and then recasting the equations so that the
track width appears on the left-hand side. The fillet radius for Group V and VI airports is as-
sumed to be 25.9 metres [2], where a minimum clearance distance of 4.5 metres is needed [31].
The locus of track widths that provide a radius of 30.4 metres will therefore provide the bound-
ary for the feasible region of wheel base and track widths. Appendix C contains clearance
distances for different aircraft types.
Figure 5.8 depicts such an envelope for the different types of airports and exits. The area to the
right of a hatched line indicates a clearance distance of less than 4.5 metres for the particular
case. The wheel base that is depicted on the x-axis could also be replaced by the distance from
the cockpit to the main gear reference position, which would represent the COC method. We
have assumed that the nose gear will follow the centreline. It is possible to overlay the data
for any aircraft onto the graph, making the method useful for comparative purposes. We have
overlaid the data (given in Appendix C), and it can be seen that the COC minimum clearance
distance for the A380 lies to the right of the required margins for Group V airports, hence the
JOS method needs to be employed for this type of airport; this is consistent with the Airplane
Characteristics Manual for the A380 [2], contained in Figure 5.1. Note the data point for the
B747-8, for which it is predicted that the aircraft will have sufficient clearance on a Group V
airport with a 90◦ exit, but not for a 135◦ exit. It can be concluded that Figure 5.8 can be used
as a useful design tool at the early stages of an aircraft programme. On the other hand, the
final clearance distances that are published in the Aircraft Characteristics Manual should still
be obtained from detailed dynamic simulations.

5.4 Load Factors During Runway Exit Turns

The following section explores how the International Civil Aviation Organization (ICAO) de-
termines the maximum exit velocity at a runway exit. Dynamic simulations of the A320 and
A380 are then conducted at a typical runway exit in the sections thereafter. The results are also
compared to steady-state values that are obtained from continuation analysis. These results are
of importance for the assumptions that are made in Section 4.3 related to towing, as well as
the following chapter, which studies the lateral loads that can be obtained during high-speed
turns. The regulatory cases for lateral loading do not specify any dynamic behaviour during
a turn. Rather, they only specify that the aircraft needs to be configured in such a way that a
56 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

34 i
32 ii,iii
Steady-state radius
Inner-gear reference 30
position 28
O iv
26
-21 -19 -17 -15 -13

Figure 5.7. Minimum clearance location of inner gear reference position for a 90◦ exit, COC method
for different calculation techniques: (i) dynamic , (ii) kinematic , (iii) empirical, (iv) steady-state, based
on the FAA method.

25

20
G

Out of bounds
rp
V
Gr

I9
pV
Track width lm (m)

15
I1

A380
35
°

C5 B747-8 B777-300ER
B787
A330-200 MD-11 A340-600
10
An-124
A320 A321
B737-900
Gr
pV

5 MD-81
Gr

90
°
pV
13

0
10 15 20 25 30 35 40
Wheel base ln (m)

Figure 5.8. Envelopes for Group V and VI runway exits with data points for selected aircraft, where the
COC method is employed.
5.4. Load Factors During Runway Exit Turns 57

Table 5.2. Runway exit velocities obtained from the ICAO Aerodrome Design Manual [31].

Airplane Design Group Exit Type Radius (m) Vn ( m/s)


A 90◦ /135◦ 22.5 5.4
B 22.5 5.4
C 30.0 6.3
D 45.0 7.7
E 45.0 7.7
F 51.0 8.2
1,2 High-speed 275.0 18.9
3,4 500.0 25.5

lateral side load factor of 0.5g is maintained. It is therefore implied that the dynamic loads do
not exceed the steady-state loads. We aim to support this assumption in this section.

5.4.1 Runway Exit Design Velocities

The accelerations that are generated during a turn are highly dependent on the entry velocity
into the turn. The International Civil Aviation Organization (ICAO) uses a steady-state lateral
load factor (defined as the ratio of the lateral load at the CG to the weight of the aircraft)
of 0.133 to determine the maximum design velocities that can be used at runway exits [31].
These velocities are obtained by using the formula for centripetal acceleration, inserting the
load factor of 0.133, and then rewriting the formula in the form


Vn = 1.1422 rn , (5.16)

where rn represents the radius of the turn. Table 5.2 provides the design velocities that are
obtained when this formula is used. The author has not been able to ascertain why the specific
value of 0.133 was chosen for the load factor, but it does provide airport designers with a means
of specifying runway exit velocities. The design rules stipulate that the operational velocities
shall be below these values, and that they need to be determined empirically [31]. Each airline
therefore stipulates its own rules with regards to exit velocities, which shall always be below
the design values. Maximum operational velocities are typically 60% of the design velocity to
ensure that all aircraft will exit runways in a safe operating region [31].
Sections 5.2 and 5.3 showed how kinematic methods can be used to obtain estimates for the
steering angle evolution, as well as minimum clearance distances, for a generic runway exit
manoeuvre. The results showed that the evolution of the steering angle during a runway exit
manoeuvre resembles an exponential function, which may suggest that the forces build up
towards a steady-state value without any overshoot. The baseline dynamic models are used
to determine the types of loads that are generated during a turn. We test this hypothesis by
considering exit manoeuvres for the A320 and A380 aircraft. A 90◦ turn to the right at a Group
VI airport is used, while a realistic operational velocity of 4 m/s (8 knots) is imposed at the nose
landing gear position, which is approximately 50% of the exit velocity prescribed in Table 5.2.
The radius of the turn is 51m.
58 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

5.4.2 Load Factors for an A320

Figure 5.9 contains the steering angle, yaw rate, and lateral load factors, at different points of
the aircraft, for an A320 that is conducting the prescribed turn. This is for the MRW and aft
CG-position. The panels on the left in Figure 5.9 use time as the independent variable, while
the panels on the right use the steering angle. Figure 5.9(a) therefore shows the evolution of
the steering angle over time. Note that dynamic data from simulations are represented by solid
curves, while steady-state values from continuation runs are represented by dashed curves. The
dashed curve in Figure 5.9(a) represents the steering angle that is needed to maintain a radius
of 51m at the nose gear, for the required velocity; it is obtained from continuation runs. Similar
to those in Chapter 3; graphs with time as the independent variable show only the final steady-
state value of steering angle that produces a turn radius of 51 m, while the dashed curves in
the graphs that contain the steering angle as the independent variable represent the steady-state
values for a set of steering angles. (This is equivalent to obtaining the graph by inserting a very
slow ramp input, and then recording the state values at each steering angle.) It can be seen
that the steering angle reaches the steady-state steering angle during the turn, maintains this
condition for a while, and then reduces to zero when the exit position from the turn is reached.
Figure 5.9(b) shows the yaw rate at the CG, and it illustrates that the yaw rate never exceeds the
final steady-state value. Figures 5.9(c1) and (c2) depict the lateral load factor at the CG, used
in the ground loads regulations in the following chapter. The graph shows that the lateral load
factor converges to the final steady-state value without any overshoot. This result is unexpected,
as there might be an overshoot that is larger than the final steady-state value. Overshoot usually
occurs when step inputs are provided to the system, which is not the case in this analysis where
an exponential input is used. Ramp inputs are also often used in simulations, where the nose
gear is allowed to deviate from the centreline within certain margins [56]. The current approach
of an exponential function that was developed in Section 5.2 for the steering angle is deemed
to be the most realistic, as the pilot would act as a controller that maintains the nose gear close
to the centreline.
Figures 5.9(d1) and (d2) show that the nose gear initially has to provide enough force to decel-
erate the aircraft along the fuselage axis, while accelerating the aircraft around the yaw-axis.
The dashed curves once again represent the steady-state values when the nose gear follows a
radius of 51m. Both components of the force are provided by the nose gear tyres. The aircraft
will not be able to conduct the turn at the required radius if these forces cannot be provided.
The force consequently builds up rapidly at the onset of the turn, and then reduces to the
steady-state value. A significant overshoot is therefore present.
Figures 5.9(e1) and (e2) contain the lateral load factors for the outermost main gear tyre, W1,
and innermost main gear tyre, W4. The main gear tyres initially resist the rotational motion at
the onset of the turn, which can be seen from the negative values for the load factor. The forces
then change direction and orientate to the general direction of the turn centre. In this case the
forces build up towards the steady-state values. Note that the outer gear load factor is less than
the inner gear load factor, even though the outer gear experiences higher vertical loads. This
is due to a larger slip angle on the inner wheel (W4) when compared to the outer wheel (W1).
An opposing force is also present when the aircraft tries to straighten out at the exit point of
the turn.
The hysteresis in the graphs is indicative of nonlinearity in the system, which is mainly due to
the tyre characteristics in these cases. The overshoot value for the main gear tyres do not exceed
5.4. Load Factors During Runway Exit Turns 59

20 6

ωz at CG (deg/s)
(a) (b)
15
δ (deg) 4
10
2
5
0 0
0 10 20 30 40 50 0 4 8 12 16

0.04 0.04
(c1) (c2)
0.03 0.03
ny at CG

0.02 0.02
0.01 0.01
0 0
0 10 20 30 40 50 0 4 8 12 16

0.2 0.2
(d1) N1 (d2)
ny at NLG

0.1 N1 0.1 N2
N2
0 0
-0.1 -0.1
-0.2 -0.2
0 10 20 30 40 50 0 4 8 12 16

0.05 0.05
(e1) W4 (e2)
0.04 W4 0.04
ny at MLG

0.03 0.03
0.02 W1 0.02 W1
0.01 0.01
0 0
-0.01 -0.01
0 10 20 30 40 50 0 4 8 12 16
Time (s) δ (deg)

Figure 5.9. Evolution of the lateral forces of an A320 aircraft conducting a turn at a Group VI runway
exit, while maintaining a nose gear velocity of 4 m/s. Panel (a) shows the evolution of the steering angle
δ over time, where the dashed line indicates the steady-state steering angle that is obtained from the
bifurcation method. Panel (b) shows the yaw rate ωz around the CG as the steering angle is varied. The
dashed line again depicts the steady-state values obtained from the bifurcation method. Panels (c1) and
(c2) show the lateral load factor ny at the CG-position. Panels (d1) and (d2) show the load factors for
the outer (N1) and inner (N2) nose gear tyres. Panels (e1) and (e2) show the load factors for W1 and
W4 .

10% of the steady-state value. It is possible to obtain the steady-state values directly from
bifurcation diagrams, and then we can assume that the dynamic value will be approximately
10% larger than this value. This dynamic effect is ignored in the regulation for high-speed turns
in the following chapter, but we will include this effect for the analysis of the A320. Figure 5.10
shows the maximum load factors that occur in the main gear tyres, associated with the peak
values of the opposing force as the aircraft straightens out. Shown are the maximum dynamic,
60 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

0.05
Maximum steady-state force
0.04 Maximum dynamic force

0.03
ny

0.02

0.01

0
W1 W2 W3 W4
Wheel Number

Figure 5.10. Dynamic and steady-state load factors on the main gear tyres for the A320 and Vn = 4 m/s.
The nose gear maintains a radius of 51m.

and steady-state values when the steering angle is maintained to follow a constant radius of 51
m. The steady-state values are not higher than 10% of the maximum dynamic values.

5.4.3 Load Factors for an A380

The next step is to look at an aircraft with more than two main gears. An A380 model at MRW
and aft CG position is used for this purpose. A similar velocity and exit radius is used to that
of the case for the A320, with a turn direction to the right. Figure 5.11 contains the steering
angle, yaw rate, and lateral load factors at different points of the aircraft. In Figure 5.11(a) it
can be seen that the steady-state steering angle is not reached for the case of a 90◦ exit, due
to the larger wheel base; this is consistent with the steering angle results from Appendix C,
where a steering angle of 32.72◦ is predicted, at the exit position, for the A380. This is in good
agreement with Figure 5.11(a). It is interesting to note that the yaw-rate from the simulations
in Figure 5.11(b) (solid line) is very close to the steady-state values that are obtained from
continuation analysis (dashed line). The yaw rate at the CG-position does not reach the final
steady-state value, and consequently it can be assumed that the steady-state value could be used
as a maximum value for design purposes. Figures 5.11(c1) and (c2) show that the lateral load
factor at the CG builds up towards a value that is close to the steady-state value, and contains
no overshoot. The simulation results are once again close to the continuation results.
The largest difference in load factor occurs between the inner-most, and outer-most tyres, hence
these tyres are used for the A380 comparisons that follow. The comparisons always progress
from the nose gear tyres to the first row of the wing gear, then to the middle row of the body
gear, and finally, the aft row of the body gear. The load factors in the other rows are significantly
lower and are consequently not shown in the detailed comparisons. Figures 5.11(d1) and (d2)
show that a large hysteresis loop exists in the nose gear tyres, N1 and N2, indicating nonlinear
effects. Even though the dynamic values from simulations are significantly higher than the
5.4. Load Factors During Runway Exit Turns 61

50 6

ωz at CG (deg/s)
(a) (b)
40
δ (deg)
30 4
20 2
10
0 0
0 10 20 30 40 50 0 10 20 30 40

0.04 0.04
(c1) (c2)
0.03 0.03
ny at CG

0.02 0.02
0.01 0.01
0 0
0 10 20 30 40 50 0 10 20 30 40

0.4 0.4
(d1) (d2)
0.3 N1 0.3 N1
ny at NLG

0.2 N2 0.2 N2
0.1 0.1
0 0
0 10 20 30 40 50 0 10 20 30 40

0 0
(e1) (e2)
-0.1 -0.1
ny at MLG

W1 W1
-0.2 -0.2
W4 W4
-0.3 -0.3
-0.4 -0.4
0 10 20 30 40 50 0 10 20 30 40

0.4 0.4
(f1) (f2)
0.3 0.3
ny at BLG

W16 W16
0.2 0.2
0.1 W13 0.1 W13
0 0
0 10 20 30 40 50 0 10 20 30 40

0.4 0.4
ny Aft Axle BLG

(g1) (g2)
0.3 0:3
0.2 0.2 W17
W17
0.1 0.1
W20 W20
0 0
0 10 20 30 40 50 0 10 20 30 40
2
Time (s) δ (deg)

Figure 5.11. Evolution of the lateral forces of an A380 aircraft conducting a turn at a Group VI runway
exit, where Vn = 4 m/s. The representation of the data in the different panels is as in Figure 5.9.
62 Chapter 5. Medium-Speed: Runway Exit Manoeuvres

steady-state values, they still do not exceed the steady-state values when a turn radius of 51m
is maintained.
Figures 5.11(e1) and (e2) depict the load factors for tyres W1 and W4. Figure 5.11(e1) shows
that the final steady-state is not reached, while Figure 5.11(e2) shows that the dynamic values
are close to the steady-state values during the turn. Figures 5.11(f1) and (f2) can be interpreted
in a similar manner for W13 and W16. The tyres on the aft axle of the BLG, tyres W17 and
W20, are shown in Figure 5.11(g1) and (g2), indicating similar load factors in both tyres. The
dynamic loads are also larger than the final steady-state value if a radius of 51m is maintained.
The kink in the curves indicate the point where the body gear steering switches on, indicating
that significant load relief can be obtained by adding steering onto the aft axles of the BLGs.
Figure 5.11(g2) shows that the dynamic values are not far from the steady-state values.
Figure 5.12 contains the steady-state load factors on all 20 of the main gear tyres when the nose
gear maintains a radius of 51m, and also the maximum dynamic load factor values for a 90◦
exit. The steady-state values are more critical for most of the tyres, apart from the tyres on the
aft axles of the BLGs, where the maximum dynamic values are approximately 10% higher for
the aft axle wheels, when compared to the steady-state values. The overall steady-state loads
on the BLGs are larger than the overall dynamic gear loads. The maximum tyre loads occur
at the exit point from the turn (θ90 ) for W1-W16. The maximum for W17-W20 occurs after
θ90 when the aircraft straightens out. A similar pattern to that of the A320 emerges, where the
inner tyres have a larger load factor compared to the outer tyres, even though the outer gears
experience larger vertical loads.
We can therefore conclude that the dynamic load factors on the tyres of the A320 will be
approximately 10% higher than the steady-state values. This is in contrast to the A380, where
the steady-state values are more critical for most of the tyres, which is consistent with the
interpretation of the lateral loads regulation in the following chapter. The previous chapter
showed that towing loads cause significantly higher loads in the aft tyres of the BLGs, hence
the dynamic effect on the aft axles of the BLGs are likely to be less important than for the
towing case. One could therefore use the steady-state loads on an aircraft such as the A380 for
design purposes, and add a correction to incorporate the dynamic effect for a smaller aircraft
such as the A320. Continuation analysis provides these steady-state values.

5.5 Discussion

An empirical formula that was derived from the results from kinematic simulations, was used
to evaluate the steering angle variation during a 90◦ -exit manoeuvre. This empirical formula
showed very good agreement with kinematic and dynamic simulations. A steering input was
derived from a dynamic model with a controller for additional validation purposes, where the
controller is used to maintain the nose landing gear on the taxiway centre line. The steering
angle obtained with the controller showed small differences from the steering angle obtained
from the kinematic method, due to a slip-angle that is needed when the dynamic method is
used. This slip-angle is responsible for the generation of a side force that maintains the cir-
cular trajectory of the nose landing gear. More accurate variations of the empirical formula
can be obtained, by adjusting the final steering angle. This adjustment can be obtained from
continuation; this is in close agreement with the values that are obtained from simulations. The
5.5. Discussion 63

0.3

0.2

0.1
ny

−0.1

−0.2 Maximum steady-state force


Maximum dynamic force
−0.3
W1 W2 W3 W4 W5 W6 W7 W8 W9 W10 W11 W12 W13 W14 W15 W16 W17 W18 W19 W20
Wheel Number

Figure 5.12. Dynamic and steady-state load factors on main gear tyres for the A380, when a radius of
51 m at the nose gear is maintained, for a nose gear velocity of 4 m/s.

advantage of the continuation approach, compared to simulations, is that it is more efficient in


terms of run-times and can be used to generate steady-state information.
We also used the kinematic equations for the prediction of clearance distances for runway exits.
The equations are normalised to the wheel base and are therefore valid for any track width and
wheel base combination. The only restriction is that the turn radius should not be smaller than
the wheel base, which is a reasonable assumption based on the aircraft and airport data that
are available. A comparison of clearance distances was made between a validated dynamic
model, a kinematic model, an empirical model that was derived from the kinematic model,
and a geometric method used by the FAA. It was shown that the clearance distances for the
kinematic and empirical methods are within 1.5% of the predicted clearance distance that is
obtained from the dynamic model. The empirical model can be used as a useful tool by design
engineers for the prediction of clearance distances for any type of exit or aircraft configuration.
This was followed by a comparison of the load factors that can be obtained at runway exits,
where the load factor is defined as the ratio between the lateral and vertical loads. We show
how ICAO calculates the maximum runway exit design velocities based on a maximum lateral
load factor criterion of 0.133 at the CG position. Simulations were conducted at a runway exit
for a typical operational exit velocity, and they show that the dynamic load factors at the CG
positions, for both the A320 and the A380, build up towards a steady-state value. The dynamic
load factors are 10% higher than the steady-state values, at the main gears for the A320, while
the maximum values are, in fact, the steady-state values for W1-W16, for the A380. The
dynamic load factors on the aft-axles, W17-W20, were 10% higher than the steady-state values.
Chapter 4 showed that the towing case is likely to drive the design of the aft axle components,
which negates the larger dynamic loads for the self-propelled case. Continuation methods can
therefore be used to analyse the load factors that can be obtained at the CG position for the
A320 and the A380, as well as the main gears for a large aircraft such as the A380. Overall,
the dynamic load factors at the main gears of the A320 can be obtained by multiplying the
steady-state values by a factor of 1.1.
Chapter 6
High-Speed: Ground Loads Requirements

6.1 Introduction

The current lateral ground loads requirement, FAR25.495, for an aircraft during a high-speed
turn, was written in the middle of the last century, when relatively small aircraft with tricycle
landing gear arrangements started to emerge. This requirement is known to be conservative
when applied to large modern passenger aircraft. In this chapter we assess the loads that can
be generated for an A320 and an A380 during typical operational ground manoeuvres, and
compare the results to the original requirement. We show that static balance calculations and
continuation methods can be used to assess the loads that are generated. Comparisons are
made between the two aircraft types, which show significantly different dynamics in terms of
stability and loads.
Nonlinearities (such as tyre forces) have a more significant effect at the edge of operating en-
velopes, placing a renewed interest on analysis methods that can classify the dynamics in these
regions. Nonlinear effects also ensure that aircraft do not experience the high lateral loads that
are stipulated in the lateral ground loads requirements by the FAA [13]. Section 6.2 discusses
the limitations related to the requirement. It also discusses the main findings from an opera-
tional ground loads measurement campaign of in-service aircraft [69], which was specifically
conducted to compare operational loads with the requirement. This study confirmed that an in-
crease in aircraft size is accompanied by a reduction in the maximum lateral load. A B747-400
experiences smaller lateral load factors compared to, say, an A320. The authors of this research
have also published some empirical formulae that were derived from the measured data, to help
with predictions for aircraft that were not part of the study [69]. These formulae are used here
to make predictions of the maximum lateral loads that an aircraft such as the A380 is likely
to experience in its lifetime. Section 6.3 shows how aircraft manufacturers interpret the high-
speed lateral loads requirement. Static balance equations are used to calculate accurate landing
gear loads for an aircraft with three landing gears, and it is compared with the results from a
simulation. This method can, however, not be employed for aircraft with more than two main
gears.
Section 6.4 contains numerical continuation results for the high-speed turn, as obtained from
the detailed A320 and the A380 models. The analysis is conducted in accordance with the
regulations at the maximum ramp weight (MRW) condition, with fore and aft CG positions.
The stability results in Section 6.4.1 for the A320 compare well with the results from previous

65
66 Chapter 6. High-Speed: Ground Loads Requirements

studies [13, 54, 55]. A clear boundary is formed by Hopf bifurcations, indicating a loss of grip
at the inner main gear tyres [13, 54]. These areas of instability occur at relatively low speeds,
hence the aerodynamics does not play any significant role as far as the A320 is concerned.
The results in Section 6.4.2 for the A380 are significantly different to those of the A320: no
bifurcations were detected, and the aircraft is in fact remarkably stable. Problematic areas are
however identified in terms of manoeuvring at moderate velocities. The analysis shows that
when a runway exit turn is conducted, the nose gear tyres cannot generate enough side force
above 8 m/s. The maximum prescribed velocity of 4 m/s avoids this problem. We show that
the gear loads can be classified across the entire operating envelope, where the WLGs act in the
opposite direction to the BLGs at low velocities, contrary to the assumption in the regulation.
Section 6.4.3 shows how continuation results can be used to assess the lateral load factors on
individual tyres. Only the inner tyres in a turn experience load factors that are in the vicinity
of the values stipulated by the requirement. The lateral load factors on the outer tyres are
significantly less than the requirement, confirming its conservative nature.
Section 6.5 compares the maximum load values that were obtained from the operational loads
measurement campaign with predicted results from numerical continuation analysis. A max-
imum load factor envelope was constructed from the continuation results. All the significant
data points were located within this envelope, showing that continuation methods provide a
conservative estimate of the maximum lateral load factors. However, it is still less than the
value prescribed in the regulation. This section also explains how an A320 could generate sig-
nificant lateral loads. The velocity where this extreme value occurs is then used as an extreme
case for the A380. The lateral load factor provided by continuation methods is approximately
10% larger than predicted from the operational study. Continuation methods do however pro-
vide complete coverage of the entire operating envelope. For the first time, it is now possible
to pinpoint the exact steering angle, and velocity, where the maximum load factor will occur.

6.2 Side Loads Requirements

One of the main design cases for aircraft ground loads pertains to the high-speed turn, which
tends to be the critical case for the design of the attachments of the main landing gears. Federal
Aviation Regulation (FAR) 25.495 [21] deals with this load case, and is phrased as follows:

“In the static position, in accordance with figure 7 [Figure 6.1(a)] of Appendix
A, the aeroplane is assumed to execute a steady turn by nose gear steering, or by
application of sufficient differential power, so that the limit load factors applied at
the centre of gravity are 1.0 vertically and 0.5 laterally. The side ground reaction
of each wheel must be 0.5 of the vertical reaction”.

Figure 6.1(a) depicts this requirement. The type of aircraft in the picture shows that the ori-
gins of the requirement are probably from the 1940’s or 1950’s. This load case forms one of
the bookcases for ground loads as stated in the regulations [21], and needs to be considered
by aircraft manufacturers in the design of their aircraft. All of these cases consist of static
external forces that usually require ground reactions to be balanced by applying inertia forces
and moments. We recap, by noting that the lateral load factor ny is defined as the ratio of the
6.2. Side Loads Requirements 67

THE AIRPLANE INERTIA FACTORS AT


CENTER OF GRAVITY ARE COMPLETELY
BALANCED BY THE WHEEL REACTIONS
AS SHOWN

0.5W W

0.5VM2 0.5V 0.5VM1


N
V V
M2 M1

NOSE WHEEL TYPE

(a) Original image depicting the lateral load factor requirement;


from [21].

10 3
Taxi out Corrected
Take off roll corrected
Landing roll corrected
2
10 Turn off corrected
Taxi in corrected
Combined corrected
1
10 A320
Cumulative Occuurences per Flight

33541.9 Flight hours


10066 Flights

10 0

10 -1

-2
10

10 -3

10 -4

10 -5

-6
10
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

Lateral Load Factor Ny, (g)

(b) Side loads as measured on an A320 aircraft; reproduced from [69].

Figure 6.1. FAR 25.495 side load requirements for the high-speed turn.
68 Chapter 6. High-Speed: Ground Loads Requirements

lateral force over the vertical force, which is 0.5 for this specific regulation. Reference [69]
states that little is known about the origin of this regulation, or even under what operational
conditions such a loading condition might occur. The regulation was clearly brought in when
aircraft were much smaller than they are nowadays, and it may be that aircraft with more than
two main gears did not exist at the point in history when the regulation was drafted; however,
we have no evidence for this assertion. Reference [69] also states that landing gear and aircraft
manufacturers believe that the current 0.5g lateral load factor requirement specified by FAR
25.495 is too stringent. The regulatory authorities are however extremely reluctant to modify
this requirement in the absence of data to back such a decision. The requirement has clearly
ensured safe aircraft operations, and consequently overwhelming evidence would be needed
for its relaxation.

6.2.1 Limitations of the Regulation

The loads produced by the high-speed turn form one of the limit load cases and, as such, con-
tribute towards the aircraft and landing gear limit load envelopes. The loads can be calculated
accurately for aircraft with statically determinate (tricycle) gear arrangements such as the A320,
where the fuselage and gears are assumed to be rigid [54, 55]. However, statically indetermi-
nate (more than three gears) arrangements such as the A380, pose computational challenges,
where minimum energy techniques are needed to calculate the response of the aircraft [15].
This was discussed in Section 2.3. Such an approach is needed to protect any one gear against
disproportionate loading due to the stiffness characteristics of the gears and fuselage [74]. For
this reason aircraft manufacturers tend to include other cases (in addition to the bookcases)
into their analysis. These cases are based on years of experience that try to cover day-to-day
operational scenarios. These additional cases are known as rational cases, as they utilise mod-
els that contain a more accurate representation of the real physics and dynamics of the system.
In particular, the dynamic response of the airplane is included. Structural inertia, flexibility,
and damping are accounted for, as well as distributed aerodynamic lift and moments. This is a
more realistic method of investigating the actual forces on the aircraft.
Dynamic calculations are employed for rational cases such as taxiing or landing, and may form
part of the limit load envelope. It is also normal practice to use such dynamic calculations to
establish the aircraft fatigue loads. Nonlinear effects within the tyres and landing gears make
it difficult to find the exact values for the steering angles and velocities where the maximum
lateral loading may occur. This is the reason for adhering to the bookcase approach, with the
addition of rational cases. Nonlinear effects could for instance lead to multiple steady-states for
the same thrust and steering angle settings, when considering the high-speed turn, such that the
response depends on the initial conditions and size of the perturbations on the system [13, 54].
A large perturbation could for instance cause a “jump” to another steady-state. This type of
behaviour is the motivation to use methods from the field of Dynamical Systems theory, to
analyse the dynamics of an aircraft on the ground [13, 15, 54, 55, 56].

6.2.2 FAA Operational Loads Study

In answer to the points mentioned in the previous section and in Reference [69], the FAA con-
ducted a large operational loads testing campaign, where it aimed to identify the maximum
6.3. Aircraft Loads from the Static Balance Equations 69

operational loads that can be experienced by in-service aircraft. A critical assessment of the
loads was then made against the current regulations [69]. The findings from this study showed
that it was very difficult to obtain lateral accelerations above 0.3g for any aircraft types. As an
example, Figure 6.1(b) shows the measured side loads for the A320 that was used in [69]; it
shows that they did not exceed 0.25g for the approximately 10,000 flights that were recorded.
The authors of [69] collated the information for all the assessed aircraft and derived an empiri-
cal formula to predict the likelihood of obtaining a certain g-level in 1000 flights as

s 2
N = N0 e−ln lm ny . (6.1)

Here N is the number of cumulative occurrences of any lateral load factor ny in 1000 flights,
where ny is the maximum lateral load factor measured during a turn; ln represents the wheel
base and lm the track width. N0 is the number of cumulative occurrences at ny = 0; s rep-
resents a shape parameter specific to the aircraft being studied. Each aircraft therefore has
its own factor for N0 and s, that are derived from the study. The authors of [69] average the
factors of all the aircraft to obtain a generic formula that can be used for any aircraft. Their
initial assessments show that the taxi-in phase seems to generate larger load factors, but they
point out that corrections need to be made to the taxi-in data, due to the fuel that is burned
between the maximum ramp weight (MRW) and maximum landing weight (MLW) condition.
Nonetheless, the taxi-in phase still seems to be more critical, even after such corrections are
made. The generic formula to obtain the lateral load factor for the taxi-in phase then becomes

0.498 l 2
N = 2225.7e−ln m ny
. (6.2)

This equation can be rearranged to obtain the lateral load factor as a function of the layout of the
gears, and the probability of an occurrence of the obtained factor in 1000 flights. If we assume
that the life of an aircraft is no more than 100,000 flights, then the value of N becomes 0.01.
Figure 6.2(a) contains a contour map of the maximum lateral load factor that can be expected at
least once in 100,000 flights, for different track width and wheel base combinations. Existing
aircraft are overlaid onto the contour map, which shows that smaller aircraft experience higher
lateral load factors than larger aircraft. This maximum value is still significantly lower than
the 0.5g specified by the regulation. It can be seen that a maximum value of 0.28g is predicted
for the A320, which has never been measured in the tests (see Figure 6.1(b)), and a value of
between 0.16 and 0.17 is predicted for the A380.
The authors of [69] then insert the dimensions of the Boeing 737-400 into this equation, and
calculate the probability of experiencing a 0.5g lateral load factor. This probability is used
as a baseline for a comparison with other aircraft. Equation (6.2) is then used to obtain the
maximum lateral expected load. Figure 6.2(b) shows the results from this study [69]. It is
apparent from both parts of Figure 6.2 that the gear positioning and size of the aircraft have a
major influence on the lateral load factor that can be achieved. This information can now be
used by the FAA for future considerations pertaining to the regulation.

6.3 Aircraft Loads from the Static Balance Equations


The high-speed turn regulation is interpreted by aircraft manufacturers to require the aircraft to
conduct a turn by using nose wheel steering or differential power. A lateral acceleration of 0.5g
70 Chapter 6. High-Speed: Ground Loads Requirements

20 0.6
0.15
(a) (b) Taxi-out

Lateral Load Factor ny (g)


0.5 Taxi-in
Track width lm (m)

15 0. 2 0.4
A380

C5 B777-300ER 0.3
B747-8
MD-11
A330-200
10 0.25 B787 A340-600 0.2
An-124
A320 A321 0.1
0.3 B737-900
0.35 B737-400 MD-81
5 0
10 15 20 25 30 35 0 82/83 -320
7-40M ER 400
B-73 D- A 7-200B-747-
Wheel base ln (m) B-76
Aircraft

Figure 6.2. Predicted lateral load factors obtained from FAA operational loads study [69]. The contours
in panel (a) represent the maximum load that can be expected at least once in 100,000 flights, while panel
(b) compares the equal probability lateral load factors during ground turning for five aircraft; reproduced
from [69].

and a vertical acceleration of 1g at the centre of gravity are considered. This section shows how
the forces can be obtained from the static balance equations for an aircraft with three landing
gears.
The aircraft is assumed to be in the level position, hence no roll angle is present. The lateral
load at each gear is set to be half its vertical load. This scenario is depicted in Figure 6.3(b).
The critical centres of gravity are chosen in accordance with the general requirement for ground
loads (FAR-25.471), hence the range must be selected so that the maximum design loads are
obtained in each landing gear element. Thus, both maximum forward and aft centre of gravity
positions are investigated. Also, concerning the weight of the aircraft, FAR-25.489 states that
unless otherwise prescribed, the landing gear and aeroplane structure must be investigated for
the aeroplane at the MRW. No wing lift may be considered. The shock absorbers and tyres may
be assumed to be in their static position. Finally, the runway conditions are assumed to be dry.
This is significant, because patches of ice on the runway could reduce the friction on a specific
gear, with a subsequent load transfer to other gears [24].
The loads can then be calculated by considering the static load balance of the aircraft. The
vertical loads at the nose and main gear positions can be obtained by deriving the static balance
equations from Figure 6.3(a). The thrust is ignored for these calculations (but is used later for
the continuation analysis). The sum of the vertical forces at the tyres needs to equal the weight
of the aircraft, while the moments around the CG also need to be in balance. The forces on
the main gears also need to resist the rolling moment that is created by a lateral load factor
ny , and consequently the left-hand gear will see larger forces than the right-hand gear, for a
turn to the right; as before, assume that a turn is made to the right, from the pilot’s perspective.
Figure 6.3(b) contains the forces and dimensions of interest for the lateral loading of the gears.
The forces at each main gear can then be calculated as
6.3. Aircraft Loads from the Static Balance Equations 71

mg hc
x
lcm lcn
z
Fzm Fzn

(a) Side view.

0.5mg
mg
y
0.5Fzmr 0.5Fzn 0.5Fzml
z
Fzmr Fzn Fzml
lm
(b) Front view.

Figure 6.3. FAR25.495 lateral loads requirement depicted in free-body diagrams.

lcm
 
0
lcn + lcm
  
Fzn
 

 Fzml  =  lcn ny hc  mg

. (6.3)
 mg

lcn + lcm lm

Fzmr

 lcn n y hc 

lcn + lcm lm

Figure 6.4(a) depicts the vertical forces at the main gears for an A320 that are obtained from
the static balance equations, in accordance with the regulation, as well as quasi-steady results
obtained from a dynamic model. The main differences between the two methods of calculation
is the absence of oleos, aerodynamics, and tyre properties, for the regulatory method. The
steering angle in the model is set to 15◦ , and then the velocity is ramped up gradually from
1 to 12 m/s. This provides forces that are close to the equilibrium values. An increase in the
lateral load factor ny at the CG, causes an increase of the vertical force on the outer gear, and a
decrease of the force on the inner gear. The difference between the forces on the gears is larger
for the method proposed by the regulation, due to the absence of aerodynamics. No equilibrium
results were obtained in the model for ny > 0.27, due to a loss of aircraft stability above these
values [13, 54, 55]. This is discussed in the next section. The area to the right of the dashed
72 Chapter 6. High-Speed: Ground Loads Requirements

600 0.6
Regulation Regulation
500 Model 0.5 Model
Fzml
400 0.4

ny at MLG’s
nymr
Fz (kN)

300 Unstable 0.3 Unstable


Fzmr nyml
200 0.2

100 0.1
(a) (b)
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
ny at CG ny at CG

Figure 6.4. Comparison of the results from the static balance equations (in accordance with the regu-
lation), and the results from a dynamic model of an A320. Panel (a) depicts the vertical loads on the
main gears (Fzml = force on left gear and Fzmr = force on right gear), as the lateral load factor ny is
varied at the CG. Panel (b) depicts the lateral load factors at the gears. The MRW at the furthest aft CG
position is used.

lines indicate this unstable region. The regulation requires the most extreme positions for the
CG, and therefore Table 6.1 contains the gear forces for the MRW of 73900 kg, at the extreme
fore and aft CG positions.
There are two equations that can be used to calculate the three lateral wheel forces: a balance
of the forces in the y-direction and the moments around the z-axis. The system in Figure 6.4(b)
is underdetermined, and therefore a lateral load factor needs to be chosen for the wheels. The
biggest assumption related to this regulation is that a lateral load factor of 0.5 is present at the
tyres. We assume that the load factors at the gears are equivalent to the load factor at the CG,
which is represented by the diagonal line, and is similar to the regulation. The quasi-steady
results from the simulation show that the inner gear experiences a lateral load factor that is
larger than the regulatory value, while the outer gear experiences a smaller value. If the values
are extrapolated to the regulatory ny value of 0.5 at the CG, the inner gear load factor would
be larger than the 0.5 regulatory value for the gear. The inner gear factor is approximately 0.35
at the onset of limit cycles, where ny = 0.3. The 0.5 regulatory factor is therefore adequate.
Tyre properties are often altered to unrealistic values in simulations, to enable the generation
of such large side load factors. This approach was not followed here.
Statically indeterminate gear arrangements cannot be solved by the previous calculation method,
and therefore dynamic simulations are used for an aircraft such as the A380. Note that the
original regulations were written in the days when tricycle arrangements were prevalent, hence
the implementation of the regulation using static balance would have been adequate for most
aircraft in operation. Simulations and continuation methods would fall under the banner of
rational cases, which are discussed in the next section.
6.4. Continuation Analysis of the High-Speed Turn 73

Table 6.1. Vertical loads for A320 gears at maximum ramp weight and 0.5 lateral loading condition.

CG-fwd CG-aft
lcn (m) 10.96 11.92
lcm (m) 1.83 0.87
Fzn (N ) 104019 49334
Fzml (N ) 519772 547115
Fzmr (N ) 101168 128511

6.4 Continuation Analysis of the High-Speed Turn

Chapter 5 showed that the lateral load factor at the CG position tends to build up gradually
towards the steady-state condition when a typical runway exit manoeuvre is conducted, this is
due to the exponential nature of the input function. A runway exit manoeuvre is essentially a
transition from a straight line motion to a steady-state circular trajectory. Section 5.4.2 shows
that the steering angle for the A320 converges to the steady-state value for a 90◦ exit, hence
this is the maximum value that will be reached. The steering angle for the A380, on the other
hand, does not reach the steady-state value at a 90◦ exit for a category VI airport, but does come
close to this value when a 135◦ exit is used [14]. We can therefore conclude that the steady-
state steering angle values would in fact be the maximum steering angle values that could be
experienced during a turn. We will also assume that the steady-state lateral load factor at the
CG position will be the maximum value. The steady-states can therefore be used to study the
loads that can be experienced during ground manoeuvres. This is consistent with FAR 25.495.
Continuation analysis provides these steady-state values for different steering angle and thrust
values, and hence provides a means in Sections 6.4.1 and 6.4.2 for the analysis of the high-
speed turn. Previous ground dynamics studies of the A320 aircraft [13, 54] attributed fold
and Hopf bifurcations to certain wheels that could not maintain the required force at the tyre-
runway interface. The Hopf bifurcations could, for instance, be attributed to the loss of grip of
the inner main gear tyres. A further study by Rankin et al. [56] showed that there was a strong
correlation between the measured data from the FAA operational loads study and the results
from dynamic simulations. The models that were used in this study did not contain oleos,
while the axle widths on all the gears were set to zero. This was done to obtain significant
improvements in the simulation times, while maintaining enough accuracy with regards to the
stability characteristics. In the following sections we include all the effects that were omitted
in the A320 model in References [55, 56], and also expand the analysis to the A380.

6.4.1 Load Factors for an A320

A similar approach is taken here as in the earlier studies [13, 54]. The initial steering angle is
set to zero and then a velocity controller is used to find equilibrium states at this target velocity,
as explained in Section 2.4. The velocity controller is then switched off and the engine thrust is
set to a constant value. The steering angle is then used as the continuation parameter, increasing
to the maximum steering angle. The combination of all these different runs then allows for the
construction of a bifurcation surface [54]. In this study we adhere to the configurations as
74 Chapter 6. High-Speed: Ground Loads Requirements

stipulated by FAR 25.495, by considering the MRW condition at the extreme forward and aft
CG positions.
Figure 6.5 shows the lateral load factor at the CG in a (δ, Vn )-projection of equilibria, con-
structed by bifurcation analysis. The top row represents the forward CG position, while the
bottom row represents the aft CG position, at the MRW condition. The regulation states that
the analysis needs to be done without aerodynamics and thrust. Cases that omit (left-hand pan-
els) and include (right-hand panels) the aerodynamics are included here, to highlight the impact
of the aerodynamics. Thrust is essential for the continuation analysis to work correctly, and is
therefore included in all the analysis. The inclusion of thrust does in fact represent more severe
loading conditions and would be more representative of reality [13]. The boundary between the
shaded and white areas in each figure represent the Hopf bifurcations (labelled H) that were
found in the original studies, indicating the onset of oscillatory behaviour [13, 15, 54, 55].
The contours labelled 30, 45, 51, 275 and 500, in Figure 6.5, indicate the steady-state radius in
metres that the nose gear will follow, and they are related to the exit radii for different airport
categories in Table 5.2. The line labelled 51 represents a runway exit at a category VI airport;
it shows that a steering angle of approximately 15◦ is needed to maintain a radius of 51m at
the nose gear. The thick contour line between the 0.1 and 0.15 lines in the plots, represents
the 0.133 ICAO lateral load condition, which is used for runway exit designs, as discussed
in Section 5.4.1. The intersection points between the radii contours and the 0.133 contour
provide the maximum steering angles and design exit velocities that can be used, according to
the ICAO design rules. For example, point C in Figure 6.5(a2) indicates that a design runway
exit velocity of approximately 8 m/s should be used for a 90◦ exit at a Category VI airport,
which is consistent with the values in Table 5.2 that were derived from the ICAO method.
A comparison of Figure 6.5(a1) and (a2) shows that the aerodynamics has a stabilising effect
at higher velocities: the unstable region is moved to the right. Both diagrams are very similar
for velocities below 10 m/s. This added stability at higher speeds is not of any real benefit for
this configuration due to the restrictions placed by the design velocities. Point A represents the
25.5 m/s design velocity for a high-speed exit as contained in Table 5.2. This intersection point
shows that a maximum steering angle of approximately 2◦ would be required for a high-speed
exit. In both cases point A falls within the stable region. The aft CG position is more critical,
as is shown in Figures 6.5(b1) and (b2), as is evident from the movement of the Hopf-curve
towards the lower left-hand corner. Panel (b1) is for no aerodynamics, and panel (b2) has this
effect included. It is clear from panel (b1) that the 25.5 m/s exit velocity, as prescribed by the
ICAO design rules, would be too high: point A falls within the unstable region. The aerody-
namics once again plays a stabilising effect; panel (b2) shows that an aft CG configuration is
less stable than a forward CG configuration. This is to be expected.
FAR 25.495 aims to cover the extreme loading cases, hence another useful comparison is to
plot the envelope of the maximum load factors at the CG and tyre positions. The data points
from the bifurcation analysis can be represented as a cloud of individual points, where each
point is associated with a steering angle, velocity and lateral load factor. All the data points
are projected onto the (δ, ny ) plane. Maximum load factor envelopes for the CG and tyres
are automatically generated by the Convex-Hull algorithms available in Matlab. Note that all
the stable and unstable solutions are considered, giving the maximum possible values that can
be generated. Figure 6.6 shows the maximum lateral load factors that can be achieved across
the entire envelope for the MRW condition. The same configurations are considered as in
6.4. Continuation Analysis of the High-Speed Turn 75

35 35
Aero: Off (a1) Aero: On (a2)
CG: Forward CG: Forward
30 30

25 A 25 A
Unstable Unstable
Vn (m/s)

20 20

15 15
0.1 0.15 0.2 0.1 0.150.2 B
10 0.25
0.3
H 10 0.25
0.3 H
C
0.133 0.133
5 5 D

500
275
500
275

51
45

30
51
45

30

0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35

35 35
Aero: Off (b1) Aero: On (b2)
CG: Aft CG: Aft
30 30

25 A 25 A
Unstable Unstable
Vn (m/s)

20 20

15 15

0.1 0.1 0.15


B
10 0.15 10
0.2 0.25 H 0.2 0.25 H
0.133 0.133
5 5
500
275

500
275
51
45

30

51
45

30

0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
δ (deg) δ (deg)

Figure 6.5. Lateral load factor contours at the CG for the A320 at the MRW, obtained from bifurcation
analysis. The onset of instability is characterised by limit point and Hopf bifurcations. Panel (a1) is
for forward CG position without aerodynamics; Panel (a2) is for the same configuration with aerody-
namics included. Panel (b1) is with an aft CG position; panel (b2) is for the same configuration with
aerodynamics included. The behaviour at points A, B, C and D are compared.
76 Chapter 6. High-Speed: Ground Loads Requirements

0.5 0.5

0.4 0.4

0.3 0.3
ny

0.2 W1 0.2
W2
W3
0.1 W4
0.1
CG
(a1) (a2)
0 0
0 15 30 45 60 75 0 15 30 45 60 75
0.5 0.5

0.4 0.4

0.3 0.3
ny

0.2 0.2

0.1 0.1

(b1) (b2)
0 0
0 15 30 45 60 75 0 15 30 45 60 75
δ (deg) δ (deg)

Figure 6.6. Maximum possible lateral load factors at the CG and wheel positions for the A320, MRW
configuration. Panel (a1) is with a forward CG position, without aerodynamics. Panel (a2) if for the
same configuration with aerodynamic effects included. Panel (b1) is for an aft CG position, while panel
(b2) is for the same configuration, with aerodynamics included.
6.4. Continuation Analysis of the High-Speed Turn 77

Figure 6.5. Note that the load factors in the tyres are increased by 10% from the steady-state
values to account for the overshoot that was observed in Section 5.4.
It is clear that the load factor at the CG of the A320 is not anywhere near the 0.5 value that
is prescribed in the regulation. The inner-most tyres in the turn can experience lateral load
factors that are close to the 0.5 value, but the tyres on the outside of the turn experience load
factors that are significantly lower than the inside tyres, and slightly higher than the factors
experienced at the CG, even though the vertical loads on the outside tyres are larger than the
vertical loads on the inside tyres. This is due to smaller slip-angles on the outside gear. The
velocity of the outside gear is larger than the inside gear, hence a smaller slip-angle is created.
The forward CG position creates a gradual decline in the load factor as δ is increased, while the
aft CG position seems to create a reasonably constant value, with a sudden drop after 60◦ of
steering. The aerodynamics makes a significant difference at small steering angles and relates
to the area around point A in Figure 6.5.

6.4.2 Load Factors for an A380

The A380 nose gear velocity drops very steeply towards lower velocities at small steering an-
gles, when a constant thrust is used. Hence a large section of the envelope is not covered. This
is not the case for the A320. The analysis technique for the A380 is therefore different to the
approach taken for the A320. The velocity controller is not switched off during the continua-
tion runs, and consequently the thrust is allowed to change as the steering angle is varied. A
constant velocity is maintained at the nose gear, allowing for complete coverage of the enve-
lope. This approach is only valid if no bifurcations are found. Bifurcations could indicate that
some dominant engine modes are present if a velocity controller is used, which was indeed the
case when the original studies were done for the A320 [13]. A lack of bifurcations for this
case would mean that the results from a constant thrust or constant velocity approach would
provide equivalent results. Note that the thrust was only applied to the inboard engines, which
is similar to the way in which pilots taxi the aircraft. Figures 6.7(a1) to (b2) contain the re-
sults for the A380 lateral load factor at the CG, with similar MRW configurations as before.
No bifurcations were found, hence no region of instability is present when compared to the
dynamics of the A320. A lack of bifurcations indicate how remarkably stable this aircraft is
when compared to a tricycle arrangement. The area below the 0.133 lateral load factor contour
is again used to define the design envelope. This is the first time that a comprehensive map of
the lateral load factor has been constructed for such a large aircraft, where such a wide range
of steering angles and velocities are covered.
The areas below the 0.133 contour in panels (a1) to (b2) are almost equivalent, showing that the
CG position does not have any significant influence on the loads within the design envelope.
The left-hand panels (a1 and b1) have no aerodynamics included. The inclusion of aerody-
namics in the right-hand panels (a2 and b2) shows that the aerodynamics causes a significant
reduction in the overall loads that can be achieved. The aerodynamics reduces the maximum
load factor by approximately 21% to 0.26 for the forward CG position, and by approximately
34% to 0.23 for the aft CG position, at the MRW. This may seem significant when compared
to the A320, but note that these differences occur again at high-speed outside of the design
envelope.
78 Chapter 6. High-Speed: Ground Loads Requirements

The most significant areas in the graphs in Figure 6.7 are in the bottom right hand corner
of each panel, which corresponds to 90◦ and 135◦ exit manoeuvres at Category VI runway
exits. The lines marked 51 are therefore of relevance. In panels (a2) and (b2) the effect of
the aerodynamics pushes the 51 line closer to the 0.133 contour (the design envelope), when
compared to panels (a1) and (b1). This is surprising when one considers that the velocity is only
approximately 8 m/s. This may seem like a low velocity, but the huge aerodynamic surfaces
of the A380 cause significant forces, even at such low velocities. In this particular case, this is
due to the vertical tailplane. In Figure 6.7(b2) it can be seen that the combination of the aft CG
position and the aerodynamics causes interesting ground handling problems at velocities in the
region of 8 m/s. The 51 m curve transitions from an almost vertical gradient at low velocities
to an almost horizontal gradient at approximately 8 m/s. A smaller force is present at the NLG
due to the aft CG position, hence the NLG tyres saturate earlier when compared to the forward
CG position. This means that no additional force can be generated from the NLG tyres, which
are operating at the limit of their performance [13, 54]. The implication is that the aircraft
cannot conduct a tighter turn than 51m when a velocity of 8 m/s is maintained at the NLG. The
recommended velocity of 4 m/s (8 knots) ensures that these types of turns can be conducted
safely for all configurations.
The curves representing the 275 m and 500 m radius high-speed exits show interesting be-
haviour in regions that the aircraft will never venture into. As an example, let us examine
Figure 6.7(b2). If the aircraft maintains a velocity of 25 m/s at the NLG, and the steering angle
is gradually increased, the radius of the nose gear trajectory will decrease, until the steering
angle reaches a value of approximately 11◦ . The trajectory of the nose gear will maintain an
almost constant radius between 11◦ and 20◦ degrees. This radius will increase after 20◦ as the
steering angle is increased. This is due to the nonlinear nature of the tyre, and it can be seen
in all the panels; the maximum side force available from the tyres is limited, as indicated by
the white region in panels (b1) and (b2). The nose gear tyres in panels (a1) and (a2) are close
to saturation in a similar region, and are operating at approximately 95% of the total available
force. The aircraft is stable within the design envelope at high-speed exits. One could argue
that the close spacing of the contours at low steering angles and high velocities make it easy to
generate significant lateral load factors at high-speed exits. However, a pilot would not over-
steer easily at such exits due to large radii of the turns, and also due to envelope protection laws
in the flight control system.
The lateral loads requirement assumes that the forces on the gears act towards the turn centre,
hence all the forces act in the same direction. This is however not the case for the A380.
Figure 6.8 contains the lateral load factor contours in the (δ, Vn )-plane. Panels (a1) and (a2)
contain the load factors on the WLGs, and (b1) and (b2) the BLGs. The negative contours
at low velocities in panels (a1) and (a2) indicate that the gear forces on the WLGs act in an
opposite direction to that of the BLGs, and change direction (to act in the same direction as
the BLGs) at nose gear velocities above 10 m/s. This effect can be mainly attributed to the
geometric layout of the landing gears, and to a lesser extent the tyre properties. The slip-
angle is positive at low velocities, creating negative lateral loads. An increase in the velocity
decreases the slip-angles, and consequently the lateral loads decrease. This can be observed
in the reduction of the magnitude of the negative contours, in the bottom right-hand corner
of Figures 6.8(a1) and (a2). Velocities above 10 m/s create negative slip-angles, with positive
loads on the WLGs. If we start in the lower left-hand corner of panels (b1) and (b2), an increase
in velocity and steering angle would lead to an increase in lateral load. The opposite is true in
6.4. Continuation Analysis of the High-Speed Turn 79

35 35
0.33
500
(a1) 0.26
(a2)
30 30 500

25 25
275
275
Vn (m/s)

20 20

15 15
B
10 10

5 5
275
500

275
500

51
51

0 0
0 10 20 30 40 50 0 10 20 30 40 50

35 35
0.35 500 (b1) (b2)
30 30 500
0.23
Nose
25 Gear
275
25
Saturated
Nose Nose
Gear Gear 275
Vn (m/s)

20 Saturated 20 Saturated

15 15
B
10 10

5 5
275
500
275
500

51
51

0 0
0 10 20 30 40 50 0 10 20 30 40 50
δ (deg) δ (deg)

Figure 6.7. Lateral load factors at the CG obtained from continuation analysis for the A380 in the MRW
configuration. Panel (a1) is for a forward CG position without aerodynamics; panel (a2) if for the same
configuration with aerodynamics included. Panel (b1) is for an aft CG position; panel (b2) is for the
same configuration with aerodynamics included.
80 Chapter 6. High-Speed: Ground Loads Requirements

35 35
WLGL (a1) WLGR (a2)
30 30

5
0.1

0.2
25 25

0.1

0.15

0.1
Vn (m/s)

0.05
20 20

0.05
15 15
0.05 0.05 0.1
10 0 10
-0.05 -0.05 0
-0.1 -0.1
-0.15
5 5 -0.2

-0.22 -0.26
0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
35 35
BLGL (b1) BLGR (b2)
30 30

0.2
0.2

25 0.23 25
0.2

0.28
Vn (m/s)

0.25
20 20

0.05
0.05
0.2

15 15

0.1
0.
1

0.15 0.1
0.05
0.05

10 10 5
0. 0.1 0.1
1
0.

5 5
1

5 5

0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
δ (deg) δ (deg)

Figure 6.8. Lateral load factor on each gear for the MRW aft CG configuration of an A380.

the top right-hand corners of these panels. The effect of the steering system on the aft axles
of the BLGs is not readily apparent from these graphs; it will be more pronounced in the next
section, when the focus shifts to the tyres.

6.4.3 Individual Tyre Loads for an A380

The extensive amounts of information provided from bifurcation methods allows one to present
the results in a different way. Figure 6.9 is an example. It shows a map of the tyre that is
carrying the largest lateral load as the steering angle and velocity is varied. A complex pattern
emerges for the A380. For example, for Vn = 5 m/s the maximum load switches between five
different tyres. The inner tyre on the aft axle of the body gear, W20, initially generates the
highest load. The body wheel steering system engages at 20◦ and a consequent shift occurs
to W18, then W4, progressing on to W16, and finally to W15. This type of diagram allows
6.4. Continuation Analysis of the High-Speed Turn 81

35

30

25
W17
20 W9
Vn (m/s)
15
W15
10 W9 W13
W15
5 W20
W18 W16 W4
0
0 10 20 30 40 50 60 70
δ (deg)

Figure 6.9. Map of tyres of the A380 that carry the highest load as the steering angle and velocity are
varied. The number in each area indicates the tyre with the highest lateral load. The MRW condition
with an aft CG position is used.

engineers to gain a much improved understanding of the loading in the system, and it is a
pertinent example of how complex the behaviour of such a large aircraft can be.
The last piece of the puzzle is to consider the lateral load factor on individual tyres. Only the
inner-most (W4,W8,W12,W16,W20) and outer-most (W1,W5,W9,W13,W17) tyres are con-
sidered in Figure 6.10. The absolute values are shown. We can observe that the inner tyres
experience a higher load factor than the outer tyres. This is due lower slip-angles on the
outer tyres. The maximum forces at the tyres can easily be obtained for a Group VI runway
exit. Chapter 5 showed that a steering angle of approximately 37◦ represents the steady-state
steering angle for a Group VI exit. If the exit velocity is restricted to 4 m/s, the coordinates
(δ, Vn ) = (37, 4), would indicate the point where the maximum tyre forces can be expected for
this type of exit. These coordinates are represented by the points labelled A in Figure 6.10. If
we then look at the load factors at these points, it can be seen that the first row of tyres on the
wing gear, and the middle row of tyres on the body gear, carry the highest loads. This would be
the most likely scenario in terms of operational velocities and steering angles. An unrealistic
scenario would be at approximately 11 m/s, where the second row of tyres on the wing gears
carry the highest load factors, and most of the other tyres appear to have low load factors.
The body wheel steering does not seem to have much of an effect on the load factors at the wing
gear tyres. There is however a significant effect on the body gear tyres, as can be observed by
the patterns in the contours around a 20◦ steering angle. The body wheel steering does provide
significant load alleviation in the body gear tyres, W17 and W20, which was also apparent in
the simulation results in Section 5.4. The difference between the load factors on the inner and
outer gears are less than for the A320. Most of the wheels are not anywhere near the 0.5 factor
stipulated by the FAA regulation, apart from the inner wheels of the body gear.
We see, therefore, that continuation methods allow for a full classification of the load factors
82 Chapter 6. High-Speed: Ground Loads Requirements

35 35
30 (W1) 30 (W4)

2 5
0.
25 25

0.2
Vn (m/s)

0.1

0.1
0.1
5
20 20

0.05
0.0
15 15
0.
10 0.05 10 0.050.1 05

Wing Landing Gear


0.0.15
1 0.150.2
0.2 0.25
5 A 5 A 0.3
0.30 0.37
0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
35 35
30
0.17 (W5) 30 0.29 (W8)
5

25 25
0.1

0.25
Vn (m/s)

0.2
0.1

0.15
20 20

0.05
0.05

15 15 0.1
10 0.05 0.1 10 0.15
0.050.1
5 0.05 5 0.05
A A 0.1
0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
´ (deg) ´ (deg)

35 35
30 (W9) 30 (W12)
0.2
Vn (m/s)

25 25
0.25

0.2
0.15

20 20
0.05

15 0. 15 0.1 0.05
0.150.21 0.15
0.2
0.25
10 0.0 0.2 0.3 10 0.0 0.1 0.1 0.3
5 0.1 5 0.33 5 5 0.35 0.40
5 A 5 A
0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
35 35
30 (W13) 30 (W16) Body Landing Gear

25 25
Vn (m/s)

20 20
0.05
0.05

0 0.1
15 0.15 .1 15
0.1

0.2 5
0.2
0.250.3 0.25 0.3
0.05

0.35
0.05

0.4
0.1

0.1 1

0.2.2

10 35 10
0.2

0.
0.1

0.
5

0.44
5 A 0.36
5 A
0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
35 35
30 (W17) 30 (W20)
0.15
0.15

25 25
Vn (m/s)

0.15
0.1

0.05

0.1
0.2

0.3

0.25

0.05
0.2

20 20
15 0.2 15
5
0.05
0.05

0.152
0.10.
5
0.15
0.1

0.1

0.25

10 10
0.3
0.0
0.0

0.27 0.43
5 A 5 A
0 2 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
´ (deg) ´ (deg)

Figure 6.10. Contour maps of the absolute steady-state lateral load factor, for the outer-most and inner-
most tyres, during a turn. The panel numbers depict the wheel numbering convention from Figure 2.2.
6.5. Relating the Continuation Results to the FAA Study 83

within the tyres. The exact conditions where these maximum values occur can be identified
very efficiently.

6.5 Relating the Continuation Results to the FAA Study

The final step is to relate the statistics from the FAA study with the results obtained in the
previous sections. Only the more realistic cases with aerodynamics are considered here. Fig-
ure 6.11 compares the envelope that is obtained from the continuation analysis, with the A320
data from the FAA study in [59]. The lateral load factors were recorded over 10066 flights,
where approximately half of the runway exit angles were below 60◦ , a quarter were between
60−120◦ , and a quarter were larger than 120◦ . The outliers at zero velocity are likely to be due
to measurement errors, or could possibly be attributed to gusts on the apron; they can be safely
ignored. All the data of significance therefore lies within the envelope. Runway exits smaller
than 60◦ seem to generate the highest loads; however, there are significantly more data points
for this type of exit. Similar maximum values may arise when more data points are added
for the other exits. It is interesting to note that [59] could not show any statistical correlation
between the exit velocity and the lateral load factor that is generated. The large void at higher
velocities shows that the analysis method covers cases that would not occur operationally. The
lower maximum load factor of 0.33, when compared to the 0.5 value from the regulation, indi-
cates that the method is less restrictive than the regulation, yet seems to be adequate to cover
the operational cases.
The results can also be used to determine the maximum likely lateral load factor in operation. It
can be seen from Figure 6.2(a) that a lateral load factor of approximately 0.28 can be expected
at least once in a lifetime for an A320, and a load factor of 0.17 for an A380. The fact that
the exit velocity and exit type seems to be statistically insignificant, makes the choice of an
operating point difficult. We therefore choose a point that is representative. Point B in Figures
6.5(a2) and (b2) would correspond to a load factor of approximately 0.28 at a runway exit for a
Category VI airport. We will therefore assume that 12 m/s is an extreme exit velocity at a 90◦
exit for the A320 and A380, and also that all 90◦ runway exit manoeuvres are conducted below
this velocity.
The lateral load factor of 0.28 at point B in Figure 6.5(a2) can be reached in one of two ways.
The first scenario is where the entry velocity into the turn is approximately 12 m/s, while the
pilot adjusts the thrust through the turn to try and maintain the velocity. The second scenario
is that the pilot enters the turn at 6 m/s (point D), then increases the thrust to try and maintain
the velocity, over-correcting the thrust above the required value and accelerating through the
turn up to point B. This scenario seems more plausible if it is assumed that pilots adhere to the
rules.
If we assume a maximum nose gear velocity of 12 m/s for the A380 in a turn, and then de-
termine the maximum load factor in the region below 12 m/s, the maximum condition is then
indicated by point B in Figures 6.7(a2) and (b2). Even when we ignore the fact that the air-
craft would be unable to maintain a radius of 51 m at this velocity, we can see that it would be
impossible to obtain a load factor of more than 0.2 for the forward CG configuration in panel
(a2), and 0.17 for the aft CG position in panel (b2). Note that the contours are also spaced
further apart from each other for the case of the A380. Hence, a change in thrust would not
84 Chapter 6. High-Speed: Ground Loads Requirements

0.35

0.30

0.25
Max ny

0.20

0.15

0.10 < 60◦ Turns


60 − 120◦ Turns
0.05
>120◦ Turns
0
0 10 20 30 40 50 60 70 80
Vx (m/s)

Figure 6.11. Comparison of maximum lateral load factor envelope (solid line) obtained from continu-
ation analysis, and data points obtained from FAA study [59]. The MRW aft CG configuration is used
for the envelope calculations.

have as much significance when compared to the closely spaced contours for the A320. The
closely spaced contours of the A320 mean that it is easier to obtain higher load values close to
the unstable area. The value of 0.2 for the forward CG is 10% higher than the expected 0.17
value, while the aft CG correlates exactly with the initial predictions in Figure 6.2(a).

6.6 Discussion

An operational loads study by the FAA showed that large aircraft with statically indeterminate
gear arrangements, such as the A380, do not generate the high loads that are stipulated in the
requirements for a high-speed turn. We therefore compared the loads that can be generated
by a relatively small (A320) and large aircraft (A380) to see how the results compare with the
findings of the FAA. Static equilibrium equations were used to calculate the vertical forces on
the gears of the A320. We showed that assumptions were needed with regards to the lateral
load factor on each gear, due to the underdetermined nature of these equations in the lateral
direction. The FAA therefore assume a 0.5 load factor, and it is this factor that gives rise to
the conservative nature of the regulation. Simulations of the A320 showed that the inner gear
experiences a higher lateral load factor than the outer gear. Extrapolation showed that a higher
lateral load factor than the 0.5 value from the regulation would be experienced at the inner
gear, if it were feasible for the aircraft to generate 0.5g at the CG position. This is however not
possible, as the aircraft loses lateral stability at approximately 0.27g. Static balance equations
cannot be used for the analysis of the gear loads of the A380, due to the underdetermined
nature of the equations in all directions. Continuation methods were therefore used to obtain
the lateral load values for this aircraft type.
Contour maps of the lateral load factor were constructed as a function of the steering angle
and velocity at the nose gear of the A320, for the MRW condition. Different CG positions
6.6. Discussion 85

and aerodynamic configurations were considered. The aerodynamics did not have a significant
effect on the maximum load that could be generated, as the maximum condition tends to occur
at relatively low velocities (below 10 m/s). The results confirmed the findings from the simu-
lations and showed that the load factors on the outer tyres are significantly lower than the load
factors on the inner tyres. This is due to the larger velocity of the outer tyres, and consequent
lower slip-angles. The results were also used to show how the maximum lateral load factors
from the FAA study might have occurred. Similar maps were constructed for the A380 at the
MRW condition with different aerodynamic configurations; they showed far less interesting
dynamics from the dynamical systems perspective.
The demonstrated stability of the A380 during ground manoeuvres is of course desirable from
an engineering perspective. We showed that the aerodynamics plays a significant role in the
alleviation of the lateral loads. It was shown that the aerodynamics causes the nose landing gear
tyres to saturate, moving the effective steering envelope closer to the design envelope specified
by ICAO. The aerodynamics also causes a significant reduction in the lateral load factor, when
compared to the case without aerodynamics. The analysis also showed that, at low velocities,
the lateral loads on the WLGs act in an opposite direction to that of the BLGs. This is contrary
to the assumption made in the regulation, where it is assumed that the forces on the gears
act towards the turn centre. The forces on the WLGs reduce with an increase in velocity and
eventually act in the same direction as the BLGs. Continuation analysis also allows for the
construction of complex maps that show how the tyre forces evolve as the steering angle and
velocity is varied. This is very useful when parameter studies are conducted. We therefore
showed that an aircraft such as the A380 would not exceed a lateral load factor of 0.26 across
the entire envelope – almost half the value of the stipulated regulation.
The last section defined a typical operating envelope for the A380, and it compared the results
from the analysis with the original FAA studies. These results show that a maximum lateral
load factor between 0.17 and 0.2 would be experienced over the life time of the aircraft, and
this correlates very well with the results from the FAA study. This provides additional evidence
that a lateral load factor of 0.5 cannot be reached for such a large aircraft.
Chapter 7

Conclusions and Outlook

7.1 Summary

The objective of this work was to classify the ground dynamics of different sized aircraft across
the entire operational and design envelope. We identified the different methods that can be used
for the analysis of aircraft ground manoeuvres. The nonlinear nature of the tyres, oleos and
aerodynamics makes this a particularly challenging problem. Different ground phases were
defined based on the velocity of the aircraft, consisting of a low-speed (0-6 knots), medium-
speed (6-14 knots) and high-speed (>14 knots) category. The main emphasis was on the use
of industrially tested models of the A320 and A380 in conjunction with a newly developed
dynamical systems software environment (AUTO in Matlab). We showed how these models
are constructed and validated, and how the different analysis methods can be used during each
ground phase. The advantages and disadvantages of each method were highlighted. Bifurca-
tion and continuation methods provided new insights into the dynamics of the A380, and also
showed vastly different dynamics between the A320 and the A380. The bifurcation analysis of
an aircraft with more than three landing gears is new to the literature. The discovery of new
dynamics will lead to a better understanding of the behaviour of in-service aircraft, and will
allow for informed decision making at all stages of an aircraft programme. The use of contin-
uation methods for the analysis of the high-speed turn regulation is also new. It can be used
to determine the maximum lateral load factor across the entire envelope. The exact conditions
where this maximum load factor will occur can now be identified more easily.
In Chapter 2 we discussed the different analysis methods that could be used; kinematic meth-
ods, multibody simulations and bifurcation analysis. We showed how kinematic models are
derived from equations that were originally used for the analysis of truck-trailer jackknifing
studies. We then presented detailed dynamic models that are used on different test platforms,
and showed how these models are constructed. The nonlinear nature of the tyres and oleos
was also discussed. Computational challenges surrounding the use of detailed models were
presented, especially with regards to the construction of diagrams that describe the dynamics
across the entire operating region. Bifurcation and continuation methods emerged as an al-
ternative means for the construction of such diagrams, hence the same dynamic models are
used in a different way. A new Dynamical Systems Toolbox that has incorporated AUTO into
the Matlab environment, will hopefully promote more widespread use of dynamical systems
methods amongst the engineering community.

87
88 Chapter 7. Conclusions and Outlook

In Chapter 3 we analysed the U-turn manoeuvre, which is particularly challenging for a large
aircraft such as the A380. A comparison was made between a widely used geometric method,
a simulation-based approach, and a bifurcation analysis approach. The geometric method is a
simple method to use, but engine thrust, tyre and brake inputs are ignored. Hence, the computed
turn radii are generally not reliable as a result of the highly nonlinear nature of the A380
landing gear systems. We then showed how an industrially tested SimMechanics model is
used for simulations of U-turn manoeuvres. We then used the same model to demonstrate
how bifurcation analysis can be used to obtain turn width results that are sufficiently close to
that of the simulations. The advantage of the bifurcation analysis approach is that it is more
efficient (in terms of run-times) and is also able to find qualitative changes in the dynamics.
The presence of a fold bifurcation signifies a change in the dynamics, where the inner wing
landing gear transitions from a forward to a sideways movement. This was not picked up from
simulations. Bifurcation analysis therefore provides additional insights into the dynamics, and
allows one to identify points for further detailed analysis.
In Chapter 4 the aim was to establish criteria for the jackknifing of an aircraft with a towbarless
tug. We presented a physical interpretation of towing manoeuvres, and showed that pushback
manoeuvres are inherently unstable, and that an aircraft that is being towed along a circular
arc will eventually reach a steady-state. The final steady-state radius and towing angle at the
nose landing gear were derived from the kinematic equations in Chapter 2. We concluded that
jackknifing can be avoided by maintaining a towing radius that is larger than the wheel base.
A comparison was made between the tyre forces that are generated when the nose landing gear
follows a circular trajectory – equivalent to an exit radius – under its own power, and when
an aircraft is towed at an equivalent circular trajectory by a towbarless tug. We assumed that
a comparison of the steady-state values that are obtained from a continuation analysis would
be sufficient, even if the transient effects are ignored. For the A380 it was shown that, when
compared to the self-propelled case, the aft tyres of the body landing gear seem to experience
significantly higher forces for the towing case, while the tyres on the wing landing gear seem to
experience lower forces. We can therefore conclude that towbarless towing could be a design
case for the aft axle of the body landing gear, which is offset by load alleviation on other parts
of the gear.
In Chapter 5 we showed how empirical formulae can be derived for the prediction of steering
angles and clearance distances at runway exits. This is useful in the early design stages of an
aircraft project, when very little data is available. The empirical equations were derived from
the results of kinematic simulations, and they showed very good agreement with kinematic and
dynamic simulations. More accurate predictions of the final steering angle can be obtained by
using continuation methods. The results from the kinematic equations were also used for the
prediction of clearance distances at runway exits. The power of the method lies in the fact that
the equations are normalised to the wheel base. Hence, the predictions are valid for any track
width and wheel base combination. We derived a novel graph that can immediately indicate
whether an aircraft would have adequate clearance distances at Category V and VI runway
exits, without the need for any detailed analysis. Simulations were conducted at a runway exit
for a typical operational exit velocity. They showed that the dynamic load factors at the CG
positions, for both the A320 and the A380, build up towards steady-state values. This is also
true for most of the tyres on the main gears of the A380, (apart from the tyres on the aft axle),
where the dynamic load factors are approximately 10% higher than the steady-state values.
The load factors on the main gears of the A320 are approximately 10% higher than the steady-
7.2. Future work 89

state values for the A320. The conclusion is that continuation methods can be used to analyse
the load factors that may occur at runway exits. The obtained values can be used directly on
tyres W1-W16 of the A380, while the steady-state load factors at tyres W17-W20 need to be
factored by 1.1, to include the dynamic effects. A similar approach was followed for the A320,
where the load factor on the main gears are factored by 1.1.
In Chapter 6 the aim was to assess the lateral loading requirement for high-speed turns, as
prescribed by the FAA. We showed in Chapter 5 that the steady-state values can be assumed
to be the most critical values for lateral loading conditions of the gears. Consequently the
results from a continuation analysis can be used to assess the lateral loads that can be generated
during a high-speed turn. Contour maps of the lateral load factors for different parts of the
aircraft were constructed as a function of the steering angle and velocity at the nose gear of
the A320, for the MRW condition. The results show that the load factors at the outer tyres are
significantly lower than the load factors at the inner tyres. This is due to the larger velocity of
the outer tyres, and consequent lower slip-angles. These results were also used to show how
the maximum lateral load factors from the FAA study might have occurred. Similar maps were
constructed for the A380 at the MRW condition with different aerodynamic configurations.
They show far less interesting dynamics from the dynamical systems perspective in that no
bifurcations were detected. On the other hand, we obtained data of operational and design
significance. We showed that the aerodynamics plays a significant role in the alleviation of
the lateral loads. Continuation analysis also allowed for the construction of complex maps that
show how the tyre forces evolve as the steering angle and velocity is varied. This is very useful
when parameter studies are conducted. The maps show that the forces acting on the WLGs
act in the opposite direction when compared to the BLGs, which is contrary to the directions
stipulated in the regulations. These forces on the WLGs decrease with an increase in velocity,
eventually acting in the same direction as the BLGs. The analysis shows that an aircraft such as
the A380 would not exceed a lateral load factor of more than half of the value stipulated by the
regulation. The results correlate very well with the results from the FAA study and provides
additional evidence that a lateral load factor of 0.5 cannot be reached for such a large aircraft.

7.2 Future work

Increased automation of ground operations will mean that different functions (on the aircraft,
and at the airport) will need to interact to obtain optimal performance. A system at the airport
could for instance calculate optimal routing of all the aircraft to the runway threshold, where
each aircraft is given information on arrival times pertaining to specific waypoints. A system
on the aircraft could then calculate the most efficient manner in which these waypoints could
be reached, where braking and fuel burn could be minimised. Such a system would have to
account for delays and weather conditions, as well as random events that could affect the traffic
flow. The ultimate goal from an aircraft point of view would be to design control systems that
can adapt to such operating conditions and changes.
The boundaries that are described by Hopf bifurcations for the A320 could, for instance, be
used as a safety envelope. The velocities and steering angles where the A380 nose gear tyres
saturate could be used as an upper limit for manoeuvring. Tyres contain the largest nonlinear
effects, hence a parametric tyre model would be of great benefit, where future studies could
90 Chapter 7. Conclusions and Outlook

use tyre parameters as the bifurcation parameters. It would be of great interest to see how the
bifurcation diagrams change with changes in the lateral stiffness of the tyre. Work has been
done in this area on nose gear vibrations [66], but not on ground manoeuvres. Further work is
also required to understand the transient effects of exit manoeuvres. We know that the aircraft
needs to be decelerated along the fuselage axis, and accelerated around the yaw axis, where
the nose gear tyres need to provide this combined force. This transient effect is not currently
accounted for at present. We are now able to categorise the dynamics for any aircraft without a
detailed control system. The next logical step would be the inclusion of such control systems.
An understanding of how the regions of safe operation change when controllers are added will
be vital. The fact that the dynamics of the A320 is significantly different from that of the A380
indicates that very different control systems may be needed on each aircraft. Hence a “one-
size-fits-all” philosophy will likely not apply. In any evaluation of controlled ground dynamics,
continuation methods will be very useful.
It is the author’s opinion that a great challenge lies in the industrialisation of dynamical sys-
tems methods. In spite of their huge potential, bifurcation methods are presently being used
only by small pockets of engineers in the aviation industry. In fact, when one wants to intro-
duce nonlinear dynamics into the engineers’ normal toolsets one encounters both societal and
technological challenges. Primarily, the societal ones relate to management support and educa-
tion. The technology needs to be supported by all tiers of management, and a strong business
case needs to be made to gain this support. The technological challenge is one of education and
development of the right tools. Training is needed to familiarise engineers with the vocabulary
and tools of dynamical systems theory, which are still largely unknown to the average engineer.
Indeed, there is a need to learn how to formulate a problem in a way conducive to nonlinear
analyses, and how to interpret the results. A level of intuition similar to that concerning, say,
Bode diagrams, needs to be developed for the interpretation of bifurcation diagrams. At the
same time more emphasis should be placed on the development of well-documented, indus-
trial, integrated toolsets for nonlinear analysis. The Dynamical Systems Toolbox is a first step
towards industrialisation, but it needs additional examples of relevance for the aerospace engi-
neer. The formation of an interest group may be needed in the long-term, to lead a coordinated
industrialisation effort.
References

[1] Advisory Council for Aeronautics Research in Europe, 2008 Addendum to the Strategic
Research Agenda, On the WWW, October 2008, URL http://www.acare4europe.
com/docs/ACARE/2008/Addendum.pdf [Retrieved on 15 October 2009].

[2] Airbus Operations, A380 Airplane Characteristics Manual, On the WWW, Octo-
ber 2009, URL http://www.airbus.com/fileadmin/media/gallery/files/
tech/data/AC/AC/A380oct09.pdf [Retrieved on 15 October 2009].

[3] , Aircraft Characteristics, On the WWW, October 2009, URL


http://www.airbus.com/en/services/customer-services/
maintenance-engineering/tech-data/aircraft-characteristics [Re-
trieved on 15 October 2009].
[4] , Global Market Forecast 2009-2028, On the WWW, September 2009, URL
http://www.airbus.com/en/corporate/gmf/ [Retrieved on 18 September 2009].

[5] , Airbus MoU with IAI to Explore Eco-efficient “Engines-off” Taxiing, On


the WWW, March 2010, URL http://www.airbus.com/en/presscentre/
pressreleases/pressreleases/items/09/06/17/eco/efficient/
taxiing.html[Retrieved on 25 March 2010].

[6] J.C. Alexander and J.H. Maddocks, On the Maneuvering of Vehicles, SIAM Journal on
Applied Mathematics 48 (1988), no. 1, 38–51.
[7] Anon., Antonov Basic Technical Data, On the WWW, June 2010, URL http://www.
aviastar-sp.ru/aviastar/en/aircraft/an124/an124.htm [Retrieved on 24
June 2010].
[8] , Lockheed C-5 Galaxy, On the WWW, June 2010, URL http://www.
militaryaviation.eu/transporter/Lockheed/C-5.htm [Retrieved on 24 June
2010].
[9] M. Blundell and D. Harty, The Multibody Systems Approach to Vehicle Dynamics, SAE,
September 2004.
[10] Boeing, Commercial Airplanes, On the WWW, October 2009, URL http://www.
boeing.com/commercial [Retrieved on 15 October 2009].

[11] S.T. Chai and W.H. Mason, Landing Gear Integration in Aircraft Conceptual Design,
Tech. Report MAD 96-09-01, Virginia Polytechnic Institute and State University, Multi-
disciplinary Analysis and Design Center for Advanced Vehicles, 1996.

91
92 REFERENCES

[12] G.A. Charles, M.H. Lowenberg, D.P. Stoten, X. Wang, and M. di Bernardo, Aircraft
Flight Dynamics Analysis and Controller Design using Bifurcation Tailoring, AIAA
Guidance Navigation and Control Conference (2002), no. AIAA-2002-4751.

[13] E.B. Coetzee, Nonlinear Aircraft Ground Dynamics, Master’s thesis, University of Bris-
tol, 2006.

[14] E.B. Coetzee, B. Krauskopf, and M. Lowenberg, Analysis of Medium-Speed Runway


Exit Maneuvers, AIAA Modelling and Simulation Technologies Conference (AIAA, ed.),
no. AIAA-2010-7616, American Institute of Aeronautics and Astronautics, AIAA, Au-
gust 2010.

[15] , Application of Bifurcation Methods to the Prediction of Low-Speed Aircraft


Ground Performance, Journal of Aircraft 47 (2010), no. 0021-8669, 1248–1255.

[16] , The Dynamical Systems Toolbox: Integrating AUTO into Matlab, 16th
US National Congress of Theoretical and Applied Mechanics (USNCTAM, ed.),
no. USNCTAM2010-827, US National Congress of Theoretical and Applied Mechanics,
AIAA, July 2010.

[17] N.S. Currey, Aircraft Landing Gear Design: Principles and Practices , American Institue
of Aeronautics and Astronautics, 1988.

[18] D. De Falco, E. Pennestri, and L. Vita, The Udwadia-Kalaba Formulation: A Report on


its Numerical Efficiency in Multibody Dynamics Simulations and on its Teaching Effec-
tiveness, Multibody Dynamics (2005), 21–24.

[19] E.J. Doedel, with major contributions from A.R. Champneys, T.F. Fairgrieve,
Y.A. Kuznetsov, B.E. Oldeman, R.C. Paffenroth, B. Sandstede, X.J. Wang, and C. Zhang.
AUTO-07P: Continuation and Bifurcation Software for Ordinary Differential Equations;
available at http://cmvl.cs.concordia.ca/.

[20] European Comission, CleanSky Joint Technology Initiative, On the WWW, October 2002,
URL http://www.cleansky.eu/ [Retrieved on 24 June 2010].

[21] Federal Aviation Administration, Part 25, 1980, pp. 252–439.

[22] , Advisory Circular, Airport Design, Tech. Report 150/5300-13, 1989.

[23] , Next Generation Air Transportation System (NextGen), On the WWW, February
2007, URL http://www.faa.gov/about/initiatives/nextgen/ [Retrieved on
24 June 2010].

[24] E. Finn, R. Gleich, K. Green, R. Saccarelli, and M. Szot, Investigation of Limit Design
Lateral Ground Maneuver Load Conditions, Tech. Report DOT/FAA/AR-07/38, Federal
Aviation Administration, June 2007.

[25] T.V. Fossum and G.N. Lewis, A Mathematical Model for Trailer-truck Jackknifing, SIAM
Review 23 (1981), no. 1, 95–99.

[26] T.D. Gillespie, Fundamentals of Vehicle Dynamics (R114), SAE International, March
1992.
REFERENCES 93

[27] A. Gomes de Barros and S. Chandana Wirasinghe, New Aircraft Characteristics Re-
lated to Airport Planning, On the WWW, URL www.ucalgary.ca/EN/Civil/
NLAircraft/Atrgpap.pdf, [Retrieved on 1 May 2010].

[28] W. Govaerts and Y.A. Kuznetsov, MATCONT Continuation Software in MATLAB, On the
WWW, April 2009, URL http://www.matcont.ugent.be/.

[29] J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Systems and Bifur-
cations of Vector Fields, Applied Mathematical Sciences Vol. 42, Springer, 1983.

[30] G.F. Hayhoe and J. Patterson, Towbarless Towing Vehicle Operations - Evaluation of
Braking Action and Vehicle Conspicuity, Tech. Report PB2010-100844, Federal Aviation
Administration, 2009.

[31] International Civil Aviation Organization, Aerodrome Design Manual (Doc 9157) Part
2: Taxiways, Aprons and Holding Bays, Tech. Report 9157,AN/901, International Civil
Aviation Organization, 2005.

[32] T. Jones, J.W. Rustenburg, D.A. Skinn, and D.O. Tipps, Study of Side Load Factor During
Aircraft Ground Operations, Tech. Report DOT/FAA/AR-05/7, Federal Aviation Admin-
istration, March 2005.

[33] D.H. Klyde and R.W. Allen, Tire Manuevring Properties and their Effect on Vehicle Han-
dling and Stability, March 2001, Paper no 577.

[34] D.H. Klyde, T.T. Meyer, R.E. Magdaleno, and J.G. Reinsberg, Identification of the Dom-
inant Ground Handling Charactersitics of a Navy Jet Trainer, Journal of Guidance, Con-
trol, and Dynamics 25 (2002), no. 3, 546–552.

[35] D.H. Klyde, T.T. Myers, R.E. Magdaleno, and J.G. Reinsberg, Development and Eval-
uation of Aircraft Ground Handling Maneuvers and Metrics, AIAA Atmosperic Flight
Mechanics Conference N/A (2001), no. AIAA-2001-4011.

[36] D.H. Klyde, E. Sanders, J.G. Reinsberg, and A. Kokolios, Flight test Evaluation of a Sta-
bility Augmentation Steering System for Aircraft Ground Handling, Journal of Guidance,
Control, and Dynamics 27 (2004), no. AIAA-2003-5318.

[37] Y.A. Kuznetsov, Elements of Applied Bifurcation Theory, Applied Mathematical Sciences,
Vol. 112, Springer-Verlag, September 1998.

[38] H.G. Kwatny, J. Dongmo, B.C. Chang, G. Bajpai, M. Yasar, and C. Belcastro, Aircraft
Accident Prevention: Loss-of-Control Analysis, 2009.

[39] Z. Liu and G. Payre, Global Bifurcation Analysis of a Nonlinear Road Vehicle System,
Journal of Computational and Nonlinear Dynamics 2 (2007), no. 4, 308–315.

[40] Z. Liu, G. Payre, and P. Bourassa, Nonlinear Oscillations and Chaotic Motions in a Road
Vehicle System with Driver Steering Control, Nonlinear Dynamics 9 (1996), no. 3, 281–
304.
94 REFERENCES

[41] M.H. Lowenberg, Bifurcation Analysis of Multiple-Attractor Flight Dynamics, Philo-


sophical Transactions: Mathematical, Physical and Engineering Sciences 356 (1998),
no. 1745, 2297–2319.

[42] Mathworks, Model and Simulate Mechanical Systems with SimMechanics, On the WWW,
2010, URL http://www.mathworks.com/products/simmechanics/ [Re-
trieved on 3 December 2010].

[43] R.K. Mehra, J.V. Carroll, and SCIENTIFIC SYSTEMS INC CAMBRIDGE MA., Global
Stability and Control Analysis of Aircraft at High Angles-of-Attack, Defense Technical
Information Center, 1978.

[44] D.J. Mitchell, Calculation of Ground Performance in Take-off and Landing, Data Sheet
85029, ESDU, 1985.

[45] MSC Software Corporation, ADAMS Multibody Dynamics, On the WWW, 2010, URL
http://www.mscsoftware.com/Products/CAE-Tools/Adams.aspx [Re-
trieved on 3 December 2010].

[46] NASA, National Plan for Aeronautics Research and Development and Related Infras-
tructure, On the WWW, December 2007, URL http://www.ostp.gov/aeroplans
[Retrieved on 15 October 2009].

[47] D. Negrut and B. Harris, ADAMS Theory in a Nutshell, Department of Mechanical Engi-
neering, The University of Michigan (2001).

[48] V. Nguyen, G. Schultz, and B. Balachandran, Lateral Load Transfer Effects on Bifurca-
tion Behavior of Four-Wheel Vehicle System, ASME Journal of Computational and Non-
linear Dynamics 4 (2009), no. 4, 041007, URL http://link.aip.org/link/?CND/
4/041007/1.

[49] V. Nguyen and G. A. Schultz, Vehicle Handling, Stability, and Bifurcaiton Analysis for
Nonlinear Vehicle Models, Master’s thesis, Department of Mechanical Engineering, Uni-
versity of Maryland, 2005.

[50] H.B. Pacejka, Tyre and Vehicle Dynamics, Elsevier, 2006.

[51] A. Prencipe, A. Davies, and M. Hobday, The Business of Systems Integration, Oxford
University Press, USA, 2004.

[52] J. Pritchard, Langley Research Center, United States. National Aeronautics, Space Ad-
ministration, and US Army Research Laboratory, An Overview of Landing Gear Dynam-
ics, Journal of Aircraft 38 (2001), no. 1, 130–137.

[53] J. Rankin, E. Coetzee, B. Krauskopf, and M. Lowenberg, Nonlinear Ground Dynamics


of Aircraft: Bifurcation Analysis of Turning Solutions, AIAA Modeling and Simulation
Technologies Conference (2008), no. AIAA-2008-7099.

[54] , Bifurcation and Stability Analysis of Aircraft Turning on the Ground, AIAA
Journal of Guidance, Control, and Dynamics 32 (2009), no. 2, 500–511.
REFERENCES 95

[55] J. Rankin, B. Krauskopf, M. Lowenberg, and E. Coetzee, Operational Parameter Study


of Aircraft Dynamics on the Ground, ASME Journal of Computational and Nonlinear
Dynamics (2010), no. CND-09-1022.
[56] J. Rankin, B. Krauskopf, M.H. Lowenberg, and E.B. Coetzee, Nonlinear Analysis of
Lateral Loading During Taxiway Turns, Accepted for publication - Journal of Aircraft
(2010).
[57] J.W. Rustenburg, D.A. Skinn, and D.O. Tipps, Statistical Loads Data for a Boeing 737-
400 Aircraft in Commercial Operations, Tech. Report DOT/FAA/AR-98/28, Federal Avi-
ation Administration, August 1998.
[58] , Statistical Loads Data for Boeing 767-200ER Aircraft in Commercial Opera-
tions, Tech. Report DOT/FAA/AR-00/10, Federal Aviation Administration, March 2000.
[59] , Statistical Loads Data for the Airbus A320 Aircraft in Commercial Operations,
Tech. Report DOT/FAA/AR-02/35, Federal Aviation Administration, April 2002.
[60] A.A. Shabana, Dynamics of Multibody Systems , 2nd ed., Cambridge University Press,
October 2003.
[61] J. Stergiopoulos and S. Manesis, Anti-jackknife State Feedback Control Law for Nonholo-
nomic Vehicles with Trailer Sliding Mechanism, International Journal of Systems, Control
and Communications 1 (2009), no. 3, 297–311.
[62] S.H. Strogatz, Nonlinear Dynamics and Chaos, Springer, 2000.
[63] J.M.T. Thompson and F.B.J. Macmillen (Eds.), Nonlinear Flight Dynamics of High-
Performance Aircraft, Philosophical Transactions of the Royal Society A 356 (1998),
no. 1745, 2165–2333.
[64] P. Thota, B. Krauskopf, and M. Lowenberg, Modeling of Nose Landing Gear Shimmy
with Lateral and Longitudinal Bending and a Non-Zero Rake Angle, AIAA Modeling
and Simulation Technologies Conference (2008), no. AIAA-2008-7099.
[65] , Interaction of Torsion and Lateral Bending in Aircraft Nose Landing Gear
Shimmy, Nonlinear Dynamics 57 (2009), no. 3, 455–467.
[66] , Nonlinear Analysis of the Influence of Tyre Inflation Pressure on Nose Land-
ing Gear Shimmy, AIAA Modeling and Simulation Technologies Conference (2009),
no. AIAA 2009-6230.
[67] , Parameter Studies of Mode Interactions in Nose Landing Gear Shimmy, MATH-
MOD 2009 - 6th Vienna International Conference on Mathematical Modelling 284
(2009), 2643–2646.
[68] , Bifurcation Analysis of Nose-Landing-Gear Shimmy with Lateral and Longitu-
dinal Bending, Journal of Aircraft 47 (2010), no. 1, 87–95.
[69] D.O. Tipps, J. Rustenburg, D. Skinn, and T. DeFiore, Side Load Factor Statistics From
Commercial Aircraft Ground Operations, Tech. Report UDR-TR 2002-00119, Federal
Aviation Administration, U.S. Department of Transportation, Federal Aviation Adminis-
tration, Office of Aviation Research, Washington, DC 20591, January 2003.
96 REFERENCES

[70] D.A. Venn, The Application of Bifurcation Analysis To Road Vehicle Dynamics, Ph.D. the-
sis, Department of Aerospace Engineering, University of Bristol, 2006.

[71] S. Vergnat, Heading Control Function (HCF), 16th Performance and Operations Confer-
ence, Airbus, Airbus, 2009.

[72] Y. Wang, J.A. Linnett, and J.W. Roberts, Motion Feasibility of a Wheeled Vehicle with a
Steering Angle Limit, Robotica 12 (2009), no. 03, 217–226.

[73] J.Y. Wong, Theory of Ground Vehicles, 3rd ed., Wiley-Interscience, March 2001.

[74] J.R. Wright and J.E. Cooper, Introduction to Aircraft Aeroelasticity and Loads, Wiley,
2007.

[75] Q. Zhu and M. Ishitobi, Chaos and Bifurcations in a Nonlinear Vehicle Model, Journal of
Sound and Vibration 275 (2004), no. 3-5, 1136–1146.
Appendix A

Aircraft Definition

The following pages contain drawings and mass information for the A320 and A380, obtained
from [3].

97
98 Chapter A. Aircraft Definition

SCALE METERS SCALE FEET


0 3 6 9 0 10 20 30

37.57 m
(123 ft 3.12 in)

3.95 m
(12 ft 11.4 in)
5.755 m
(18 ft 10.32 in)

11.19 m
CFM56
(36 ft 8.52 in)
11.12 m
V2500
(36 ft 5.76 in) NOTE: FOR DOOR SIZES
SEE CHAPTER 2 7
4.14 m
(13 ft 7 in)

SEE
CHAPTER
23

5.07 m 12.64 m
(16 ft 7.56 in) (41 ft 5.52 in)
34.10 m
(111 ft 10.44 in)

12.45 m
(40 ft 10.08 in)

7.59 m
(24 ft 10.8 in)

Aircraft Characteristics
WV010 WV011 WV012 WV013 WV014
Maximum Landing Kilograms 64 500 66 000 66 000 64 500 64 500
Weight (MLW) Pounds 142 198 145 505 145 505 142 198 142 198
Maximum Zero Fuel Kilograms 61 000 62 500 62 500 61 000 61 500
Weight (MZFW) Pounds 134 482 137 789 137 789 134 482 135 584
Estimated Operational CFM Engines 41 244 kg (90 927 lb)
Empty Weight (OEW) IAE Engines 41 345 kg (91 150 lb)
Estimated Maximum Kilograms 19 756 21 256 19 756 20 256
Payload CFM 56 Pounds 43 555 46 861 43 555 44 657
Estimated Maximum Kilograms 19 655 21 155 19 655 20 155
Payload IAE V2500 Pounds 43 332 46 639 43 332 44 434

Figure A.1. A320 dimensions and mass information.


99

72.571 m (238.09 ft)


68.854 m (225.90 ft)
46.972 m (154.11 ft) 3.983 m
(13.07 ft)
52.065 m (170.82 ft)

29.943 m (98.23 ft)

22.228 m (72.92 ft)

72.727 m (238.61 ft)

70.400 m (230.97 ft)


8.557 m (28.07 ft)
8.410 m (27.60 ft)

METERS
4.972 m (16.31 ft) 0 4 8 12 16
33.578 m (110.16 ft) WLG
FEET
36.854 m (120.91 ft) BLG 0 10 20 30 40 50

79.750 m (261.65 ft)


30.372 m (99.65 ft)
7.142 m
(23.43 ft)

NOTE: FOR DOOR


DIMENSIONS AND LOCATION,
5.264 m
SEE CHAPTER 2 7.
(17.27 ft)
12.456 m
(40.87 ft)
29.598 m (97.11 ft)
51.400 m (168.64 ft)
DB1A L_AC_020200_1_0010101_01_01

Aircraft Characteristics
WV004 WV005
Maximum Ramp Weight 562 000 kg 562 000 kg
(MRW) (1 238 998 lb) (1 238 9998 lb)
Maximum Taxi Weight
(MTW)
Maximum Take Off 560 000 kg 560 000
Weight (MTOW) (1 234 588 lb) (1 234 588 lb)
Maximum Landing 391 000 kg 386 000 kg
Weight (MLW) (862 007 lb) (850984 lb)
Maximum Zero Fuel 366 000 kg 366 000 kg
Weight (MZFW) (806 892 lb) (806 892 lb)
Estimated Operating With Trent 970 Engines : 270 364 kg (596 050 lb)
Empty Weight (OEW) With GP 7270 Engines : 270 630 kg (596 637 lb)

Figure A.2. A380 dimensions and mass information.


Appendix B

State Definition

Table B.1: State names. † These states are held constant.

Component Constraint Degrees Description A320 A380


of freedom State State

Nose Landing Gear Prismatic 2 Oleo stroke 1 1


Oleo velocity 1 1
Revolute 2 Oleo rotation angle 1 1
Oleo rotational velocity 1 1
Rotational motion -2 Oleo rotation angle -1 -1
Oleo rotational velocity -1 -1

Left Wing Landing Gear Prismatic 2 Oleo stroke 1 1


Oleo velocity 1 1
Revolute 2 Oleo rotation angle 1
Oleo rotational velocity 1
Revolute 2 Bogie rotation angle 1
Bogie rotational velocity 1

Right Wing Landing Gear Prismatic 2 Oleo stroke 1 1


Oleo velocity 1 1
Revolute 2 Oleo rotation angle 1
Oleo rotational velocity 1
Revolute 2 Bogie rotation angle 1
Bogie rotational velocity 1

Left Body Landing Gear Prismatic 2 Oleo stroke 1


Oleo velocity 1
Revolute 2 Oleo rotation angle 1
Oleo rotational velocity 1
Revolute 2 Bogie rotation angle 1
Bogie rotational velocity 1
Revolute 2 Rear axle rotation angle 1
Rear axle rotational velocity 1
Rotational motion -2 Rear axle rotation angle -1
Rear axle rotational velocity -1

100
101

Table B.1: State names. † These states are held constant.

Component Constraint Degrees Description A320 A380


of freedom State State

Right Body Landing Gear Prismatic 2 Oleo stroke 1


Oleo velocity 1
Revolute 2 Oleo rotation angle 1
Oleo rotational velocity 1
Revolute 2 Bogie rotation angle 1
Bogie rotational velocity 1
Revolute 2 Rear axle rotation angle 1
Rear axle rotational velocity 1
Rotational motion -2 Rear axle rotation angle -1
Rear axle rotational velocity -1

CG None 12 X-position in body-axis† 1 1


Y-position in body-axis† 1 1
Z-position in body-axis 1 1
Pitch angle 1 1
Roll angle 1 1
Yaw angle† 1 1
X-velocity in body-axis 1 1
Y-velocity in body-axis 1 1
Z-velocity in body-axis 1 1
Roll-rate in body axis 1 1
Pitch-rate in body axis 1 1
Yaw-rate in body axis 1 1
Total Number of States 18 38
Appendix C

Kinematic Dimensions

The wheel base and track width data in Table C.1 were obtained from [3, 7, 8, 10, 27]. Table C.2
contains the calculated steering angles that are achieved for different aircraft types, based on
the equations in Section 5.2. Tables C.3 and C.4 contain estimates for the minimum radial
location of the inner gear reference position, from the origin of the centreline, based on the
equations in Section 5.3.

Table C.1. Aircraft dimensions.

AC Type Wheel base Track Width Lm


ln (m) lm (m)
A320 12.64 7.59 0.60
A321 16.91 7.59 0.45
A330-200 22.20 10.69 0.48
A340-600 32.89 10.69 0.33
A380 30.40 14.30 0.47
An-124 23.00 9.00 0.39
C5 22.22 11.42 0.51
B737-900 17.17 6.72 0.39
B747-8 29.67 12.00 0.40
B777-300ER 31.22 11.97 0.38
B787 22.78 10.80 0.47
MD-81 22.10 5.10 0.23
MD-11 24.60 10.70 0.43

102
103

Table C.2. Steering angles for Group V and Group VI Airports.

AC Type Grp V Grp VI


(rn = 45.7m) (rn = 51.0m)
Rn δf δ90 δ135 Rn δf δ90 δ135
A320 3.62 16.06 15.99 16.05 4.03 14.35 14.32 14.35
A321 2.70 21.72 21.30 21.66 3.02 19.36 19.14 19.34
A330-200 2.06 29.06 27.43 28.68 2.30 25.80 24.83 25.61
A340-600 1.39 46.03 38.20 42.80 1.55 40.16 34.92 38.27
A380 1.50 41.70 35.82 39.49 1.68 36.59 32.72 35.33
An-124 1.99 30.22 28.30 29.74 2.22 26.81 25.65 26.57
C5 2.06 29.09 27.45 28.70 2.30 25.83 24.85 25.64
B737-900 2.66 22.07 21.61 22.00 2.97 19.67 19.43 19.65
B747-8 1.54 40.48 35.12 38.53 1.72 35.57 32.06 34.47
B777-300ER 1.46 43.09 36.61 40.58 1.63 37.75 33.46 36.30
B787 2.01 29.90 28.06 29.44 2.24 26.53 25.42 26.30
MD-81 2.07 28.92 27.32 28.54 2.31 25.68 24.72 25.49
MD-11 1.86 32.57 30.01 31.85 2.07 28.84 27.26 28.47

Table C.3. Minimum clearance location of inner-gear reference position at Group V airports.

AC Type θm90min rm90min θm135min rm135min


(deg) (m) (deg) (m)
A320 75.33 40.16 119.33 40.12
A321 72.07 38.82 114.78 38.76
A330-200 69.16 35.06 109.99 34.56
A340-600 65.61 29.89 103.62 27.96
A380 66.26 29.34 104.80 27.77
An-124 68.81 35.54 109.37 34.95
C5 69.15 34.69 109.97 34.18
B737-900 71.90 39.15 114.52 39.08
B747-8 66.46 30.85 105.18 29.39
B777-300ER 66.04 30.10 104.40 28.41
B787 68.90 34.74 109.54 34.18
MD-81 69.21 37.90 110.07 37.41
MD-11 68.15 33.95 108.22 33.17
104 Chapter C. Kinematic Dimensions

Table C.4. Minimum clearance location of inner-gear reference position at Group VI airports.

AC Type θm90min rm90min θm135min rm135min


(deg) (m) (deg) (m)
A320 76.50 45.55 120.54 45.37
A321 73.30 44.45 116.64 44.46
A330-200 70.30 40.87 111.92 40.57
A340-600 66.52 35.89 105.28 34.29
A380 67.21 35.30 106.53 34.05
An-124 69.93 41.37 111.29 41.00
C5 70.29 40.50 111.90 40.20
B737-900 73.13 44.80 116.39 44.80
B747-8 67.43 36.81 106.93 35.65
B777-300ER 66.97 36.07 106.11 34.71
B787 70.03 40.57 111.46 40.21
MD-81 70.35 43.71 112.00 43.42
MD-11 69.23 39.82 110.11 39.28
Appendix D

Dynamical Systems Toolbox, Integrating


AUTO into Matlab

D.1 Introduction

Dynamical systems theory provides a methodology for studying systems of nonlinear ordinary
differential equations (ODEs). A key method is that of bifurcation analysis, where one identi-
fies different ways in which the dynamics of the system can change. In combination with the
numerical technique of continuation, one can perform a nonlinear stability analysis by follow-
ing solutions and detecting their stability changes (bifurcations). The bifurcations can then be
followed in more parameters to identify regions in parameter space that correspond to different
behaviour of the system. See, for example [29] and [62] as entry points to the literature.
The development of a Dynamical Systems Toolbox (DST) within the Matlab environment is a
first attempt to consolidate previous efforts at the University of Bristol to create a user-friendly
environment for engineers, and follows on previous work by Ryan Bedford in the Aerospace
Engineering Department. Other efforts around the world to develop dynamical systems soft-
ware in Matlab exist, such as MATCONT [28], but it appears that this software has not been
widely adopted by the engineering community. This opinion needs to be tested by further
investigation.

D.2 Toolbox Development

We believe that the following aspects are needed to create an environment that will enable the
wider adoption of dynamical systems methods within the engineering community:

1. The software environment needs to be familiar, hence an environment such as Matlab is


ideal, as many engineering students learn Matlab at university, and Matlab is also used
in many companies.

2. The software constructs need to be familiar. Software such as AUTO [19] is widely used,
hence it would make sense to use similar terms and constructs to that of AUTO.

105
106 Chapter D. Dynamical Systems Toolbox, Integrating AUTO into Matlab

3. The software constructs need to be easy to maintain. The structure or Object Oriented
Programming (OOP) features in Matlab can contribute to this end.

4. Detailed examples and an extensive help section are essential.

5. A wide enough user-base is needed. People with experience in a specific software tool
are needed to support new users. AUTO is therefore a good candidate.

6. The run-time requirements need to be balanced with usability. Lower level languages
such as Fortran are more efficient, but the correct programming techniques in Matlab
can make program execution very efficient, especially if vectorization is extensively used
and loops are avoided.

The optimal fit to these needs seems to be achieved by integrating the existing Fortran AUTO
code into Matlab via mex-functions. This allows one to use the speed of a lower level language
with the user-friendly interface of Matlab. The user creates similar files to that of the original
AUTO code, i.e. a constants file and an equations file. Another important aspect of the tool-
box is that engineering tools such as Simulink and SimMechanics can be integrated with the
dynamical systems software. In this way, industrially tested models (both existing and newly
developed) can be used directly in this environment, without the need for converting models
to a format that can be used by stand-alone AUTO. More specifically, AUTO has direct access
to the states of the Simulink/SimMechanics models. Similar output files to that of AUTO are
also generated. An additional feature was added to the fort.7 file, where the outputs from
Simulink output ports are also written alongside the continuation parameters and states.
The toolbox can be downloaded from http://seis.bris.ac.uk/~ec1099.

D.3 Benchmarks

The “ab” demonstration in AUTO was used as a benchmark problem where a comparison is
made between the run-times of different sub-problems (i.e. stationary or periodic solutions),
as well as different mex-integration schemes between Fortran and Matlab. The stationary and
locii solutions of the DST run approximately twice as slowly as the original Fortran code, while
limit-cycle solutions take significantly longer. Figure D.1 contains comparisons for limit-cycle
continuations. Different interface schemes between Fortran and Matlab are depicted, as well as
additional features such as error checking and the ability to write additional outputs from the
Simulink output ports into the AUTO output files. The first bar indicates the original AUTO07P
code. The second to the fifth bars all contain DST benchmarks where the function file is called
via a mexCallMatlab command. A mexEvalString command was originally used and
was proven to be very inefficient, due to extra function calls being invoked. In bars two to four
all the functions that are usually contained in the AUTO function definition file are split into
separate Matlab functions, and then directly called from the mex-file. It can be seen that error
checking (second bar) and additional outputs (bar four) do not make a significant difference to
the execution time. The last case (fifth bar) is where all the functions are contained in one file,
and then called via a ‘case’ selection at the entry into the file. It can therefore be concluded that
the effect of additional functionality such as error checking, and additional outputs, is far less
D.4. Future Directions 107

Demo ‘ab’ bifurcation diagram

Periodic solutions

Figure D.1. Limit-cycle benchmark comparisons for the ‘ab’ demo in AUTO.

than the effect of the data transfer across the function interface. It therefore stands to reason
that any future work should focus on other interface schemes between Fortran and Matlab,
such as the use of function handles, or any other options that might increase the rate at which
function calls can be made.

D.4 Future Directions

More widespread use of the Dynamical Systems Toolbox will be promoted by providing doc-
umentation and reference material that is easy to use, while concrete examples will act as
additional training material for the user. We have combined most of the user manual of AUTO
into the toolbox, which is integrated into the Matlab help environment. Future versions of the
toolbox will contain our own examples related to aerospace applications [13, 54, 64]. The Dy-
namical Systems Toolbox therefore feels like any other toolbox that has been developed for
Matlab, where the user can select the toolbox from the menu, with the accompanying help and
search functionality. We have also started to develop components with the new OOP capability
in Matlab, which we feel will enhance the pace at which new applications will be developed
in the future. The nature of the problems tend to be “massively parallel”, which lends itself
towards use on GPU’s. This will allow for the construction of a complete bifurcation surface,
across all steering angles and velocities, in a matter of two or three minutes.

Potrebbero piacerti anche