26.12.2012 Views

Model based control of a flight simulator motion system - DCSC

Model based control of a flight simulator motion system - DCSC

Model based control of a flight simulator motion system - DCSC

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong><br />

<strong>system</strong>


<strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong><br />

<strong>system</strong><br />

PROEFSCHRIFT<br />

ter verkrijging van de graad van doctor<br />

aan de Technische Universiteit Delft,<br />

op gezag van de Rector Magnificus pr<strong>of</strong>. ir K.F. Wakker<br />

voorzitter van het College voor Promoties,<br />

in het openbaar te verdedigen op maandag 10 december 2001 om 16.00 uur<br />

door<br />

Sjirk Holger KOEKEBAKKER<br />

werktuigkundig ingenieur<br />

geboren te Ermelo


Dit proefschrift is goedgekeurd door de promotor:<br />

Pr<strong>of</strong>. ir O.H. Bosgra<br />

Samenstelling promotiecommissie:<br />

Rector Magnificus voorzitter<br />

Pr<strong>of</strong>. ir O.H. Bosgra Technische Universiteit Delft, promotor<br />

Dr ir A.J.J. Van der Weiden Technische Universiteit Delft, toegevoegd promotor<br />

Pr<strong>of</strong>. dr ir M. Steinbuch Technische Universiteit Eindhoven<br />

Pr<strong>of</strong>. dr ir J.A. Mulder Technische Universiteit Delft<br />

Pr<strong>of</strong>. dr ir P.M.J. van den H<strong>of</strong> Technische Universiteit Delft<br />

Pr<strong>of</strong>. dr ir J.H. de Leeuw University <strong>of</strong> Toronto<br />

Ir P.C. Teerhuis Technische Universiteit Delft<br />

Ir P.C. Teerhuis heeft als begeleider in belangrijke mate aan de<br />

totstandkoming van het proefschrift bijgedragen.<br />

Keywords: model <strong>based</strong> <strong>control</strong>, <strong>flight</strong> simulation, parallel <strong>motion</strong> <strong>system</strong>s<br />

ISBN 90-370-0194-7<br />

Copyright c 2001 by S.H. Koekebakker<br />

All rights reserved. No part <strong>of</strong> the material protected by this copyright notice may be reproduced or<br />

utilized in any form or by any means, electronic or mechanical, including photocopying, recording or<br />

by any information storage and retrieval <strong>system</strong>, without written permission from the copyright owner.<br />

Printed in the Netherlands by Ponsen & Looijen b.v.


To faith<br />

We’re on a road to nowhere<br />

Come on inside<br />

Takin’ that ride to nowhere<br />

We’ll take that ride<br />

Talking Heads, 1985<br />

Running is no sport<br />

but a way <strong>of</strong> travelling<br />

displacing body and mind<br />

Vrij naar Jan Knippenberg


Voorwoord<br />

In de ruwweg zeven jaren, maanden, dagen en uren tussen de start van dit onderzoek en het<br />

moment waarop het verdedigd zal worden is veel gebeurd. Beginnend vanuit een kamertje<br />

in een studentenhuis in Delft kon ik me weinig voorstellen bij de grote, relatief eenzaam<br />

verrichte inspanningen waar kond van werd gedaan in het voorwoord van de verschillende<br />

proefschriften. Onderzoek doen is leuk en zeker in het kader van een multidisciplinair<br />

project werk je daarbij met veel mensen samen. Pas nadat het onderzoek na vier jaar<br />

afgerond maar nog niet helemaal opgeschreven was en naast een aantrekkelijke voltijds<br />

baan en een fantastisch gezin met inmiddels twee dochters vanuit Venlo het werk in groter<br />

geheel vastgelegd moest worden, kon ik mij vinden in vele van die eerder gelezen frasen.<br />

Alhoewel een groot aantal mensen je op een <strong>of</strong> andere manier bijstaan tijdens de gehele periode,<br />

moet het schrijfwerk toch grotendeels alleen werkend achter de computer gebeuren.<br />

Hier een punt achter zettend is eindelijk de tijd ook gekomen om al deze mensen uitdrukkelijk<br />

te bedanken voor de hulp die ze me op verschillende wijze hebben geboden. Een<br />

aantal mensen zal ik alleen als groep noemen gezien de vele veranderingen die zich over<br />

deze lange periode hebben voorgedaan en het risico daarbij individuele namen over te slaan.<br />

Ten eerste wil ik mijn promotor Okko Bosgra bedanken die me in staat heeft gesteld<br />

mijn promotieonderzoek te verrichten binnen zijn vakgroep. Vervolgens bedank ik mijn<br />

directe begeleiders, Ton van der Weiden met zijn directheid en politiek gevoel voor tact<br />

waar nodig in dit project en Piet Teerhuis met zijn aanstekelijke liefde voor de techniek en<br />

in het bijzonder de hydraulische <strong>system</strong>en.<br />

Zoals in de referenties terug te vinden is, heeft een grote groep studenten het laatste (anderhalf<br />

<strong>of</strong> meer) jaar van de ingenieursopleiding een afstudeerproject in het kader van mijn<br />

onderzoek kunnen verrichten. Dit heb ik als zeer prettig ervaren. Zeker op het moment dat<br />

er gelijktijdig vijf, zes enthousiaste mensen de ’Simona <strong>motion</strong> <strong>control</strong> room’ bij werktuigbouwkunde<br />

bevolkten en we door slim tactisch spel een (audiovisio)hardloopwedstrijd konden<br />

winnen zonder als eerste aan te komen. Boris Rijnten, Vasken der Kevorkian, Sander<br />

Bettendorf, Philippe Piatkiewitz, Jan van Hulzen, Maris Franken, Riad Al-Saidi, Etienne<br />

van Zuijlen en Arne Scheffer bedankt voor jullie mede- en vooral samenwerking.<br />

Alle groeps- en ex-groepsleden van de vakgroep Systeem- en regeltechniek wil ik bedanken<br />

voor de collegialiteit, ook in de periode dat ik al uit Delft weg was. Met name door<br />

de combinatie van een grote groep promovendi, de ervaren staf en de rechtdoorzee support<br />

groep was er sprake van een levendige en productieve sectie. Marco Dettori bracht daar<br />

als langstzittende kamergenoot nog wat Italiaanse cultuur in die gecombineerd met Carsten<br />

Scherers internationale en vooral ook wiskundige achtergrond tot prettig pittig smakende<br />

lunchdiscussies leidden. Thomas de Hoog wil ik ook speciaal noemen vanwege het feit dat<br />

vii


viii Voorwoord<br />

ik de template van zijn proefschrift en allerhande LateX-advies goed heb kunnen gebruiken.<br />

Door het interfacultaire karakter van het <strong>simulator</strong>project is er interactie met veel mensen<br />

die ik als zeer leerzaam heb ervaren. Ik bedank alle mensen betrokken bij Simona, vooral<br />

ook door de eigenlijk nooit ontbrekende en ook zeker noodzakelijke passie voor het maken<br />

van een technisch complex apparaat waarbij altijd vele hobbels genomen moeten worden.<br />

Ik wil een aantal mensen hiervan met name noemen waarmee ik door hun directe betrokkenheid<br />

bij mijn onderzoek nauw samengewerkt heb. Sunjoo Advani, die met zijn enthousiasme<br />

het Simonaproject als projectleider lange tijd over pieken en dalen getrokken<br />

heeft en ook tot nu nog grote betrokkenheid bij dit onderzoek heeft getoond. Gert van<br />

Schothorst, die, twee jaar eerder als werktuigkundig promovendus bij het <strong>simulator</strong>project<br />

begonnen, veel werk op het gebied van de modelvorming rond hydraulische <strong>system</strong>en heeft<br />

verricht waarop ik met mechanica en regeling verder kon bouwen. Met Peter Valk kon ik,<br />

gegeven ook de gelijkgestemdheid op sportief gebied, eindeloze discussies voeren. Rens<br />

de Keijzer, Kees Slinkman, Rolf van Overbeek, Fred den Hoedt, John Dukker en Ad van<br />

der Geest brachten allen vanuit onze faculteit een praktisch technische insteek waar ik nog<br />

steeds niet aan kan tippen. Ook de wizardous meet- en communicatiekastjes van de prettig<br />

nuchtere Henk Huisman moet ik niet vergeten.<br />

Buiten het werk brachten in de eerste jaren vooral de gezamenlijk gelopen ettelijk<br />

honderden trainingskilometers, tientallen etentjes en (estafette)wedstrijden onder andere in<br />

Londen, Parijs, Barcelona en een tweede keer ook weer tactisch winnend rond het IJsselmeer<br />

met het losse en toch ook weer vaste groepje jonge oude Delvers de nodige afleiding<br />

en zeker ook vriendschap. Daarnaast leefden ook veel andere vrienden, kennissen en<br />

familie mee. Allen bedankt!<br />

Alhoewel de frequentie verlaagd is houdt nu een groep enthousiaste collega’s de sportiviteit<br />

en de felle en gelukkig lekker oeverloze en daarmee gevarieerde discussies tijdens de<br />

gezamenlijke etentjes erin. Ik wil graag hierbij ook mijn ho<strong>of</strong>d bij de researchafdeling van<br />

Océ, Jo Geraedts, bedanken voor zijn steun ten aanzien van mijn werk voor dit onderzoek<br />

en dan met name zijn motivatie om er vooral op een goede manier een punt achter te zetten.<br />

Tenslotte kan ik van hieruit nooit mijn lieve Els, Hanneke en Ingrid voldoende bedanken<br />

voor het geduld dat ze met me en het vertrouwen dat ze in me hebben gehad. Ik zal van<br />

’pappa z’n werk’ met vier PC’s en meer dan twintig meter stapels papieren archief op twaalf<br />

vierkante meter maar (g?)een museum maken.<br />

Sjirk Koekebakker<br />

Venlo, oktober 2001.


Summary<br />

The extremely high safety requirements in air transportation require a proper understanding<br />

<strong>of</strong> human pilot behaviour under extreme weather conditions. Advanced research in this<br />

area must utilize <strong>flight</strong> <strong>simulator</strong> equipment that is able to reproduce critical <strong>flight</strong> conditions<br />

with great fidelity. The turbulence and wind shear phenomena in the lower part <strong>of</strong><br />

the atmosphere contain a velocity spectrum that induces large forces on the aircraft over a<br />

wide frequency range. Thus, <strong>motion</strong> reproduction and generation <strong>of</strong> desired accelerations<br />

in a high quality <strong>flight</strong> <strong>simulator</strong> constitutes a complicated <strong>control</strong> task that must exploit<br />

the capabilities <strong>of</strong> an available actuation <strong>system</strong> to its limits. Predominately, these <strong>system</strong>s<br />

are comprised <strong>of</strong> six hydraulically actuated linear servo motors, similar to many robotic<br />

manipulators. The <strong>control</strong> <strong>of</strong> modern robotic devices employs advanced model <strong>based</strong> <strong>control</strong><br />

strategies that allow higher performance to be achieved than less structured approaches,<br />

and can provide more insight into the <strong>system</strong> limitations. In <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>control</strong>,<br />

application <strong>of</strong> such methods has been almost absent. One <strong>of</strong> the reasons might be that<br />

the <strong>motion</strong> <strong>system</strong>s have a degree <strong>of</strong> complexity, which does not allow exact modelling or<br />

detailed models as part <strong>of</strong> the <strong>control</strong>ler. Application has required an intermediate step <strong>of</strong><br />

extracting a reasonably detailed model, which described the most relevant dynamic characteristics.<br />

The central problem <strong>of</strong> this research was to investigate which relevant <strong>system</strong> knowledge<br />

should be used in a model <strong>based</strong> <strong>control</strong> strategy, and to quantify the extent to which this improves<br />

the <strong>control</strong>led dynamics <strong>of</strong> the <strong>simulator</strong>. First, the full design process was defined,<br />

structured and evaluated. This began with modelling and analysis, followed by <strong>control</strong> synthesis<br />

and, finally, testing the <strong>system</strong>. Many solutions and procedures <strong>of</strong> the steps taken<br />

were given in literature, but an integral approach, having the specific properties and requirements<br />

<strong>of</strong> <strong>flight</strong> simulation <strong>motion</strong> generation in mind, was still lacking. In this research,<br />

integratibility and applicability were the main arguments in choosing the most suitable alternative<br />

or, if necessary, newly proposed variant, <strong>of</strong> each step in the <strong>system</strong>/<strong>control</strong> design<br />

procedure.<br />

First, a relevant model structure for analysis and <strong>control</strong> through physical modelling was<br />

derived. Next, the structure was evaluated by experiments. <strong>Model</strong> parameters were identified<br />

and boundaries were provided to the extent for which the model validity holds. Then,<br />

the most appropriate model <strong>based</strong> robot <strong>control</strong> strategy was chosen and, given the particular<br />

<strong>system</strong> properties and performance requirements, modified for this application. The next<br />

step was actual application within the <strong>simulator</strong> environment, after which the testing procedure<br />

was defined in order to quantify the performance with respect to the requirements, and<br />

to conventional <strong>system</strong>s.<br />

ix


x Summary<br />

Structural insight into the <strong>system</strong> kinematics and dynamics was required in order to<br />

maintain at least the level <strong>of</strong> robustness <strong>of</strong> conventional <strong>control</strong> strategies, and simultaneously<br />

increase <strong>motion</strong> performance and predictability. The physical modelling began with<br />

an analysis <strong>of</strong> the <strong>system</strong> kinematics and, as a contribution <strong>of</strong> this thesis, a method was proposed<br />

to exclude singular points from the operational workspace and, subsequently, safely<br />

and efficiently determine platform pose from the actuator positions in real time. Calibration<br />

<strong>of</strong> the physical <strong>system</strong> improved further its positioning accuracy by an order <strong>of</strong> magnitude.<br />

The dynamics <strong>of</strong> the Stewart Platform <strong>of</strong> this <strong>simulator</strong> were expressed by a set <strong>of</strong> explicit<br />

differential equations using platform coordinates. It was shown in theory and practice<br />

that the rigid body modes resulting from interaction <strong>of</strong> hydraulics and mechanics form the<br />

most relevant <strong>motion</strong> <strong>system</strong> dynamics in <strong>simulator</strong> applications. The platform to actuator<br />

coordinate jacobian is a central operator here. The research showed also how the parallel<br />

actuation can be viewed from a new set <strong>of</strong> coordinates as six nearly independent hydraulic<br />

<strong>system</strong>s driving a single mass.<br />

The applied model <strong>based</strong> <strong>control</strong>ler attained several goals, which could not be implemented<br />

conventionally. First, it directs the required forces for the desired accelerations<br />

along the appropriate vectors experiencing largely different masses. Then, the feedback<br />

paths <strong>of</strong> the positional errors also use these decoupled directions. Finally, the interaction<br />

between hydraulics and mechanics is minimised over the feed forward path primarily by<br />

applying compensation <strong>of</strong> the required oil flow given the desired velocity. Implementation<br />

in practice showed an even higher bandwidth, less peaking, and a more equalised response<br />

<strong>of</strong> the model <strong>based</strong> <strong>control</strong>ler over each degree <strong>of</strong> freedom compared with a conventional<br />

strategy. Even more can be gained since model <strong>based</strong> <strong>control</strong> can more effectively use<br />

predictive information from the vehicle simulation model.<br />

Performance limitations to this technique can be found at frequencies where parasitic<br />

effects such as structural flexibility, and hydraulic valve and transmission line dynamics<br />

become significant, or when the <strong>system</strong> no longer behaves as being rigidly attached to the<br />

floor. However, through the application <strong>of</strong> a light weight and rigid design <strong>of</strong> the moving<br />

platform, and with the <strong>control</strong> strategy proposed in this thesis, the <strong>system</strong> bandwidth can<br />

be raised to a sufficiently high level, thereby enabling a wider range <strong>of</strong> <strong>motion</strong> <strong>based</strong> <strong>flight</strong><br />

simulation research to be executed than heret<strong>of</strong>ore.


Contents<br />

Voorwoord vii<br />

Summary ix<br />

1 Introduction 1<br />

1.1 Flight simulation . .............................. 1<br />

1.1.1 The art <strong>of</strong> simulation . . ....................... 2<br />

1.1.2 Mechanism <strong>of</strong> <strong>motion</strong> perception . . . ............... 4<br />

1.1.3 Historical perspective . . ....................... 8<br />

1.1.4 Motion simulation . . . ....................... 9<br />

1.1.5 SIMONA project . . . . ....................... 12<br />

1.2 Motion <strong>control</strong> . . .............................. 17<br />

1.2.1 Motion <strong>control</strong> structure ....................... 17<br />

1.2.2 Motion <strong>control</strong> specifications . . ................... 19<br />

1.2.3 Conventional <strong>simulator</strong> <strong>motion</strong> <strong>control</strong> ............... 21<br />

1.2.4 Robot <strong>control</strong> . . . . . . ....................... 22<br />

1.3 Problem statement .............................. 25<br />

1.3.1 General problem statement . . . ................... 25<br />

1.3.2 Structuring the problem statement;<br />

approach in research . . ....................... 25<br />

1.4 Outline <strong>of</strong> the thesis .............................. 28<br />

2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s 31<br />

2.1 Parallel <strong>motion</strong> <strong>system</strong>s . . . . . ....................... 32<br />

2.2 Kinematics .................................. 33<br />

2.2.1 Notation and definitions ....................... 33<br />

2.2.2 Calculating velocity and acceleration by differentiation . . . .... 34<br />

2.2.3 Parametrising orientation by euler angles . . . . . . ........ 36<br />

2.2.4 Parametrising orientation by euler parameters . . . . ........ 38<br />

2.2.5 Jacobian matrices . . . . ....................... 40<br />

2.2.6 Parallel, serial and singular configurations . . . . . . ........ 41<br />

2.2.7 Stewart platform definitions and assumptions . . . . ........ 43<br />

2.2.8 Stewart platform kinematics . . ................... 44<br />

xi


xii Contents<br />

2.2.9 Interpretation and use <strong>of</strong> the jacobian matrix, J l�x .......... 45<br />

2.2.10 Velocity and acceleration <strong>of</strong> the actuator joints . . . ........ 47<br />

2.2.11 The Simona <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> kinematics . . . .... 49<br />

2.3 Analysis <strong>of</strong> the Stewart platform kinematics . . ............... 50<br />

2.3.1 Convergence NR-iteration . . . ................... 53<br />

2.3.2 NR-convergence Stewart platform . . . ............... 54<br />

2.3.3 Lipschitz condition on jacobian ................... 56<br />

2.3.4 Exclusion <strong>of</strong> singular points . . ................... 58<br />

2.3.5 Sufficient update frequency . . ................... 59<br />

2.4 <strong>Model</strong>ling the mechanical part <strong>of</strong> the <strong>system</strong> dynamics . . . ........ 60<br />

2.4.1 General theory in modelling the dynamics <strong>of</strong> mechanical <strong>system</strong>s . 60<br />

2.4.2 Example in modelling using Kane’s method . . . . . ........ 64<br />

2.4.3 Stewart platform dynamics . . . ................... 66<br />

2.4.4 Basic Stewart platform mechanical model dynamics ........ 66<br />

2.4.5 Influence <strong>of</strong> the actuator inertial forces ............... 67<br />

2.4.6 The Stewart platform model including actuator inertia . . . .... 69<br />

2.4.7 Parasitic mechanical aspects: modelled as linear flexibility . .... 70<br />

2.5 Chapter Resume . . .............................. 73<br />

3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s 75<br />

3.1 The basic structure <strong>of</strong> hydraulic actuators . . . ............... 76<br />

3.1.1 Leakage . . .............................. 78<br />

3.1.2 Basic hydraulically driven mechanical <strong>system</strong> model ........ 79<br />

3.2 Extensions . .................................. 80<br />

3.2.1 Servo valve .............................. 80<br />

3.2.2 Transmission lines . . . ....................... 85<br />

3.3 Integrated <strong>system</strong> . .............................. 89<br />

3.3.1 Passive input/output pairs in <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s .... 90<br />

3.3.2 The 1-d.o.f. hydraulically driven mechanical <strong>system</strong> ........ 94<br />

3.3.3 Additional mechanical modes . ................... 97<br />

3.4 Hydraulically driven Stewart platform . ................... 101<br />

3.4.1 Basic model structure hydraulically driven <strong>system</strong>s . ........ 102<br />

3.4.2 Dynamical and kinematical properties <strong>of</strong> the SRS-<strong>motion</strong> <strong>system</strong>;<br />

Design aspects . . . . . ....................... 104<br />

3.4.3 Actuator inertial effects . ....................... 110<br />

3.4.4 Analyzing the SRS hydraulically driven <strong>motion</strong> <strong>system</strong> model . . . 112<br />

3.4.5 Connection <strong>of</strong> a flexible foundation to the SRS <strong>system</strong> model . . . 116<br />

3.5 Chapter Resume . . .............................. 118<br />

4 Parameter identification and model validation 119<br />

4.1 Calibration .................................. 120<br />

4.1.1 Calibrating the Stewart platform ................... 121


Contents xiii<br />

4.1.2 Redundant measurement <strong>of</strong> the platform pose . . . . ........ 123<br />

4.1.3 Stewart platform pose measurement in practice . . . ........ 126<br />

4.1.4 Identification <strong>of</strong> the kinematical parameters . . . . . ........ 128<br />

4.1.5 Results in calibrating the SRS <strong>motion</strong> <strong>system</strong> . . . . ........ 131<br />

4.2 Stewart platform model parameter identification ............... 133<br />

4.2.1 Gravitational force determination . . . ............... 133<br />

4.2.2 Identification <strong>of</strong> the inertial properties . ............... 136<br />

4.3 Frequency response model validation . . ................... 141<br />

4.3.1 Frequency response dummy platform . ............... 142<br />

4.3.2 Additional dynamics into the higher frequency area . ........ 147<br />

4.4 Chapter Resume . . .............................. 161<br />

5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> 163<br />

5.1 Introduction .................................. 163<br />

5.2 Another look at the <strong>control</strong> problem . . ................... 164<br />

5.3 Control Strategy . . .............................. 166<br />

5.4 Inner loop pressure <strong>control</strong> . . . ....................... 168<br />

5.5 Multivariable feedback linearisation . . ................... 173<br />

5.5.1 Feedback linearising robotic manipulators . . . . . . ........ 173<br />

5.5.2 Feedback linearisation <strong>of</strong> a Stewart platform . . . . ........ 174<br />

5.5.3 Implicit state measurement requirements . . . . . . ........ 176<br />

5.5.4 Outer loop <strong>control</strong> . . . ....................... 177<br />

5.6 Reference model <strong>based</strong> feed forward . . ................... 178<br />

5.6.1 Reference model <strong>based</strong> predictors . . . ............... 179<br />

5.6.2 Construction <strong>of</strong> the reference model <strong>based</strong> <strong>control</strong>ler ........ 181<br />

5.7 Implementational issues . . . . . ....................... 185<br />

5.8 Performance quantification . . . ....................... 185<br />

5.8.1 Motion <strong>system</strong> evaluation methods and requirements ........ 187<br />

5.8.2 New test procedure . . . ....................... 189<br />

5.9 Experimentally evaluating performance ................... 191<br />

5.9.1 Experimental set up . . . ....................... 191<br />

5.9.2 Characteristics <strong>of</strong> the Simona <strong>motion</strong> <strong>system</strong> . . . . ........ 193<br />

5.10 Chapter Resume . . .............................. 200<br />

6 Review and discussion on the results 203<br />

6.1 Flight simulation . .............................. 203<br />

6.2 Motion <strong>system</strong> specifications . . ....................... 204<br />

6.3 Theoretical modelling . . . . . . ....................... 205<br />

6.4 Experimental modelling and validation . ................... 206<br />

6.5 Control strategy and evaluation . ....................... 208<br />

7 Conclusion and recommendations 211


xiv Contents<br />

7.1 Conclusion .................................. 211<br />

7.2 Recommendations . .............................. 213<br />

A Frequency domain measurements 215<br />

B Derivation <strong>of</strong> actuator inertial properties 227<br />

Bibliography 231<br />

Glossary <strong>of</strong> symbols 241<br />

Samenvatting en CV 247


Chapter 1<br />

Introduction<br />

”As man travels, he vibrates. For as he travels, he generally experiences accelerations<br />

not perfectly uniform in magnitude nor direction, and these he feels as varying forces or<br />

vibrations. Travelling faster, these vibrations become more severe, until he learns new<br />

ways to reduce them. But the pioneers will vibrate considerably.” [23] These phrases by<br />

Clark, originally used to motivate research into the tolerable limits <strong>of</strong> <strong>motion</strong> cues for human<br />

beings in the early development <strong>of</strong> the U.S. space program, exactly point at the aspect <strong>of</strong><br />

piloted simulation, which is the subject <strong>of</strong> this research. The <strong>motion</strong> cues, which are a<br />

consequence <strong>of</strong> travelling, can be stimulated without actually travelling, thereby enhancing<br />

perceived realism in a <strong>simulator</strong>. Further, since proper <strong>control</strong> <strong>of</strong> a vehicle requires sufficient<br />

reduction <strong>of</strong> vibration, the actual sensation <strong>of</strong> these resonances forms an essential part <strong>of</strong><br />

basic skill training.<br />

The tool to achieve this, is a <strong>motion</strong> <strong>system</strong>. By automatic <strong>control</strong>, the dynamics <strong>of</strong><br />

the <strong>motion</strong> <strong>system</strong> will have to be transformed into the dynamics <strong>of</strong> the vehicle with its<br />

environment. This requires knowledge <strong>of</strong> both <strong>system</strong>s. With the developments in computer<br />

technology, more and more information can be directly incorporated in the <strong>control</strong>ler. Given<br />

feasible trajectories generated by a reference generating <strong>system</strong> <strong>of</strong> e.g. an aircraft model in<br />

our case, this research considers the use <strong>of</strong> <strong>motion</strong> <strong>system</strong> knowledge in a model <strong>based</strong><br />

<strong>control</strong>ler.<br />

In this chapter, first an overview <strong>of</strong> the most important aspects <strong>of</strong> <strong>flight</strong> simulation will be<br />

given in Section 1.1. Then, the specific requirements and approaches towards <strong>motion</strong> <strong>control</strong><br />

will be discussed in Section 1.2. This will lead to the problem statement and approach <strong>of</strong><br />

this research in Section 1.3. Finally, the outline <strong>of</strong> this thesis will be given in Section 1.4.<br />

1.1 Flight simulation<br />

First, the different aspects <strong>of</strong> the complex <strong>system</strong> <strong>of</strong> <strong>flight</strong> simulation will be discussed.<br />

With an emphasis on the <strong>motion</strong> aspect, the evolution <strong>of</strong> these <strong>system</strong>s over the years will<br />

be sketched. Then, the Simona project centred around the development <strong>of</strong> a modern research<br />

<strong>simulator</strong>, in which this investigation was embedded, will be introduced. Finally, the specific<br />

features <strong>of</strong> the six degrees-<strong>of</strong>-freedom Stewart-Gough Platform <strong>motion</strong> <strong>system</strong>, nowadays<br />

generally used in <strong>flight</strong> simulation, will be stated.<br />

1


2 1 Introduction<br />

1.1.1 The art <strong>of</strong> simulation<br />

The fact that simulation is <strong>of</strong>ten still considered an art instead <strong>of</strong> a science [25],[131], reflects<br />

the fact that this complex mechanism <strong>of</strong> creating a virtual environment is not fully<br />

understood yet. The <strong>flight</strong> <strong>simulator</strong> or more specifically piloted simulation is, however,<br />

widely used nowadays, especially in the area <strong>of</strong> aerospace and astronautics. Following Advani<br />

[7], several areas <strong>of</strong> application can be recognised.<br />

Area <strong>of</strong> application<br />

Engineering <strong>simulator</strong>s are used to evaluate the characteristics <strong>of</strong> a vehicle. In the development<br />

<strong>of</strong> a new aeroplane, the <strong>simulator</strong> is currently intensively used during the whole<br />

process <strong>of</strong> design, due to their ability to predict problems and support <strong>flight</strong> clearance. With<br />

simulation, large reduction in <strong>flight</strong> testing could be achieved, which saved costs and allowed<br />

earlier certification [131]. Most engineering <strong>simulator</strong>s are fixed base i.e. without<br />

inertial <strong>motion</strong> since <strong>motion</strong> cues are not always important and necessary.<br />

However, the <strong>motion</strong> <strong>system</strong> is essential in predicting handling qualities, e.g. a pilot<br />

induced oscillation (PIO) due the dynamics <strong>of</strong> the vehicle, in a proper way [121]. Low<br />

frequent modes <strong>of</strong> the aircraft can be damped more appropriately with <strong>motion</strong>, while vibrations<br />

at higher frequencies amplified by the pilot in the loop, e.g. <strong>of</strong>ten modes due to<br />

structural aeroelasticity, can usually only be detected if the <strong>control</strong>led <strong>motion</strong> <strong>system</strong> leaves<br />

these modes unaltered.<br />

Training <strong>simulator</strong>s are used to train pilots both in the procedural (e.g. <strong>flight</strong> management)<br />

as well as the basic skill tasks (e.g. manual <strong>control</strong>). In military applications the procedural<br />

training is <strong>of</strong>ten done in fixed base <strong>simulator</strong>s. The basic skill training is mostly performed<br />

in-<strong>flight</strong> since the discrepancy between the <strong>simulator</strong> and aircraft is considered too large in<br />

extreme manoeuvres and <strong>of</strong> course actual <strong>flight</strong> training is much more exciting. In commercial<br />

airlines, the economic attractiveness <strong>of</strong> <strong>motion</strong> base <strong>simulator</strong>s is preferred over<br />

actual <strong>flight</strong> time and many hours are spent in training <strong>simulator</strong>s. Pr<strong>of</strong>iciency checks and<br />

re-currency training i.e. skill examinations and transfer to another type <strong>of</strong> plane, are allowed<br />

by the regulating authorities without actually flying for pilots with over 500 hours experience<br />

on an aircraft <strong>of</strong> the same group or 1500 (Europe) to 2500 hours (USA) grand total<br />

[25]. There is a tendency to further reduce this prescribed number <strong>of</strong> hours [143], which<br />

allow this so-called Zero-Flight-Time (ZFT) training, but exact knowledge <strong>of</strong> the required<br />

simulation fidelity is still lacking.<br />

Research <strong>simulator</strong>s are <strong>of</strong>ten applied for fundamental investigation into pilot/vehicle interaction<br />

and human perception research. These kind <strong>of</strong> <strong>simulator</strong>s most <strong>of</strong>ten require the<br />

highest level <strong>of</strong> performance w.r.t. the <strong>motion</strong> <strong>system</strong> as the resulting sensory input is essential<br />

in the manual <strong>control</strong> task [47].<br />

Apart from this fundamental scientific work, the <strong>simulator</strong> has also proved valuable<br />

for accident investigations [131]. Research <strong>simulator</strong>s have been used to study the circum-


1.1 Flight simulation 3<br />

stances <strong>of</strong> incidents due to clear air turbulence and other types <strong>of</strong> severe turbulence like gust,<br />

and come up with more beneficial strategies in <strong>control</strong> and better operational procedures.<br />

In these experiments, the <strong>motion</strong> <strong>system</strong> had to be capable <strong>of</strong> producing vibrations which<br />

made the pilot unable to read the instruments.<br />

Advantages <strong>of</strong> <strong>simulator</strong>s<br />

Reviewing the area <strong>of</strong> application reveals the various specific qualities <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong>.<br />

Over the years it has proved to have several advantages as compared to flying an actual<br />

aircraft.<br />

-Cost effectiveness. Both with respect to initial investment as to operational costs, the <strong>flight</strong><br />

<strong>simulator</strong>, though still requiring a considerable amount <strong>of</strong> money (<strong>of</strong>ten over 10M$<br />

and about 1k$/h), is still an order <strong>of</strong> a magnitude less expensive than most public<br />

transportation aircrafts. Further, the effectiveness <strong>of</strong> training and testing is in many<br />

cases even better.<br />

-Safety requirements. The <strong>simulator</strong> is a safe tool to test and train in conditions, which<br />

are dangerous in actual <strong>flight</strong>. Training <strong>simulator</strong>s can safely be used to show pilots<br />

the best operating procedures in unusual circumstances. Inexperienced pilots can<br />

be taught the bulk <strong>of</strong> basic skills and procedures avoiding the risk <strong>of</strong> putting them<br />

into the air. Many potential <strong>flight</strong> design errors can be detected in advance. Further,<br />

the fact that accidents can be analysed in detail with a <strong>simulator</strong>, enhances <strong>flight</strong><br />

safety. An indirect safety aspect is the cost effectiveness enabling a thorough and still<br />

economically feasible training program.<br />

-Environmental considerations. Every hour spent in a <strong>simulator</strong> instead <strong>of</strong> an aircraft,<br />

directly saves a considerable amount <strong>of</strong> kerosine burnt in the air. Indirectly, environmentally<br />

less harmful <strong>flight</strong> operation can be tested and trained in a <strong>simulator</strong>.<br />

Since the objective is not to duplicate, but to simulate an environment which creates the<br />

best and most effective training and testing results [25],[117], the <strong>flight</strong> <strong>simulator</strong> should<br />

also be considered as a training or testing tool in its own right, rather than as a substitute for<br />

the aircraft [143]. In this respect the <strong>simulator</strong> is <strong>of</strong>ten more effective.<br />

Training can be split up in parts and these modules can be taught separately, in a step by<br />

step approach. The <strong>simulator</strong> provides an objective tool to evaluate different subjects under<br />

equivalent conditions as the repeatability <strong>of</strong> the <strong>system</strong> is much higher. A <strong>simulator</strong> is very<br />

flexible in performing specific tests e.g. under desired weather conditions or at a certain<br />

airport.<br />

Flight <strong>simulator</strong> components<br />

In piloted simulation, a complex environment is created, which has to reflect the impression<br />

<strong>of</strong> travelling in the actual vehicle with an emphasis on the perceived realism <strong>of</strong> the task to<br />

be investigated or learned. This environment can be divided into several sub<strong>system</strong>s:<br />

-Interior. First <strong>of</strong> all the static environment <strong>of</strong> the pilot has preferably the look <strong>of</strong> the interior<br />

<strong>of</strong> the vehicle cockpit.


4 1 Introduction<br />

-Visual <strong>system</strong>. The exterior out-<strong>of</strong>-the-window sight is simulated with an image generating<br />

<strong>system</strong> in case <strong>of</strong> civil aircraft simulation <strong>of</strong>ten projected via a wide angle mirror<br />

attached in front <strong>of</strong> the <strong>simulator</strong> cockpit.<br />

-Instruments. The instruments in the cockpit give information about the state <strong>of</strong> the simulated<br />

vehicle and should actively be changed accordingly.<br />

-Motion <strong>system</strong>. The required inertial <strong>motion</strong>, the actual accelerations <strong>of</strong> the <strong>system</strong>, can<br />

be generated to some extent by actively moving the <strong>system</strong>.<br />

-Control loading. The pilot <strong>control</strong>s to the vehicle <strong>of</strong>ten responds to the given input by<br />

feedback <strong>of</strong> forces which reflect some <strong>of</strong> the external forces on the vehicle e.g. the<br />

aerodynamic force on the rudder, etc. These forces have to be generated artificially<br />

and in this case called <strong>control</strong> loading in a <strong>simulator</strong> and vehicles, where this has been<br />

decoupled mechanically (fly-by-wire).<br />

-Audio <strong>system</strong>. The generation <strong>of</strong> the sound a vehicle together with its environment produces<br />

adds to the perceived realism in the <strong>simulator</strong> and, above this, is used in the<br />

<strong>control</strong> task (gas throttle). In some cases, the audio <strong>system</strong> can also mask the parasitic<br />

noise produced by other sub <strong>system</strong>s [10].<br />

Apart from the static interior, all these sub<strong>system</strong>s add to the <strong>motion</strong> awareness and all cues<br />

(noticeable stimuli), which can be generated by these sub<strong>system</strong>s, have to be considered in<br />

evaluating a manual <strong>control</strong> task.<br />

The main unknown remains to what extent the <strong>simulator</strong> environment reflects the inthe-air<br />

conditions and to what extent this is required to achieve a certain training or test<br />

objective. First, the <strong>simulator</strong> is only as good as the data it contains i.e. the quality <strong>of</strong><br />

the model representing the aircraft and the <strong>flight</strong> environment to be simulated. Secondly,<br />

inherent limitations <strong>of</strong> the <strong>system</strong> used to simulate can be the cause <strong>of</strong> mismatches. The<br />

<strong>motion</strong> <strong>system</strong> <strong>of</strong> a <strong>simulator</strong> does posses one <strong>of</strong> the most obvious constraints with its very<br />

limited amount <strong>of</strong> travel. Fortunately, the pilot can be made to perceive the <strong>motion</strong> desired<br />

if the main cues are correctly simulated, allowing for some deviation in the others.<br />

1.1.2 Mechanism <strong>of</strong> <strong>motion</strong> perception<br />

Perception <strong>of</strong> <strong>motion</strong> has been studied for more than hundred years [88]. At that time,<br />

mainly experimental data was obtained. In the late 60’s <strong>of</strong> the past century models were<br />

proposed, which described the human behaviour in a <strong>motion</strong> <strong>control</strong> loop [69],[94]. With<br />

the crossover model <strong>of</strong> McRuer et al. [94], the pilot acts according to a describing function,<br />

H(j!), over the frequency, !,<br />

H(j!)=Kh 1j! +1<br />

2j! +1 e; dj! : (1.1)<br />

This model assumes only one main cueing input and considered additional dynamics <strong>of</strong> the<br />

sensory and response <strong>system</strong> as a simple time delay, e ; dj! . The parameters, gain, Kh, lead<br />

time constant, 1, and lag time constant, 2, are adapted such that a slope <strong>of</strong> -20 dB/decade


1.1 Flight simulation 5<br />

in the describing function frequency plot around the crossover frequency, ! c is attained <strong>of</strong><br />

the open loop <strong>system</strong> <strong>of</strong> human and the <strong>system</strong> to be <strong>control</strong>led, which usually results in<br />

a sufficient stability margin in the closed loop. This frequency is approximately 2 rad/s to<br />

5 rad/s in case <strong>of</strong> a <strong>flight</strong> <strong>control</strong> task (involving tracking and disturbance rejection) [53, 94,<br />

149]. With aircraft vehicle <strong>control</strong>, on which this model was experimentally validated, the<br />

open loop transfer function with a double integrator, 1=(j!) 2 , can be <strong>control</strong>led with PD,<br />

proportional/differential action in which 2 < 1=(!c) < 1.<br />

The optimal <strong>control</strong> model <strong>of</strong> Kleinman et al.[69] assumes that the pilot operates in an<br />

optimal way, given all the perceived cues with their respective limitations. This motivated<br />

the research into the mechanism <strong>of</strong> <strong>motion</strong> perception ([159] and the references therein) with<br />

a model <strong>based</strong> approach in which the interaction between the different stimuli, e.g. inertial<br />

<strong>motion</strong> and visual inputs, is taken into account. These models are still being refined [53]<br />

as the <strong>system</strong> is incompletely understood. Though the quantitative values are seen to be<br />

dependent on the subject, task and environment involved, the qualitative <strong>control</strong> behaviour<br />

<strong>of</strong> a well trained subject is reasonable consistent and reproducible.<br />

Motion results in stimuli inputs to the sensory organs. Above some sensory threshold,<br />

the sensory organs start to fire pulses to the central nervous <strong>system</strong>. The subject perceives<br />

<strong>motion</strong> from the set <strong>of</strong> cues generated in this way. These cues have to be reasonably consistent<br />

relative to each other. Otherwise disorientation and <strong>motion</strong> sickness can occur. Some<br />

deviations can be tolerated as long as the coherence functions <strong>of</strong> the subject from each cue to<br />

the corresponding movement have sufficient overlap [148]. Advantage from these tolerable<br />

differences is taken in simulation, which allows perceived <strong>motion</strong> to be simulated without<br />

the need for very large <strong>simulator</strong> excursions.<br />

The main sensory mechanisms to detect <strong>motion</strong> as given in the AGARD Advisory report<br />

159, [2] are:<br />

- Semicircular canals and Otoliths are called the vestibular organs. Both reside in the<br />

inner ear. The function <strong>of</strong> the semicircular canals is roughly that <strong>of</strong> rate gyros, from<br />

which angular velocity can be measured. The otoliths sense the specific force which<br />

is the combination <strong>of</strong> inertial acceleration and gravity.<br />

- Tactile or somatosensory receptors permit sensing a change <strong>of</strong> force on the body, e.g. due<br />

to a change in orientation. These sensors have a high-pass characteristic, which causes<br />

sustained uniform pressure to be faded to a certain reference level. Especially the fingertips<br />

are very capable detecting high frequency vibrations.<br />

- Proprioceptive and kinesthetic senses lead to information about the relative positions <strong>of</strong><br />

the parts <strong>of</strong> the body as well as their movements i.e. the body acceleration. This is<br />

attained through the double sided muscular coupling <strong>of</strong> force and velocity/position<br />

and an internal model <strong>of</strong> the body<br />

- The eyes primarily detect <strong>motion</strong> through the peripheral retina. With uniform <strong>motion</strong> <strong>of</strong><br />

a wide visual field, it is possible to create a so-called self <strong>motion</strong> sensation, which<br />

is called vection. The foveal area <strong>of</strong> the retina or central part is primarily associated<br />

with scanning and recognition i.e. the cognitive sense <strong>of</strong> self-<strong>motion</strong>.


6 1 Introduction<br />

Advani [7] also mentions the auditory <strong>system</strong> in the ears, from which through sound <strong>motion</strong><br />

can be detected.<br />

With the visual and auditory <strong>system</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong>, the pilot can be isolated from<br />

detecting inertial <strong>motion</strong> through sound or sight. With the other organs, inertial <strong>motion</strong> cues<br />

will be perceived as <strong>motion</strong>. In case <strong>of</strong> required high forces, the proprioceptive and tactile<br />

senses can also be stimulated with a so-called g-suit, mainly used in military applications.<br />

Control forces and vibrations provided by the <strong>control</strong> loading <strong>system</strong> will also influence the<br />

tactile senses.<br />

The vestibular <strong>system</strong> is considered to generate the main cues in the frequency area<br />

around the crossover frequency (:3 to 1Hz), essential to learn skill <strong>based</strong> <strong>control</strong> behaviour.<br />

This is the central issue in <strong>motion</strong>-<strong>based</strong> <strong>flight</strong> simulation. In order to recreate the skill<br />

<strong>based</strong> behaviour, the pilot in the <strong>simulator</strong> must be provided with well coordinated visual<br />

and vestibular cues. Motion cues around the crossover frequency frequency area can only<br />

be stimulated with actual inertial <strong>motion</strong> and this is the main reason a <strong>motion</strong> <strong>system</strong> is<br />

essential in case <strong>of</strong> piloted <strong>control</strong> tasks in <strong>flight</strong> simulation.<br />

In the low frequency area ( :1 Hz), <strong>motion</strong> is much easier to detect through the visual<br />

information <strong>of</strong> position and orientation at the instrument displays and exterior screen. Some<br />

mismatch with inertial <strong>motion</strong>, with the already low accelerations related to these frequencies,<br />

is allowed, even slightly superthreshold, without the pilot noticing discrepancies.<br />

Schroeder [126] notes that the review <strong>of</strong> different studies resulted in a wide range <strong>of</strong><br />

threshold values. This is confirmed by the study reported in [54] in which it is shown that a<br />

threshold depends on several factors such as the frequency contents <strong>of</strong> the stimulus and the<br />

task the subject has to fulfil i.e. mental load. Values range for the translational accelerations<br />

from :08 m=s 2 in heave, no task involved to :7 m=s 2 in a very busy situation experiencing<br />

surge (longitudinal direction). Rotational thresholds are found to be in a closer range <strong>of</strong><br />

:8 deg =s 2 to 1:5 deg =s 2 for pitch and roll, though Schroeder [126] reports a range <strong>of</strong><br />

:0035 deg =s 2 to 2 deg =s 2 for yaw <strong>motion</strong> (vertical axis rotation).<br />

Hosman [53] provides some transfer functions for the vestibular <strong>system</strong>. Both the semicircular<br />

channels as the otholits are modelled according to a function specified in the frequency<br />

domain:<br />

xout<br />

=<br />

xin<br />

K( Lj! +1)<br />

� (1.2)<br />

( 1j! + 1)( 2j! +1)<br />

where the ratio between (actual) input acceleration, x in, and noticed acceleration (xout in<br />

impulses per second, ips), is given by a second order transfer function with gain, K, two<br />

lag time constants, 1 and 2, and a lead time constant, L.<br />

Although the models <strong>of</strong> the two vestibular organs, which give the sensitivity over the frequency,<br />

have the same structure, the characteristic is different. This is due to the parameters,<br />

which are given by Hosman [53] as in Table 1.1.<br />

This means the sensitivity <strong>of</strong> the semicircular channels drops over the range from 0.025 Hz<br />

to 1.4 Hz after which it starts to drop again at 30 Hz, while the otholits become more sensitive<br />

from .16 Hz to .3 Hz, after which it starts to drop again at 10 Hz. This opposite effect<br />

in the relative important frequency range could maybe explain the fact that different conclusions<br />

can be drawn towards the question whether translational accelerations or angular


1.1 Flight simulation 7<br />

Semicircular Channels Otholits<br />

K 2 ips=( =s 2 ) 3:4 ips=(m=s 2 )<br />

L 0:11 s 1 s<br />

1 5:9 s 0:5 s<br />

2 0:005 s 0:016 s<br />

Table 1.1: The parameters <strong>of</strong> the models <strong>of</strong> the vestibular organs<br />

accelerations are more important. In case <strong>of</strong> roll and lateral <strong>motion</strong> in large aircrafts and yaw<br />

and lateral <strong>motion</strong> in helicopters, lateral <strong>motion</strong> cues are shown to be more important [126].<br />

In other cases reported in [54], [126] rotational cues are considered more relevant. Chung<br />

et al. [22] conclude that the translational and rotational <strong>motion</strong> relative to each other should<br />

not differ in phase nor amplitude since these types <strong>of</strong> <strong>motion</strong> are coupled as a function <strong>of</strong><br />

place.<br />

The <strong>motion</strong> cues provided through the vestibular <strong>system</strong> with inertial <strong>motion</strong>, change<br />

the describing function <strong>of</strong> the pilot [53], [149]. With the introduction <strong>of</strong> <strong>motion</strong>, the time<br />

delay in the describing function is seen to drop from 0.23 s to 0.12 s in case <strong>of</strong> a disturbance<br />

rejection task. The crossover frequency a pilot can achieve is lower in the absence <strong>of</strong> inertial<br />

<strong>motion</strong> and, as a result, also the performance is. Hall [47] argues and confirms by an<br />

example that, even with equivalent performance, the pilot can develop a different and false<br />

strategy if the cues in a <strong>simulator</strong> deviate from those in the aircraft. Especially in case <strong>of</strong><br />

relatively inexperienced pilots, all cues are important in the learning process <strong>of</strong> picking the<br />

most relevant [143].<br />

There is no doubt inertial <strong>motion</strong> is necessary in the appropriate perception <strong>of</strong> <strong>motion</strong>.<br />

Not much is known to what extent the inertial <strong>motion</strong> <strong>of</strong> the <strong>simulator</strong> is allowed to differ<br />

from the aircraft. In the low frequency area it is concluded that high pass filtering at .08 Hz is<br />

acceptable [149]. The crossover region between .1 Hz and 1 Hz is likely to be most sensitive<br />

to deviations. At higher frequencies, the situation is less clear. Pilots are seen to provide<br />

lead well over 3 Hz and consequently still respond to <strong>motion</strong> with increasing gain in this<br />

frequency area. Further, pilots respond much earlier when inertial <strong>motion</strong> is present. In the<br />

more general requirements <strong>of</strong> the international standard ISO 2631 [138], which evaluates<br />

human exposure to whole-body vibration, it is noted that humans are most sensitive to heave<br />

accelerations in the frequency area between 4 Hz and 8 Hz. The bandwidth <strong>of</strong> many <strong>motion</strong><br />

<strong>system</strong>s in this frequency area makes these <strong>system</strong>s prone to errors exactly in the area where<br />

the level <strong>of</strong> comfort is most sensitive.<br />

Over the years, many discrepancies in the <strong>motion</strong> <strong>system</strong>s <strong>of</strong> <strong>simulator</strong>s have been considered<br />

acceptable. There was no economical or technical feasible alternative. Other sub<strong>system</strong>s<br />

were also unable to represent reality in a satisfactory manner. A pilot was required<br />

to fly an aircraft on the basis <strong>of</strong> instrument displays and external sight without <strong>motion</strong> cues.<br />

Being able to fly a <strong>simulator</strong> without appropriate inertial <strong>motion</strong> qualified since it is more<br />

difficult than with an aircraft. In-<strong>flight</strong> training still formed a considerable part <strong>of</strong> the learning<br />

process. This point <strong>of</strong> view is changing as can be seen by putting the current state <strong>of</strong> the<br />

art in a historical perspective.


8 1 Introduction<br />

1.1.3 Historical perspective<br />

In [4], Adorian et al. provide a nice overview <strong>of</strong> the development <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong>s over<br />

the years. Almost directly with the introduction <strong>of</strong> the first aeroplanes, the first <strong>simulator</strong>s<br />

were built. Already in those days, a <strong>motion</strong> <strong>system</strong> was part <strong>of</strong> the <strong>simulator</strong>. Around 1910,<br />

these <strong>motion</strong> <strong>system</strong>s had to be driven by hand by the instructor, or wind, and mainly had<br />

to give the experience <strong>of</strong> a changing orientation around one to three axes. Soon the inertial<br />

<strong>motion</strong> was provided automatically. The Ruggles Orientator, developed in 1917, had full rotation<br />

around three axes and additionally vertical movement, all driven by electrical motors.<br />

The most successful device, the Link trainer, developed in 1927-29, was used well into the<br />

50’s. It was driven pneumatically and operated by the stick and rudder in the cockpit. The<br />

simulated effects were adjusted by trial and error independently for ailerons, elevators and<br />

rudder. In this way, the aircraft’s coordinated behaviour could not be represented properly.<br />

After 1930, for thirty years, the main developments concerned other sub<strong>system</strong>s than the<br />

<strong>motion</strong> <strong>system</strong> <strong>of</strong> the <strong>simulator</strong>. Early in the Second World Ward, hundreds <strong>of</strong> the so-called<br />

Celestrial Navigation Trainers, massive bomber crew trainers, were built. A moving dome<br />

provided for the ’correct’ out-<strong>of</strong>-the-window sight in case <strong>of</strong> a nightly atlantic ocean crossing.<br />

During the war, the visual <strong>system</strong> was developed further into actual image projection<br />

<strong>system</strong>s.<br />

With the instruments in the cockpit, all basic flying behaviour, all engine, electrical and<br />

hydraulic <strong>system</strong>s could be simulated. The computation, at first pneumatically, could solve<br />

the equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> an aircraft with the introduction <strong>of</strong> the analog computer.<br />

The first full aircraft <strong>simulator</strong> operated by an airline was installed in 1948 (a Boeing<br />

377 by PanAm). Most full aircraft <strong>simulator</strong>s did not have a <strong>motion</strong> <strong>system</strong> at that time,<br />

which was justified by the statement that ”modern pilots should not fly by the seat <strong>of</strong> their<br />

pants.” Partly, the <strong>control</strong> loading <strong>system</strong> was thought to compensate for the lack <strong>of</strong> <strong>motion</strong><br />

and provide for a realistic feel.<br />

With the improvement <strong>of</strong> <strong>flight</strong> test data and the increasing complexity <strong>of</strong> the aircrafts,<br />

the analog computer became the bottle-neck in the <strong>simulator</strong>. The demand for increased<br />

fidelity and reliability motivated the introduction <strong>of</strong> the digital computer in the <strong>flight</strong> <strong>simulator</strong><br />

<strong>system</strong>.<br />

However, correlation was poor with <strong>flight</strong> tests concerning low damping, stability or<br />

characteristics <strong>of</strong> pilot induced oscillations. Further, strong visual cues supplied by the<br />

improved wide field <strong>of</strong> view visual <strong>system</strong> were disorienting in the absence <strong>of</strong> <strong>motion</strong> cues.<br />

In 1959, the first <strong>motion</strong> research <strong>simulator</strong> was built at NASA, which also indicated the<br />

importance <strong>of</strong> <strong>motion</strong> cues [10]. Analysis <strong>of</strong> loss <strong>of</strong> <strong>control</strong> in extreme operations motivated<br />

the research with centrifuge <strong>simulator</strong>s [23]. The many <strong>motion</strong> deficiencies, due to the<br />

very limited vertical stroke, the vibration noise and incorrect centrifugal forces, were so<br />

distracting that most <strong>of</strong> the research was still performed with fixed base <strong>simulator</strong>s.<br />

Motion <strong>system</strong>s with alternative construction were designed. The introduction <strong>of</strong> the<br />

wide body transport aircraft, such as the B747, required lateral acceleration and led to the<br />

four and six degrees-<strong>of</strong>-freedom (d.o.f.) <strong>motion</strong> <strong>system</strong>s. The first six-degrees-<strong>of</strong>-freedom<br />

research <strong>simulator</strong> at NASA Ames was put into operation in 1964. These <strong>system</strong>s were <strong>of</strong>ten<br />

given too much stroke (up to 30 m), which was not necessary given the limited velocity and<br />

resulted in inferior dynamics <strong>of</strong> lengthy cables (low eigenfrequency). Also the turn-around-


1.1 Flight simulation 9<br />

bump due to coulomb friction prohibited unnoticeable fade to a central position.<br />

In the late sixties, the synergetic, fully parallel, hydraulically driven <strong>motion</strong> <strong>system</strong> came<br />

into operation. The most commonly used is the Stewart-Gough platform [140]. Through<br />

the hydrostatic bearings introduced by Viersma [153] in 1969 in the first (3-d.o.f.) <strong>flight</strong><br />

<strong>simulator</strong> <strong>motion</strong> <strong>system</strong> at Delft University <strong>of</strong> Technology [12], these <strong>system</strong>s had virtually<br />

no coulomb friction. The first commercially available 6-d.o.f. <strong>motion</strong> <strong>system</strong> in 1977 with<br />

hydrostatic bearings was reported on by Baret in [13]. Nowadays, almost every <strong>simulator</strong><br />

<strong>motion</strong> <strong>system</strong> operates in this manner.<br />

In the late 70’s, an attempt was made to classify the characteristics <strong>of</strong> <strong>motion</strong> <strong>system</strong>s in<br />

<strong>flight</strong> <strong>simulator</strong>s [1], [67]. Travelling limits, response time (latency), roughness and noise<br />

were considered most important. The regulating authorities such as the FAA (USA) and<br />

JAR (Europe), given confidence in the use <strong>of</strong> simulation, came up with required <strong>motion</strong><br />

<strong>system</strong> specifications, which reflect the abilities <strong>of</strong> <strong>motion</strong> <strong>system</strong>s at that time [3], [25],<br />

[118], [139]. It is argued that the fact that these standards have been basically unchanged<br />

for the past 15 years, has stopped virtually all fundamental development in <strong>flight</strong> <strong>simulator</strong>s<br />

[143]. Further, since there are still no objective tests, no prescribed manoeuvres and no<br />

criteria for accurate <strong>motion</strong> cueing, makes it difficult to establish new standards. There is,<br />

however, a market demand from the airlines to come up with additional knowledge to enable<br />

a decrease in required <strong>flight</strong> experience for zero <strong>flight</strong> time training.<br />

In the 80’s and 90’s, the main developments concerned the rapid growing abilities in<br />

the visual and computational <strong>system</strong>s. As the main selling argument in the 60’s came from<br />

the all but perfect though impressive <strong>motion</strong> <strong>system</strong>s, now most interest stems from the fast<br />

improving, marketable, visual <strong>system</strong>. As in the past, higher standards for other sub<strong>system</strong>s,<br />

such as the quick response <strong>of</strong> the image generating <strong>system</strong>s now available [100], the<br />

increased computational power enabling to include aer<strong>of</strong>lexible modes in the aircraft model<br />

and more advance <strong>control</strong> strategies for <strong>motion</strong> <strong>control</strong>, will be a drive for higher fidelity<br />

with respect to <strong>motion</strong>.<br />

In [10] it is not expected that any major changes in <strong>motion</strong> travel will be made in <strong>flight</strong><br />

simulation <strong>motion</strong> <strong>system</strong>s. However, within the current structure improvements can be attained<br />

[7] and more <strong>of</strong>ten multi-stage, cascaded, <strong>motion</strong> <strong>system</strong>s can be observed, in which<br />

the need for <strong>motion</strong> cues (with the current 6-d.o.f. structure) and travel (additional for lateral<br />

and longitudinal travel over rails as in the National Advanced Driving Simulator [20], [92])<br />

have been decoupled but still have considerable interaction in the dynamics. Reduction <strong>of</strong><br />

time delays in <strong>motion</strong> <strong>system</strong> response and interaction is needed for solving more complex<br />

problems with visual/<strong>motion</strong>/model mismatches [10]. At this moment, high frequency <strong>motion</strong><br />

cues ( 4 Hz) require special effects included in the <strong>motion</strong> <strong>system</strong> <strong>control</strong>ler and<br />

phase lag in the crossover region is compensated for by additional lead (differential action)<br />

on the reference signals [10], [24], [143]. A more structural model <strong>based</strong> <strong>based</strong> approach<br />

will be strived for in this research.<br />

1.1.4 Motion simulation<br />

As already mentioned, in <strong>flight</strong> simulation, the pilot relies on the perception <strong>of</strong> self-<strong>motion</strong><br />

through several stimuli, and uses this perception to exercise <strong>control</strong> over the aircraft. From


10 1 Introduction<br />

Experimentator<br />

Situation<br />

Aircraft<br />

<strong>Model</strong><br />

Wash-Out<br />

Filter<br />

Pilot<br />

Motion<br />

Controller<br />

Simulation computer Motion Computer<br />

Motion<br />

System<br />

Continuous World<br />

Fig. 1.1: A block diagram representation <strong>of</strong> <strong>motion</strong> simulation.<br />

the pilots point <strong>of</strong> view, the <strong>simulator</strong> should be able to replace the aircraft.<br />

In the <strong>control</strong> task, the pilot uses his visual perception to make a good estimation <strong>of</strong><br />

the aircraft’s long term attitude and velocity. The approximate frequency response <strong>of</strong> visual<br />

perception can be modelled as a first order low-pass filter with a break frequency <strong>of</strong> 0.1 Hz<br />

[92]. Fortunately, the inertial <strong>motion</strong> <strong>of</strong> the <strong>simulator</strong> has a minor role below this frequency.<br />

There is a potential pay<strong>of</strong>f in tuning the <strong>motion</strong> cues to match the human physiological<br />

sensors’ responses, rather than attempting to match the aircraft <strong>motion</strong> themselves. If done<br />

properly, the effective provision <strong>of</strong> <strong>motion</strong> can be obtained in an envelope substantially<br />

reduced from that <strong>of</strong> the aircraft <strong>motion</strong>.<br />

The stimulation through actual inertial <strong>motion</strong> with a <strong>simulator</strong> can usually be represented<br />

by the block scheme given in Fig. 1.1. The pilot responds to <strong>simulator</strong> cues and<br />

the task at hand provided through the experiment leader. With a model <strong>of</strong> the aircraft, it is<br />

calculated what the actual aircraft <strong>motion</strong> would have done due to the action <strong>of</strong> the pilot,<br />

the situation (on the ground the aircraft has different dynamics in interacting with the track<br />

than in the air) and the situation dependent disturbances such as turbulence.<br />

With washout, the calculated aircraft <strong>motion</strong> has to be translated into feasible <strong>motion</strong><br />

<strong>system</strong> trajectories regarding the limited stroke, velocity and accelerations. Washout must<br />

provide some form <strong>of</strong> high pass filtering to limit the <strong>simulator</strong> excursions given the unlimited<br />

aircraft travel. Washout does usually separate the <strong>motion</strong> cues into high (”onset”)<br />

and low (”sustained”) frequency components. Further, by also introducing a scaling factor<br />

<strong>of</strong> <strong>simulator</strong> <strong>motion</strong>, all onsets can be performed by the <strong>simulator</strong> without attaining the<br />

inherent limitations <strong>of</strong> the <strong>motion</strong> <strong>system</strong>.<br />

Most sustained cues can not be represented through inertial <strong>motion</strong> <strong>of</strong> the <strong>simulator</strong> and<br />

have to be ”washed out”. Sustained forward or lateral acceleration is sometimes represented<br />

through a trick already mentioned by Johnson in 1931 [4]. Through the help <strong>of</strong> the gravity<br />

force, a backward tilt gives the same sensation as a forward linear acceleration. Problem


1.1 Flight simulation 11<br />

Fig. 1.2: A typical modern <strong>flight</strong> <strong>simulator</strong> equipped with a synergistic (hexapod) <strong>motion</strong><br />

<strong>system</strong>. 1<br />

with this so-called tilt coordination is to attain the tilted pose without the pilot noticing<br />

i.e. pitching or rolling with subthreshold angular <strong>motion</strong>. In a critical survey, Advani [7]<br />

notices that the mostly used classical linear washout leads to considerable false cues in the<br />

important frequency area between 0.1 Hz and 1 Hz. In an overview <strong>of</strong> wash-out techniques<br />

in [92] by Martin, also non-linear wash-out is mentioned, which however has not led to<br />

completely acceptable solutions so far.<br />

The feasible washed-out <strong>motion</strong> trajectories are the reference input to the <strong>motion</strong> <strong>control</strong>ler.<br />

The task <strong>of</strong> the <strong>motion</strong> <strong>control</strong>ler here is to attain a one-to-one correspondence<br />

between the actual <strong>motion</strong> <strong>system</strong> trajectories and the desired feasible washed-out <strong>motion</strong><br />

trajectories. In the usual configuration, the references contain positional information, which<br />

is also measured and can be used in feedback. Further, the desired acceleration pr<strong>of</strong>iles are<br />

provided for, which can be applied in generating an appropriate feed forward.<br />

As it is physically impossible to generate immediate acceleration with the <strong>motion</strong> <strong>system</strong>,<br />

imperfect cues are generated, which lag the desired cues. Enlarging the bandwidth<br />

decreases the lag but has to be compromised with sufficient attenuation <strong>of</strong> false cues resulting<br />

from noise and excitation <strong>of</strong> parasitic dynamics, which mainly reside in the high<br />

frequency area. Considering a more general setting than the one given in Fig. 1.1, can possibly<br />

lead to a less severe compromise by e.g. allowing available predictive information in<br />

the simulation model to be sent to the <strong>motion</strong> <strong>control</strong>ler.<br />

The <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s, nowadays, have almost invariably a so-called 6d.o.f.<br />

synergetic construction. A <strong>motion</strong> <strong>system</strong> <strong>of</strong> this type is shown in Fig. 1.2. 1 Six<br />

1 Picture courtesy <strong>of</strong> Rexroth-Hydraudyne.


12 1 Introduction<br />

actuators are connected in parallel to a base frame at the foundation with one gimbal, a<br />

freely rotating joint, and at the other end to the <strong>simulator</strong> with another gimbal. Kinematically,<br />

the six lengths <strong>of</strong> the actuators can be driven independently. However, dynamically,<br />

this <strong>system</strong> is highly interactive i.e. an elongation force <strong>of</strong> one actuator generally induces reaction<br />

forces in all others. Further, prescribed platform <strong>motion</strong> requires coordinated <strong>motion</strong><br />

<strong>of</strong> all actuators, which varies with the platform position.<br />

The main advantages <strong>of</strong> this construction are the parallel connections resulting in a <strong>system</strong><br />

with higher stiffness given a specific actuator and its simplicity. Six identical <strong>system</strong><br />

are easy to build and maintain. The stiffness is due to the fact that springs in parallel result<br />

in higher stiffness while springs in series will result in a decrease. This is important in <strong>flight</strong><br />

<strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s since the payload to be accelerated is usually relatively high. With<br />

the synergetic construction the static load per actuator will be lower and the eigenfrequencies<br />

will shift to higher values. The forces to be applied with the actuators are usually still<br />

considerably high. Most <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s are therefore driven by hydraulic<br />

servo actuators, since these <strong>system</strong>s provide superior performance in generating high power<br />

long stroke linear <strong>motion</strong>s [124]. Further, the use <strong>of</strong> hydrostatic bearings results in very low<br />

coulomb friction forces, which minimises the velocity reversal bump. As <strong>simulator</strong> wash<br />

out <strong>motion</strong> reversal and aircraft <strong>motion</strong> do not coincide, an unnoticeable change <strong>of</strong> velocity<br />

sign is very important in <strong>flight</strong> simulation.<br />

With the introduction <strong>of</strong> improved linear electrical drives, there is a trend to use electrically<br />

driven <strong>motion</strong> <strong>system</strong> if a <strong>simulator</strong> has relatively low payload up to 2500 kg<br />

[24].<br />

1.1.5 SIMONA project<br />

Through a preliminary investigation for an upgrade programme <strong>of</strong> the older DUT 3-d.o.f.<br />

<strong>simulator</strong>, a new international centre for research into <strong>flight</strong> simulation techniques, Simona,<br />

was initiated in 1992. Simona stands for Simulation, Motion and Navigation, which reflects<br />

the areas <strong>of</strong> interest in the research programme <strong>of</strong> the three groups, who initiated this centre.<br />

Instead <strong>of</strong> an upgrade <strong>of</strong> the older <strong>simulator</strong>, which has been dismantled, the core <strong>of</strong> Simona<br />

will be a full scale 6-d.o.f. <strong>flight</strong> <strong>simulator</strong>, the Simona Research Simulator (SRS) [6]. The<br />

development <strong>of</strong> the SRS is also performed within Simona. The actual fabrication <strong>of</strong> the<br />

<strong>simulator</strong> was started in 1994, after a large donation <strong>of</strong> the Dutch Government to enable a<br />

multi-discipline international centre with a high fidelity <strong>simulator</strong>.<br />

The three groups involved, each have their specific research interest and contribution to<br />

the project.<br />

The Simulation and Control Group <strong>of</strong> the Faculty <strong>of</strong> Aerospace Engineering performs<br />

studies into <strong>flight</strong> <strong>control</strong>, the man-machine interaction <strong>of</strong> the pilot in the aircraft<br />

and modelling and real-time simulation <strong>of</strong> the aircraft and its <strong>system</strong>s. In the, now<br />

ongoing, realisation phase <strong>of</strong> the <strong>simulator</strong> the group takes part in the development<br />

<strong>of</strong> simulation s<strong>of</strong>tware, the interior and the visual <strong>system</strong>. The Group <strong>of</strong> Production<br />

Technology within Aerospace Engineering did the design <strong>of</strong> the integrated platformcockpit<br />

shuttle and planned connections <strong>of</strong> projection platform and image projection<br />

screens, which will be connected in future.


1.1 Flight simulation 13<br />

Fig. 1.3: Artist’s impression <strong>of</strong> what will be the full operational Simona <strong>flight</strong> <strong>simulator</strong>


14 1 Introduction<br />

Fig. 1.4: Simona <strong>motion</strong> <strong>system</strong> with empty shuttle on top, configuration B.


1.1 Flight simulation 15<br />

The Mechanical Engineering Systems and Control Group <strong>of</strong> the Faculty <strong>of</strong> Design,<br />

Engineering and Production performs research into the design, development, modelling<br />

and <strong>control</strong> <strong>of</strong> hydraulic servo and <strong>motion</strong> <strong>system</strong>s. The 6-d.o.f. hydraulically<br />

driven <strong>motion</strong> <strong>system</strong> has been realised and tested and so have the first model <strong>based</strong><br />

<strong>control</strong>lers as part <strong>of</strong> this research. In future, alternative <strong>control</strong> concepts will have to<br />

be developed. For one, to interconnect simulation models and <strong>motion</strong> <strong>control</strong> in an<br />

optimal way. Further, to adapt to the changing characteristics <strong>of</strong> the <strong>motion</strong> <strong>system</strong><br />

with e.g. a, to some extent flexibly attached, visual <strong>system</strong>.<br />

The Telecommunications and Traffic-Control Systems and Services Group <strong>of</strong> the Faculty<br />

<strong>of</strong> Information Technology and Systems does research into navigation technology<br />

and advanced cockpit display <strong>system</strong>s. The operational SRS can serve as an experimental<br />

set-up to test new navigation techniques and approaches to simulate these<br />

techniques appropriately. The group has been involved in the development <strong>of</strong> a realtime<br />

simulation platform.<br />

An impression <strong>of</strong> what will be the full operational <strong>flight</strong> <strong>simulator</strong> is given in Fig. 1.3.<br />

The <strong>motion</strong> <strong>system</strong> consists <strong>of</strong> six hydraulic servo actuators which are connected in parallel<br />

at the base frame and <strong>simulator</strong>. The actuators together can drive the <strong>simulator</strong> in all three<br />

translational and rotational degrees <strong>of</strong> freedom. At the <strong>simulator</strong> side, the actuators are connected<br />

to the bottom <strong>of</strong> the shuttle, an integrated platform/cockpit made <strong>of</strong> the light weight<br />

material TWARON/carbon. In front <strong>of</strong> the shuttle, a wide angle visual mirror projection<br />

screen will be attached. With projectors on top <strong>of</strong> the shuttle and a back projection screen<br />

in between, the pilot in the cockpit can be provided with a out-<strong>of</strong>-the-window view via the<br />

mirror in front <strong>of</strong> the shuttle.<br />

The new design <strong>of</strong> the integrated platform/cockpit and the use <strong>of</strong> light weight materials<br />

will lead to a relative low centre <strong>of</strong> gravity and considerably less mass to be accelerated<br />

as compared to conventional <strong>simulator</strong>s. These properties strived for in the design <strong>of</strong> the<br />

construction, will have to result in favourable <strong>motion</strong> characteristics. In what sense the<br />

expectedly 4 tons <strong>of</strong> the SRS instead <strong>of</strong> the 12 to 15 tons <strong>of</strong>ten reported for other <strong>system</strong>s<br />

[24], [92] can be taken advantage <strong>of</strong>, is one <strong>of</strong> the issues to be discussed in this research.<br />

After separate testing <strong>of</strong> all six hydraulic actuators in a special experimental set-up in<br />

1994 [124], the integration <strong>of</strong> the <strong>motion</strong> <strong>system</strong> was complete early in 1997, and the first<br />

free-run tests <strong>of</strong> the integrated Simona <strong>motion</strong> platform were performed. The <strong>motion</strong> <strong>system</strong><br />

was tested in three configurations in the Central Workshop at the faculty <strong>of</strong> Mechanical<br />

Engineering.<br />

A. The first tests were performed with a steel platform, the dummy platform <strong>of</strong> 2200 kg.<br />

B. In the autumn <strong>of</strong> 1997 the empty (only chairs) shuttle <strong>of</strong> 1700 kg was put on top <strong>of</strong> the<br />

<strong>motion</strong> <strong>system</strong>, as shown in Fig. 1.4.<br />

C. Finally, tests were performed with the estimated final load <strong>of</strong> 4000 kg with the dummy<br />

platform and additional masses (6 times 250 kg) on top, shown in Fig. 1.5, in the<br />

months <strong>of</strong> May and June, 1998.<br />

In the early spring <strong>of</strong> 1999, the <strong>motion</strong> <strong>system</strong> was to be operated at the building especially<br />

made to house the Simona centre, eventually with the full operational <strong>simulator</strong>. Main


16 1 Introduction<br />

Fig. 1.5: Simona <strong>motion</strong> <strong>system</strong> with dummy platform and additional masses, configuration<br />

C.


1.2 Motion <strong>control</strong> 17<br />

w z<br />

- -<br />

- Psp u<br />

K<br />

Fig. 1.6: A standard <strong>control</strong> design structure<br />

efforts are now in progress to complete the cockpit interior, the real-time simulation environment<br />

and the visual <strong>system</strong>.<br />

1.2 Motion <strong>control</strong><br />

Having sketched the <strong>flight</strong> simulation process and the place <strong>motion</strong> generation has in this<br />

environment, the <strong>motion</strong> <strong>control</strong> problem can be defined. First the process <strong>of</strong> <strong>motion</strong> generation<br />

is put into a generalised framework. From this perspective the <strong>motion</strong> <strong>control</strong> specifications<br />

are stated and the conventional approach towards the design <strong>of</strong> <strong>motion</strong> <strong>control</strong> in<br />

<strong>flight</strong> <strong>simulator</strong>s is discussed. Then realising that the <strong>motion</strong> <strong>system</strong> <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> can<br />

be considered a robotic <strong>system</strong>, <strong>of</strong> course with its specific demands and properties, some <strong>of</strong><br />

the modern <strong>control</strong> strategies developed in robot <strong>control</strong> are put forward, since these could,<br />

with some modification, possibly be <strong>of</strong> benefitin<strong>flight</strong> simulation.<br />

1.2.1 Motion <strong>control</strong> structure<br />

A standard <strong>control</strong> framework, adopted in many textbooks on modern <strong>control</strong> [17], [89],<br />

[160], is given in Fig. 1.6. A <strong>control</strong>ler, K, is provided with measurement signals, y, and<br />

has to stabilise a plant, Psp, with input signals, u, such that the cost variables, z, are minimal<br />

in some sense, despite the disturbance signals, w.<br />

The plant, Psp, is <strong>of</strong>ten called the generalised plant or standard plant since it usually<br />

does not only consists <strong>of</strong> the plant to be <strong>control</strong>led, the <strong>motion</strong> <strong>system</strong> in <strong>flight</strong> simulation<br />

<strong>motion</strong> <strong>control</strong>, but can also contain weightings e.g. sensory thresholds weighting on the<br />

cost variable specific force or a reference model e.g. the aircraft to be simulated. Also the<br />

other entities can be viewed in a generalised way, e.g. reference signals can be incorporated<br />

as disturbances w and additional feedback paths can be taken in case <strong>of</strong> a robust <strong>control</strong><br />

problem to describe a set <strong>of</strong> <strong>system</strong>s i.e. uncertainty.<br />

In a general simulation <strong>motion</strong> <strong>control</strong> problem, as shown in Fig. 1.7, the performance<br />

cost variables, zp, will be the difference <strong>of</strong> <strong>motion</strong> in the aircraft and in the <strong>simulator</strong> with<br />

the pilot in-the-loop weighted with the ability <strong>of</strong> the pilot to perceive these differences,<br />

Wp. Further, <strong>motion</strong> <strong>system</strong> limitations should be penalised by additional cost variables,<br />

zc. So the standard plant will consist <strong>of</strong> the <strong>motion</strong> <strong>system</strong>, the simulation model (Sim),<br />

the aircraft, the pilot and weighting functions. The generalised disturbances are both the<br />

y


18 1 Introduction<br />

wda<br />

wr<br />

wdm<br />

u<br />

-<br />

-<br />

-<br />

-<br />

-<br />

Aircraft<br />

Pilot<br />

Pilot<br />

Motion System<br />

Sim - Ww<br />

Motion Control<br />

e? - Wp -<br />

6<br />

Fig. 1.7: The <strong>flight</strong> simulation <strong>motion</strong> <strong>control</strong> problem put in the generalised plant structure<br />

tasks to be tracked, wr and the external disturbances to be rejected. There will be both<br />

external disturbances on the aircraft, wda, such as forces due to turbulence, which should<br />

be simulated, and external disturbances on the <strong>motion</strong> <strong>system</strong>, w dm, such as friction forces,<br />

which are to be rejected.<br />

The <strong>control</strong>ler receives information from both the <strong>motion</strong> <strong>system</strong>, y m, which has to<br />

be stabilised and the other sub<strong>system</strong>s <strong>of</strong> the generalised plant to construct an appropriate<br />

reference to be tracked, yr. In case <strong>of</strong> the <strong>motion</strong> <strong>control</strong>ler in the conventional setting as<br />

given in Fig. 1.1 this will be the desired <strong>motion</strong> reconstructed by the wash-out filters, W w.<br />

If the reconstruction <strong>of</strong> a feasible trajectory is taken as a part <strong>of</strong> the <strong>control</strong> problem, as is<br />

done by Idan et al. in [58], the aircraft <strong>motion</strong> can be taken as the reference and the wash<br />

out model becomes part <strong>of</strong> the <strong>motion</strong> <strong>control</strong>ler to be designed. When the whole inertial<br />

<strong>motion</strong> simulation is considered, the ’simulation <strong>control</strong>ler’ is fed by actions <strong>of</strong> the pilot and<br />

the experiment leader. The <strong>control</strong>ler can act through the actuators <strong>of</strong> the <strong>motion</strong> <strong>system</strong>, u.<br />

To calculate an appropriate <strong>control</strong>ler, a mathematical model <strong>of</strong> the sub<strong>system</strong>s in the<br />

standard plant can be derived. If these models are sufficiently simple, e.g. linear time-<br />

+<br />

-<br />

yr<br />

ym<br />

zp<br />

zc<br />

-


1.2 Motion <strong>control</strong> 19<br />

invariant, an optimal <strong>control</strong>ler can be synthesised, which for instance minimises some<br />

maximal amplification on some norm on the cost, z, given a bound on the norm <strong>of</strong> the<br />

disturbances, w [160]. The <strong>flight</strong> simulation <strong>motion</strong> <strong>control</strong> problem is yet too complex to<br />

be solved in this way, but some observations can be made from the structure <strong>of</strong> the standard<br />

plant given in Fig. 1.7.<br />

The task <strong>of</strong> tracking a reference model is a particular element in <strong>flight</strong> simulation,<br />

which requires special consideration. Several techniques, like model following [145], model<br />

matching [156], the servo mechanism [33], etc., exist to achieve such tasks. Tracking results<br />

through the servo-mechanism with linear time-invariant <strong>system</strong>s if the internal model<br />

principle [40] is satisfied, which states that the (unstable) parts <strong>of</strong> the <strong>system</strong> to be tracked<br />

have to be part <strong>of</strong> the <strong>control</strong>ler. Also with the other techniques, knowledge <strong>of</strong> the <strong>system</strong><br />

to be matched or followed is incorporated in the <strong>control</strong>ler (design). Although the reference<br />

aircraft model is available in <strong>flight</strong> simulation, the specific properties <strong>of</strong> these models are<br />

not yet taken into account in the <strong>motion</strong> <strong>control</strong>ler design.<br />

The reference tracking problem typically leads to a two degree <strong>of</strong> freedom <strong>control</strong>ler<br />

design problem [152], as given in Fig. 1.8, in which a feedforward, C 1, and a feedback part,<br />

C2, can be distinguished. With feedforward, ideally, the input, u, required to achieve the<br />

desired reference, r, should be generated, i.e. an inverse model <strong>of</strong> the <strong>system</strong>, G. Usually,<br />

however, the <strong>system</strong> does not have a causal and/or stable inverse and therefore an inverse<br />

model cannot be used. Further, the feedforward should not unnecessarily excite unmodelled<br />

dynamics in the <strong>system</strong>. Typically, this dynamics resides in the high frequency area, where<br />

an inverse model would, moreover, usually result in high amplification.<br />

The feedback part should take care <strong>of</strong> uncertainties such as unmodelled dynamics and<br />

other external disturbances and, if necessary, stabilisation <strong>of</strong> the plant using the error, e, the<br />

difference between the measured output, y, and the reference, r. Where feedforward can<br />

be seen as specific shaping <strong>of</strong> the plant w.r.t. reference signals only, the feedback shapes<br />

the sensitivity <strong>of</strong> the plant towards all disturbances. Feedback usually results in a frequency<br />

area (<strong>of</strong>ten the lower frequencies) where the <strong>control</strong>led plant becomes less sensitive, an area<br />

where sensitivity is higher (around the bandwidth) and an area with unchanged characteristics<br />

(high frequency area) [89].<br />

De Roover argues in [30] that a third degree <strong>of</strong> freedom in the design <strong>of</strong> the <strong>control</strong>ler<br />

should be considered. This should take care <strong>of</strong> the generation <strong>of</strong> an appropriate reference<br />

signal. In case <strong>of</strong> the <strong>flight</strong> simulation <strong>motion</strong> <strong>control</strong>, the causality problem in the mechanics<br />

is <strong>of</strong>ten overcome by generating an acceleration reference signal instead <strong>of</strong> position<br />

only. Calculating the forces required given a position reference would lead to a non-causal<br />

transfer function. This is not the case with acceleration. It is questionable, whether an acceleration<br />

reference signal is sufficient to fully overcome the causality problem since this<br />

would imply that immediate force generation is possible.<br />

1.2.2 Motion <strong>control</strong> specifications<br />

The main task to be accomplished after wash-out, tracking <strong>of</strong> the reference acceleration<br />

signals should not only be accomplished w.r.t. the asymptotic behaviour at infinitely long<br />

time spans but especially the transient behaviour, the generation <strong>of</strong> appropriate onset mo-


20 1 Introduction<br />

-<br />

C1<br />

r + e +<br />

- e - C2 - e?<br />

u<br />

y<br />

- G<br />

-<br />

+<br />

6<br />

Fig. 1.8: Two degree <strong>of</strong> freedom <strong>control</strong> structure<br />

tion, is most relevant. By using (model) knowledge <strong>of</strong> the reference generating <strong>system</strong>, the<br />

disturbed aircraft, the pilot and the plant, the <strong>motion</strong> <strong>system</strong>, use can be made <strong>of</strong> the two<br />

additional degrees <strong>of</strong> freedom in the <strong>control</strong>ler to enhance performance [32].<br />

As acceleration is the perceived unit, the performance <strong>of</strong> such <strong>motion</strong> <strong>control</strong> is much<br />

more sensitive as (disturbance) signals become faster, as compared to <strong>system</strong>s with position<br />

as cost variable. Attaining higher bandwidths than required can make the <strong>system</strong><br />

unnecessarily sensitive. Though it is known that below approximately 0.1 Hz <strong>motion</strong> is not<br />

perceived through inertial movement, the required bandwidth is yet unknown. Of course,<br />

the crossover frequency area <strong>of</strong> the pilot/aircraft loop around 1 Hz should be simulated with<br />

close correspondence but the question <strong>of</strong> whether the vestibular organs, which are sensitive<br />

up to 30 Hz, should be stimulated over this whole frequency area is not seen to be<br />

answered in literature. It can be expected that striving for a bandwidth <strong>of</strong> 30 Hz will lead<br />

to compromising false cues due to parasitic dynamics <strong>of</strong> the <strong>motion</strong> <strong>system</strong> with required<br />

cues.<br />

In summary, inertial <strong>motion</strong> <strong>control</strong> <strong>of</strong> the cockpit in <strong>flight</strong> <strong>simulator</strong>s is directed at<br />

providing appropriate cues in the following areas:<br />

- The onset <strong>of</strong> the aircraft response to the actions <strong>of</strong> the pilot. Both shape (amplitude) and<br />

timing (phase) <strong>of</strong> the response is important here. As the bandwidth <strong>of</strong> the pilot usually<br />

does not exceed 1 Hz, in a first attempt, a bandwidth <strong>of</strong> an order <strong>of</strong> a magnitude higher<br />

can be strived for in this respect.<br />

- A realistic generation <strong>of</strong> disturbances resulting from e.g. turbulence. At higher frequencies,<br />

above 2-3 Hz, the shape <strong>of</strong> these disturbances becomes most important while<br />

(linear) phase lags, small time delays, are not so relevant.<br />

- The dynamics related to the aircraft resulting in vibrations. This can be both due to the<br />

pilot in the loop as the disturbances. This dynamics can change drastically with the<br />

situation e.g. on the ground or airborne. This also means sufficient suppression or at<br />

least minimal excitation <strong>of</strong> <strong>simulator</strong> related dynamics i.e. parasitic dynamics, which<br />

do not correspond with the pilot in-the-loop with the actual aircraft.<br />

- Given feasible reference trajectories, the position <strong>control</strong> should be such that errors in


1.2 Motion <strong>control</strong> 21<br />

the actuator position do not result in running the actuators out-<strong>of</strong>-stroke. The corrective<br />

action should not lead to false cues and should therefore typically be <strong>of</strong> a low<br />

frequency nature and thus have very limited bandwidth.<br />

Not only the quality <strong>of</strong> the <strong>motion</strong> <strong>control</strong> is important, but also the predictability. In doing<br />

pilot/aircraft interaction experiments or in design <strong>of</strong> a training programm, it is relevant<br />

to know to what degree the simulated environment corresponds with the actual situation.<br />

Finally, the main requirement in doing piloted simulation experiments is safety, which demands<br />

stability <strong>of</strong> the <strong>motion</strong> <strong>control</strong> loop at all time.<br />

1.2.3 Conventional <strong>simulator</strong> <strong>motion</strong> <strong>control</strong><br />

Most <strong>motion</strong> <strong>control</strong> schemes, reported on in <strong>flight</strong> simulation, focus on the kinematics <strong>of</strong><br />

the mechanical <strong>system</strong> and the dynamics <strong>of</strong> independent hydraulically driven masses [12,<br />

24, 46, 51, 82]. Given a desired <strong>simulator</strong> trajectory, the required actuator trajectories can be<br />

calculated with the inverse kinematics (end-effector to actuator coordinate transformation)<br />

and with a fully parallel driven robot these relations can be given explicitly. By neglecting<br />

the interaction between the actuators, each feedback loop can be designed given the transfer<br />

function <strong>of</strong> a hydraulic actuator driving a mass. The linearised version from valve input,<br />

i, to position, q, as in [99, 153] consists <strong>of</strong> a lightly damped second order <strong>system</strong> in series<br />

with an integrator. This is described in the frequency domain as<br />

q<br />

i =<br />

K<br />

s(s2 =! 2 o +2 s=!o<br />

� (1.3)<br />

+1)<br />

considering s = j!.<br />

So at low frequencies the <strong>system</strong> acts as a velocity generator and at higher values the<br />

resonant p characteristics <strong>of</strong> second order <strong>system</strong> plays a role with the eigenfrequency, ! o =<br />

c=m, which depends on the driven mass, m, and the limited oil stiffness, c, <strong>of</strong>ten with<br />

poor damping,


22 1 Introduction<br />

the characteristics as described for the one degree-<strong>of</strong>-freedom <strong>system</strong>. Pressure feedback,<br />

therefore, still results in damping only. As the eigenfrequencies can vary considerably for<br />

each direction and <strong>simulator</strong> position, the relative damping provided by pressure feedback<br />

will differ. The positional feedback will have to be tuned according to the lowest occurring<br />

<strong>system</strong> eigenfrequency. The resulting <strong>control</strong>led <strong>system</strong> has a bandwidth corresponding to<br />

this frequency and is <strong>of</strong>ten somewhat lower since sufficient damping <strong>of</strong> the faster modes can<br />

require overdamped characteristics <strong>of</strong> the slowest.<br />

By using the acceleration reference signals as lead feedforward tuned according to a<br />

neutral <strong>simulator</strong> position 2 [24, 46], the bandwidth up to which the <strong>system</strong> reacts properly<br />

to these references can be extended. In principle, the second degree <strong>of</strong> freedom in the <strong>control</strong>ler<br />

is used to create a local inverse <strong>of</strong> the <strong>control</strong>led <strong>system</strong> up to a frequency somewhat<br />

higher than the bandwidth attained with feedback.<br />

Although the parallel structure used in <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s requires some special<br />

consideration, the <strong>system</strong> can be viewed as a robotic manipulator for which a wide field<br />

<strong>of</strong> literature exists on <strong>control</strong> strategies, which make more extensive use <strong>of</strong> <strong>system</strong>/model<br />

knowledge. Not until very recently, this is seen to be applied in <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s<br />

[20, 58, 82]. It seems worthwhile to review the robot <strong>control</strong> strategies, which can be <strong>of</strong> use.<br />

1.2.4 Robot <strong>control</strong><br />

Although the dynamics <strong>of</strong> robot manipulators is highly nonlinear, a very specific structure<br />

can be recognised. Most robot <strong>control</strong> strategies take advantage <strong>of</strong> this structure. In [136], a<br />

selection <strong>of</strong> articles provides a state <strong>of</strong> the art and an overview is given <strong>of</strong> the developments<br />

in robot <strong>control</strong> over the years. There are also numerous textbooks on the subject. In recent<br />

works like [11, 31, 81, 102, 113, 115], modern issues such as passivity, robustness, tracking,<br />

elasticity and flexibility, are treated.<br />

Computed torque<br />

One <strong>of</strong> the earliest applications <strong>of</strong> nonlinear <strong>control</strong> <strong>of</strong> robot manipulators was the method <strong>of</strong><br />

computed torque, which started in the early seventies. Using the Euler-Lagrange equations<br />

<strong>of</strong> a rigid robot manipulator with n input torques, and output positions, q given by<br />

M (q)q + C(q� _q)_q +g(q) = � (1.5)<br />

the required input torque can be calculated, compensating the coriolis/centripetal terms, C,<br />

and gravity, g, measuring the positions, q, and velocities, _q, and filling in a new input vector,<br />

a, instead <strong>of</strong> q, to end up with<br />

M (q)q = M (q)a: (1.6)<br />

2This position usually corresponds with all the actuators half stroke and is the position to which the wash-out<br />

fades


1.2 Motion <strong>control</strong> 23<br />

Since the inertia matrix, M, can be inverted, linear, parallel (decoupled) double integrator<br />

<strong>system</strong>s result from a to q,<br />

q = 1<br />

s 2 In n a (1.7)<br />

The double integrator <strong>system</strong>s can easily be <strong>control</strong>led with any desired bandwidth by a<br />

simple proportional/derivative (PD) <strong>control</strong>ler.<br />

a = Kpe + Kd _e + q d<br />

(1.8)<br />

with the positional error, e =(qd ; q), the desired positions, qd, can be tracked with step<br />

and ramp inputs.<br />

In [77], several <strong>control</strong> strategies like the computed torque method, feedback linearisation,<br />

resolved acceleration method and inverse dynamics are shown to be very similar and<br />

unifiable. In all cases, exact linearisation and decoupling is strived for. This can be done<br />

both in end effector or joint coordinates.<br />

The main disadvantages <strong>of</strong> the computed torque method are its robustness properties.<br />

Both with inexact cancellation due to parameter uncertainties (e.g. an inexactly known inertia<br />

matrix) [9] and unmodelled dynamics, such as joint elasticity [85], this <strong>control</strong>ler can<br />

go unstable.<br />

Robust robot <strong>control</strong><br />

Robustness can be regained by adding an outer loop such as in the sliding mode <strong>control</strong><br />

<strong>of</strong> Slotine and Li [133] or the saturating <strong>control</strong>ler applied by Spong [135]. Both methods<br />

employ local high gain, which is sufficiently high to track in the presence <strong>of</strong> uncertainties<br />

but on the other hand is bounded. With the discontinuous sliding mode feedback method,<br />

high frequency unmodelled dynamics is easily excited and chattering is <strong>of</strong>ten reported in<br />

practice. Transient response in saturating <strong>control</strong> is worse. Both properties do not seem<br />

most suitable for <strong>flight</strong> simulation <strong>motion</strong> <strong>control</strong>.<br />

Yet another robustifying approach is to compensate for the main nonlinearities and consider<br />

the <strong>system</strong> linear with some, possibly varying, uncertainty as was done by Van de Linden<br />

[146]. Robust <strong>control</strong> techniques developed for linear <strong>system</strong>s as described in textbooks<br />

as [132, 160] can then be used. Main problem is to arrive at a nonconservative description<br />

<strong>of</strong> an uncertainty in the nonlinear plant at the level <strong>of</strong> the outer loop.<br />

Passive robot <strong>control</strong><br />

Alternatively, use can be made <strong>of</strong> the passivity (’no energy generating’) property <strong>of</strong> a robotic<br />

<strong>system</strong> [11, 14]. Feedback with a passive <strong>control</strong>ler results in a passive closed loop <strong>system</strong>,<br />

which under some assumptions has specific stability properties. The feedback <strong>control</strong>ler can<br />

e.g. have a PD structure with possibly time varying gains (dependent on estimated mass matrix).<br />

W.r.t. feedforward, the computed torque structure can still be employed. This strategy<br />

even generalises to a much larger class <strong>of</strong> <strong>system</strong>s [147] among which are the mechanical<br />

<strong>system</strong>s with possibly flexible structures [63, 66].<br />

Most textbooks on robotic <strong>system</strong>s deal with rigid body structures only. The complexity<br />

<strong>of</strong> the <strong>system</strong> increases drastically in including deformations. The flexible structure has


24 1 Introduction<br />

more degrees <strong>of</strong> freedom than <strong>control</strong> inputs. Trajectory tracking <strong>of</strong> a robot end-effector becomes<br />

very difficult and requires accurate models [31]. The eigenfrequency <strong>of</strong> the flexible<br />

mode is usually a hard constraint for the attainable performance such as bandwidth. With<br />

passive <strong>control</strong>, though stable, performance is <strong>of</strong>ten hard to specify.<br />

Actuator dynamics<br />

Also passivity (<strong>control</strong>) <strong>of</strong> robot manipulators with dynamics in the actuators is addressed<br />

as by Arimoto [11]. The DC-actuator employed there is equivalent with a linearised version<br />

<strong>of</strong> a hydraulic actuator [137]. Although the structure <strong>of</strong> the (voltage driven) DC-motor and<br />

the hydraulic actuator may be equivalent, the parameters usually vary largely. Therefore,<br />

the DC-motor can <strong>of</strong>ten be considered relatively fast in comparison with the robot mechanics.<br />

On the contrary, the (finite oil stiffness <strong>of</strong> the) hydraulic actuator forms a relevant part<br />

<strong>of</strong> the dynamics in e.g. a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> around the achievable bandwidth.<br />

The influence <strong>of</strong> the mechanics can nominally be decoupled in the hydraulics by local<br />

velocity compensation [127]. By an inner feedback loop the hydraulic force generation<br />

becomes relatively fast compared to the mechanics. This cascaded approach was successfully<br />

implemented by Heintze and Van Schothorst [48, 124]. In another cascaded approach<br />

suggested by Qu and Dawson [113], actuator-level robust <strong>control</strong> is achieved through a<br />

recursive or backstepping design [78]. Though this method is going through rapid development<br />

[128], not much practical applications have seen to be reported yet.<br />

Parallel <strong>system</strong>s<br />

Parallel driven <strong>system</strong>s, such as the <strong>flight</strong> <strong>simulator</strong> Stewart platform <strong>motion</strong> <strong>system</strong> in<br />

Fig. 1.2, are usually not discussed in considering robotic <strong>system</strong>s. Only recently, in the late<br />

eighties, the advantages <strong>of</strong> parallel driven robotic manipulators were recognised and now<br />

become more widely applied in industry [5, 50, 96]. Such <strong>system</strong>s are inherently stiffer<br />

and due to the parallel actuation, less actuator weight has to be moved, which makes them<br />

suitable for fast assembly lines.<br />

The computed torque <strong>control</strong> laws for Stewart platforms presented by Liu et al. in [83],<br />

reveal some <strong>of</strong> the additional problems with application <strong>of</strong> model <strong>based</strong> <strong>control</strong>ler for the<br />

parallel <strong>system</strong>s. The linearising <strong>control</strong> in end-effector coordinates requires matrix inversion<br />

(<strong>of</strong> the jacobian) and the forward kinematical problem has to be solved on-line.<br />

This means computing the platform position <strong>based</strong> on joint measurements, which is easy<br />

in serial robots. No closed form explicit solution, however, has seen to be found for this<br />

problem for Stewart platforms. Choice <strong>of</strong> linearising in joint coordinates does not solve<br />

any <strong>of</strong> these problems but actually adds the requirement <strong>of</strong> calculating derivative jacobian<br />

matrices (solved in [38] by Dutre et al.).<br />

Considering general forms <strong>of</strong> parallel or closed chain mechanisms, the situation becomes<br />

even more difficult as discussed by Ghorbel in [44], where several parallel mechanisms<br />

and <strong>control</strong> strategies are classified. In some cases, it is not possible to describe such<br />

a <strong>system</strong> with a minimal number <strong>of</strong> coordinates globally and due to this fact and the specific<br />

singularities, which only appear in parallel <strong>system</strong>s, global stability results are generally<br />

hard to derive, especially for tracking problems.<br />

Many <strong>control</strong> strategies presented for parallel <strong>system</strong>s are designed for appropriate in-


1.3 Problem statement 25<br />

teraction with an environment. Force or hybrid position/force <strong>control</strong> and creating a passive<br />

compliance are <strong>of</strong>ten strived for [96, 103, 110, 116, 142]. Another field <strong>of</strong> application is the<br />

area <strong>of</strong> active vibration <strong>control</strong>, mostly in space structure applications [42, 108, 93]. This<br />

requires high bandwidth <strong>control</strong>lers, which, however, have very limited stroke and thus do<br />

not have to include typical position dependent nonlinearities.<br />

Summarising, in principle, most <strong>of</strong> the robot <strong>control</strong> strategies take advantage <strong>of</strong> the<br />

specific structure in the dynamics <strong>of</strong> mechanical <strong>system</strong>s. Though not <strong>of</strong>ten seen, the model<br />

<strong>based</strong> schemes are applicable to the parallel driven <strong>flight</strong> simulation <strong>motion</strong> <strong>system</strong>s, but<br />

all require some modifications in analysis and design. The main mechanical nonlinearities<br />

can be compensated for by a computed torque strategy, possible feedforwarded. Further,<br />

several alternatives exist to robustify the feedback loop. An important source <strong>of</strong> additional<br />

relevant dynamics can be expected from the hydraulic servo actuators, which are <strong>of</strong> main<br />

consideration in the conventional <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>control</strong> and usually not taken into<br />

account in robot <strong>control</strong>.<br />

1.3 Problem statement<br />

Due to the rapid progress in all kinds <strong>of</strong> <strong>flight</strong> <strong>simulator</strong> sub<strong>system</strong>s such as simulation<br />

computing power, image generating <strong>system</strong>s, etc., also enhanced quality <strong>of</strong> <strong>flight</strong> <strong>simulator</strong><br />

inertial <strong>motion</strong> generation is desired. At least, high fidelity <strong>motion</strong> generation would enable<br />

research into the specific requirements <strong>of</strong> <strong>motion</strong> simulation, which are still not exactly<br />

known. In similar <strong>system</strong>s, such as robotic manipulators, advanced model <strong>based</strong> <strong>control</strong><br />

strategies exist, which can attain higher performance than less structured methods, provide<br />

more insight into the <strong>system</strong>s limitations and possibly point at constructional properties,<br />

which can be taken advantage <strong>of</strong>. In <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>control</strong>, application <strong>of</strong> such<br />

methods has been almost absent. One <strong>of</strong> the reasons might be that the <strong>motion</strong> <strong>system</strong>s have<br />

a degree <strong>of</strong> complexity, which does not allow exact modelling or detailed models as part <strong>of</strong><br />

the <strong>control</strong>ler. Application requires an intermediate step <strong>of</strong> extracting a reasonably detailed<br />

model, which describes the most relevant dynamic characteristics.<br />

1.3.1 General problem statement<br />

The foregoing leads to the following problem statement for this research:<br />

Investigate what relevant <strong>system</strong> knowledge should be used in a model <strong>based</strong> <strong>control</strong><br />

strategy and to what extent does this improve the <strong>control</strong>led dynamics <strong>of</strong> a <strong>flight</strong><br />

<strong>simulator</strong> <strong>motion</strong> <strong>system</strong>.<br />

1.3.2 Structuring the problem statement;<br />

approach in research<br />

To solve this problem, the full design process has to be defined, structured and evaluated.<br />

This process starts with modelling and analysis, proceeds with <strong>control</strong> synthesis and finally<br />

testing <strong>of</strong> the <strong>system</strong>. Many solutions and procedures <strong>of</strong> the steps to be taken are given


26 1 Introduction<br />

in literature, but an integral approach, having the specific properties and requirements <strong>of</strong><br />

<strong>flight</strong> simulation <strong>motion</strong> generation in mind, still lacks. In this research, integratibility and<br />

applicability will be the main arguments in choosing the most suitable alternative or, if<br />

necessary, newly proposed variant, <strong>of</strong> each step in the <strong>system</strong>/<strong>control</strong> design procedure.<br />

Since the overall investment is high and the fidelity <strong>of</strong> the <strong>motion</strong> <strong>system</strong> is considered<br />

important, the full <strong>motion</strong> <strong>control</strong> design process <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> is allowed a fairly<br />

elaborate procedure optimised towards the specific properties <strong>of</strong> the <strong>system</strong> at hand, as<br />

compared to the industrial robot <strong>control</strong> design. Only, <strong>of</strong> course, if this leads to enhanced<br />

quality <strong>of</strong> the resulting <strong>system</strong>.<br />

In the <strong>control</strong> design procedure employed in this research, the following fairly standard<br />

subproblems in application directed <strong>system</strong>s and <strong>control</strong>, will be distinguished:<br />

- Obtain a relevant model structure for analysis and <strong>control</strong> through physical modelling.<br />

- Evaluate the structure by experiments, identify model parameters and provide boundaries<br />

for the extent to which the model validity holds.<br />

- Choose and, if necessary, modify the most appropriate model <strong>based</strong> robot <strong>control</strong> strategy,<br />

given the <strong>system</strong>s properties and requirements. Design and implement this <strong>control</strong>ler<br />

on the <strong>motion</strong> <strong>system</strong> <strong>of</strong> the Simona <strong>flight</strong> <strong>simulator</strong>.<br />

- Define a test to quantify performance <strong>of</strong> the <strong>control</strong>led <strong>system</strong> and use this test to compare<br />

with the usual <strong>control</strong> design approach.<br />

Related to the <strong>flight</strong> simulation application and the properties <strong>of</strong> the <strong>motion</strong> <strong>system</strong>s used,<br />

specific elements have to be emphasised in each <strong>of</strong> these steps.<br />

Physical modelling<br />

<strong>Model</strong>ling according to physical relations such as balance relations i.e. force balance equations<br />

in mechanics and flow balance equations in hydraulics, has two important advantages<br />

as compared to the so-called black box experimental identification. First, as the structure<br />

<strong>of</strong> the model points at the underlying physical properties <strong>of</strong> the <strong>system</strong>, analysis and (constructional)<br />

design considerations are performed much easier. Secondly, modelling <strong>of</strong> a<br />

nonlinear <strong>system</strong> such as the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> platform, does not impose additional<br />

difficulties.<br />

As already mentioned, the mechanics and hydraulics are considered the main sub<strong>system</strong>s<br />

to be taken into account. These sub<strong>system</strong>s specifically, and the integration <strong>of</strong> the<br />

two, will form the basis <strong>of</strong> the models to be derived. The mechanical <strong>system</strong> is, even if<br />

the parts are considered rigid, relatively complex. This complexity mainly stems from the<br />

parallel kinematical structure, which, however, has its influence on the <strong>system</strong>s dynamics.<br />

The kinematics <strong>of</strong> parallel <strong>system</strong>s, and <strong>of</strong> the Stewart platform specifically, will have to be<br />

analysed.<br />

Nowadays, the equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> complex rigid body <strong>system</strong>s can <strong>of</strong>ten be generated<br />

automatically in symbolic form. This way <strong>of</strong> working will be evaluated as opposed to<br />

the derivation ’by hand’ for this <strong>system</strong>.<br />

The models describing the dynamics <strong>of</strong> the mechanical <strong>system</strong> will have to be evaluated<br />

for use in a real-time model <strong>based</strong> <strong>control</strong> structure. Further, it will have to be analysed, how


1.3 Problem statement 27<br />

choices in the design <strong>of</strong> the construction resulting in the specific kinematical and inertial<br />

(mass) properties, influence the dynamics.<br />

The hydraulic servo actuators <strong>of</strong> the Simona <strong>motion</strong> <strong>system</strong> have been modelled and<br />

evaluated in detail in [124] by Van Schothorst. In this research, only models will be discussed,<br />

which describe the most relevant dynamics for <strong>control</strong>. Finally, the properties <strong>of</strong> the<br />

integrated hydraulically driven mechanical <strong>system</strong> will have to be analysed.<br />

Experimental validation<br />

Since the kinematical structure <strong>of</strong> the <strong>system</strong> forms the basis <strong>of</strong> the dynamical properties <strong>of</strong><br />

this <strong>motion</strong> <strong>system</strong>, some attention will have to be paid to the validity <strong>of</strong> the structure and the<br />

identification <strong>of</strong> the parameters to be used. An, ’after construction’, calibration procedure<br />

will be evaluated as opposed to setting tight (and <strong>of</strong>ten expensive) tolerance specifications<br />

in fabrication.<br />

Experimental identification procedures will have to be developed and applied to reconstruct<br />

the parameters such as dissipative terms, which are difficult to predict in advance,<br />

and evaluate others such as the inertial properties. The models will be evaluated through<br />

comparison with the experimental data.<br />

The area <strong>of</strong> validity <strong>of</strong> the models is also an important aspect, which has to be checked.<br />

The cause and severity <strong>of</strong> possible parasitic effects will be investigated. Especially flexibility<br />

is known to constrain the achievable <strong>control</strong> performance and difficult to predict.<br />

Control strategy<br />

To maintain structure in a model <strong>based</strong> <strong>control</strong>ler for a complex <strong>system</strong> such as the <strong>motion</strong><br />

<strong>system</strong> under consideration and to let a <strong>control</strong>ler result, which can be implemented on a<br />

real-time application, modularization <strong>of</strong> the tasks is desired. The physical structure <strong>of</strong> the<br />

<strong>system</strong> provides a reasonable way to achieve this with a multi level <strong>control</strong>ler in which for<br />

each level specific tasks are defined in close relation with the other levels.<br />

It should be possible to turn the hydraulic actuators into fast and smooth force generators<br />

by local inner loop pressure <strong>control</strong> as done by Heintze in [48]. The applicability <strong>of</strong> this<br />

design step, which e.g. assumes accurate velocity measurement, should be evaluated.<br />

Any standard robot <strong>control</strong> strategy, which assumes direct torque, can be put on top <strong>of</strong><br />

this lower level. As first basic step, the computed torque schemes will have to be made fit for<br />

use in parallel <strong>system</strong>s. This mainly concerns the choice <strong>of</strong> an appropriate set <strong>of</strong> coordinates<br />

and evaluation <strong>of</strong> safe (convergent and stable) and sufficiently fast reconstruction <strong>of</strong> these<br />

coordinates.<br />

In the outer level, the two degrees <strong>of</strong> freedom, smooth reference feedforward, possibly<br />

taking reference model knowledge into account, and position feedback for stabilisation and<br />

preventing the <strong>system</strong> from running out <strong>of</strong> stroke, should not interfere with each other in<br />

performing their basic task.<br />

Performance evaluation<br />

In evaluation <strong>of</strong> the model <strong>based</strong> <strong>control</strong> strategy the feedback oriented conventional decentralised<br />

<strong>control</strong>ler will be taken as a reference. It is understood that, in practise, this<br />

conventional <strong>control</strong>ler can be improved by design <strong>of</strong> an additional feed forward path, but


28 1 Introduction<br />

this usually leads to local improvement only and is either an ad hoc approach or takes a<br />

similar strategy as in the more structured model <strong>based</strong> <strong>control</strong> design.<br />

Evaluation test procedures exist [1], but are outdated and should be modified to current<br />

standards <strong>of</strong> <strong>control</strong> performance quantification. In this research, the following approach<br />

will be taken. By assuming the high fidelity <strong>control</strong>ler being capable <strong>of</strong> achieving a reasonably<br />

linear <strong>system</strong> response, the frequency response measurements are considered most<br />

important. A set <strong>of</strong> other tests is mainly developed to evaluate the assumption <strong>of</strong> closed<br />

loop linear <strong>system</strong> response.<br />

Further, it will be observed that the existing evaluation procedures have two main limitations.<br />

First, they have a local character and consequently do not show whether the <strong>control</strong>ler<br />

does adequately respond to the nonlinear characteristics <strong>of</strong> the <strong>system</strong> over the workspace.<br />

Secondly, the <strong>system</strong> is not evaluated performing its actual task, generating inertial <strong>motion</strong><br />

cues in <strong>flight</strong> simulation manoeuvres. A modified test will, as well as possible, have to<br />

overcome these limitations.<br />

1.4 Outline <strong>of</strong> the thesis<br />

The thesis roughly follows the line <strong>of</strong> the reasoning for the approach sketched in the preceding<br />

section, to go through the full <strong>control</strong>ler design process. Subsequently the subproblems<br />

in modelling, identification, <strong>control</strong> and evaluation <strong>of</strong> the <strong>motion</strong> <strong>system</strong>/<strong>control</strong> design with<br />

the <strong>flight</strong> simulation application in mind, will be treated.<br />

Physical modelling<br />

By considering the laws <strong>of</strong> physics, dynamical models <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

will be obtained. In Chapter 2, the mechanical part will be discussed. Specifically, attention<br />

will be paid to the kinematics and dynamics <strong>of</strong> a parallel rigid body <strong>system</strong>. In Chapter 3,<br />

the basic structure <strong>of</strong> the hydraulic servo actuators and extensions to this model will first be<br />

discussed. The integration <strong>of</strong> the hydraulic mechanical <strong>system</strong>, the full model <strong>of</strong> the <strong>motion</strong><br />

<strong>system</strong> and its properties are the subject <strong>of</strong> Section 3.3 and its specifics in a Stewart platform<br />

construction are pointed at in Section 3.4.<br />

’Open loop’ experimental verification<br />

Chapter 4 deals with the experimental validation <strong>of</strong> the physical models, the identification<br />

<strong>of</strong> model parameters and the parasitic ’additional’ dynamics. The identification <strong>of</strong> the<br />

static parameters in the kinematical structure, calibration through redundant measurements<br />

is discussed in Section 4.1. Dynamic (inertial) parameter identification by modal analysis<br />

techniques is the subject <strong>of</strong> Section 4.2. Frequency response measurements are the main<br />

ingredients in the model validation considered in Section 4.3. This includes getting track<br />

<strong>of</strong> the parasitic dynamics caused by flexibility in the <strong>system</strong> and its environment in Section<br />

4.3.2, which concludes the discussion on the ’open loop’ experiments taken in this<br />

chapter.<br />

Control strategy<br />

In Chapter 5, the <strong>control</strong> strategy is built up by defining the general approach for multi level


1.4 Outline <strong>of</strong> the thesis 29<br />

<strong>control</strong> in Section 5.3 after a short introduction to the main <strong>control</strong> aspects in Section 5.1<br />

and Section 5.2. Then, each part <strong>of</strong> this <strong>control</strong>ler will be treated. In Section 5.4, the local<br />

pressure <strong>control</strong> <strong>of</strong> the hydraulic servo <strong>system</strong>s is outlined. Section 5.5 discusses the computed<br />

torque <strong>control</strong> <strong>control</strong> level to compensate for the multivariable, nonlinear mechanics.<br />

This requires an iterative coordinate reconstruction and is followed by the feedback path defined<br />

by the problem <strong>of</strong> position stabilisation. Interconnection with the trajectory generating<br />

<strong>system</strong> with reference model <strong>based</strong> <strong>control</strong>, treated in Section 5.6, defines the forward path<br />

at the outer level.<br />

Closed loop experimental evaluation<br />

The evaluation <strong>of</strong> the <strong>control</strong>ler on the actual experimental set-up is the subject <strong>of</strong> the second<br />

part <strong>of</strong> Chapter 5 starting with discussing the implementational issues in Section 5.7. First,<br />

the procedure to test the performance <strong>of</strong> the <strong>control</strong>led <strong>system</strong> is defined in Section 5.8.<br />

Then, this procedure is applied to the Simona <strong>motion</strong> <strong>system</strong> with the newly proposed <strong>control</strong>ler,<br />

which is discussed in Section 5.9. As a reference, comparing with a conventional<br />

<strong>control</strong>ler strategy.<br />

Final considerations<br />

The main aspects in the design <strong>of</strong> <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> <strong>control</strong>lers evaluating the<br />

experience <strong>of</strong> going through the full <strong>system</strong>/<strong>control</strong> design process defined are treated in<br />

Chapter 6 giving a review and discussion on the results. Directed at the problem statement,<br />

defined in Section 1.3, the final Chapter 7 states conclusions towards the opportunities given<br />

in this research by the application <strong>of</strong> a model <strong>based</strong> <strong>control</strong>ler in <strong>flight</strong> simulation <strong>motion</strong><br />

<strong>control</strong>. Also recommendations are provided.


Chapter 2<br />

Mechanics <strong>of</strong> parallel driven<br />

<strong>motion</strong> <strong>system</strong>s<br />

The, fully parallel, hydraulically driven construction <strong>of</strong> a Stewart platform is <strong>of</strong>ten applied<br />

for <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s. In this research, the use <strong>of</strong> model <strong>based</strong> <strong>control</strong> for such<br />

<strong>system</strong>s is investigated. In this chapter, the first step in modelling and analysis is performed<br />

by considering the mechanics.<br />

In general, models <strong>of</strong> parallel robots result in combined algebraic and differential equations,<br />

which causes difficulties with simulation, analysis and model <strong>based</strong> <strong>control</strong>. This<br />

problem will be stated in discussing some literature on parallel manipulators and modelling<br />

<strong>of</strong> mechanical <strong>system</strong>s in Section 2.1.<br />

In Section 2.2, first a short introduction into basic kinematics will be given. Then, the<br />

structure <strong>of</strong> kinematic parallel and serial singularities will be treated. The, less standard,<br />

parallel singularity causes local loss <strong>of</strong> <strong>control</strong>lability and observability <strong>of</strong> the mechanical<br />

<strong>system</strong> and should be avoided. In Section 2.3, a new method for the Stewart platform is<br />

introduced to ensure exclusion <strong>of</strong> these points from the working area <strong>of</strong> the manipulator.<br />

Moreover, this is also a necessary condition for guaranteed sufficiently fast convergence <strong>of</strong><br />

an iterative scheme in reconstructing the platform coordinates. For the <strong>motion</strong> <strong>system</strong> at<br />

hand, sufficient accurate real time convergence is proven, which enables safe use <strong>of</strong> this<br />

scheme in model <strong>based</strong> <strong>control</strong> later on in Chapter 5.<br />

The platform coordinates are a proper choice in parametrization <strong>of</strong> the Stewart platform,<br />

thereby circumventing difficulties with combined algebraic and differential equations. In<br />

Section 2.4, which deals with the dynamics <strong>of</strong> the mechanical part <strong>of</strong> the <strong>motion</strong> <strong>system</strong>,<br />

it is shown that an explicit differential model results from this choice. Thereby, the use <strong>of</strong><br />

a suitable projection method, following Kane [64], is exploited to arrive at the equations<br />

<strong>of</strong> <strong>motion</strong>. In this way, a proper starting point for simulation, analysis, model reduction<br />

and model <strong>based</strong> <strong>control</strong> is provided for. Some extensions, e.g. parasitic modes <strong>of</strong> the<br />

foundation, are also included.<br />

31


32 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

2.1 Parallel <strong>motion</strong> <strong>system</strong>s<br />

The use <strong>of</strong> robot manipulators is widely spread in industry nowadays. Most <strong>of</strong> these manipulators<br />

are constructed as a series connection <strong>of</strong> joints and links. The dual form <strong>of</strong> these<br />

robots, the parallel manipulator, is less <strong>of</strong>ten seen to be applied. In (<strong>flight</strong>)simulation <strong>motion</strong><br />

<strong>system</strong>s however, the parallel construction is almost invariably in use. The Stewart platform<br />

is a 6 degrees-<strong>of</strong>-freedom (d.o.f.) parallel manipulator, which is applied in most <strong>of</strong> the current<br />

high fidelity <strong>flight</strong> <strong>simulator</strong>s. It was named after Stewart [140] who illustrated the use<br />

<strong>of</strong> such a parallel structure for <strong>flight</strong> simulation in 1965. Nowadays, it is also referred to<br />

as Gough-platform, since it was Gough who presented the practical use <strong>of</strong> such a <strong>system</strong><br />

somewhat earlier in 1949 rediscovering the structure already described around 1800 by the<br />

mathematician Cauchy [95]. As a manufacturing and fabrication robot it is also named a<br />

hexapod.<br />

There are several advantages in applying a parallel construction. This kind <strong>of</strong> manipulators<br />

have higher rigidity and accuracy due to the parallel force path and averaged link to<br />

end-effector error. The inverse kinematics (from end-effector to link coordinates) which is<br />

a problem in path generation <strong>of</strong> serial manipulators is easily solved in parallel robots. There<br />

are also disadvantages. The dual forward kinematics is a complex algebraic problem and<br />

has in general more than one solution [114]. <strong>Model</strong>ling the dynamics is also less straight<br />

forward.<br />

For several reasons, feedback <strong>control</strong> <strong>of</strong> these <strong>motion</strong> <strong>system</strong>s is still decentralized<br />

i.e. per actuator without taking mechanical coupling into account. Setting higher standards<br />

<strong>of</strong> <strong>motion</strong> realism in simulation will involve modern <strong>control</strong> strategies in order to fully benefit<br />

from recent constructional and computational improvements in <strong>flight</strong> <strong>simulator</strong>s e.g. light<br />

weight, low centre-<strong>of</strong>-gravity (c.o.g.) platforms, high frequency airplane dynamics simulation<br />

[6]. Also use <strong>of</strong> a Stewart platform as a more general robot, as presented by Nguyen<br />

et al. [104], will be enhanced with high performance <strong>control</strong> <strong>of</strong> <strong>motion</strong>, as these <strong>system</strong>s<br />

usually require minimal time tasks.<br />

By incorporating more structural <strong>system</strong> information i.e. a model into the <strong>control</strong>ler,<br />

it is possible to achieve higher performance. Most modern <strong>control</strong> strategies are therefore<br />

model <strong>based</strong> in some way (directly, in design or evaluation). In this case the quality <strong>of</strong><br />

<strong>motion</strong> depends on the fidelity <strong>of</strong> the model. Deriving a model <strong>of</strong> the mechanics <strong>of</strong> the<br />

Stewart platform manipulator for analysis, design and <strong>control</strong> will be the subject <strong>of</strong> this<br />

chapter. To arrive at a model with structure from which insight can be gained, modelling<br />

will be done <strong>based</strong> on physical laws.<br />

<strong>Model</strong>ling the dynamics <strong>of</strong> a Stewart platform as a multibody <strong>system</strong> has been done<br />

by Lee et al. [80] who claim to be the first to present a complete model and with more<br />

simplifications by Do et al. [36] and Liu et al. [83]. <strong>Model</strong>ling the mechanics <strong>of</strong> this platform<br />

can be done in several ways and with various objectives in mind. The equations <strong>of</strong> <strong>motion</strong><br />

can be derived by using the classical approach <strong>of</strong> Lagrange [80] or Newton-euler [36].<br />

In general, deriving the equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> a parallel manipulator results in combined<br />

differential and algebraic (constraint) equations (see e.g. Roberson and Schwertassek<br />

[120]). In simulation and <strong>control</strong> this formulation can cause difficulties, usually referred to<br />

as index problems [19]. Here it is shown that an explicit differential model for the Stewart


2.2 Kinematics 33<br />

platform results if one makes the right choices in parametrization. Dependent variables are<br />

explicit functions <strong>of</strong> the integrable differential equations. In this way index problems, etc.<br />

are circumvented.<br />

By using a modern projection or dual basis method [79] like Kane’s [64], which have<br />

the advantages <strong>of</strong> both the Newton-euler and Lagrange formulation but without the corresponding<br />

disadvantages [56], it will be shown that applying this approach can result in a<br />

model from which more insight can be gained.<br />

Together with the alternative parametrization, this is advantageous over the models earlier<br />

presented in literature, if one wants to apply a model for both analysis, simulation and<br />

<strong>control</strong>. <strong>Model</strong> <strong>based</strong> feedback, however, is still more complex for parallel manipulators,<br />

since the dynamics are only described in end-effector coordinates and the measured signals<br />

are link related. Next to the fairly low performance requirements <strong>of</strong> <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong><br />

<strong>system</strong>s in the past, this will probably be the reason that in most <strong>of</strong> these <strong>motion</strong> <strong>system</strong>s,<br />

feedback <strong>control</strong>lers are decentralized (one SISO loop per parallel link i.e. actuator).<br />

Through analysis <strong>of</strong> the derived model, some <strong>of</strong> the disadvantages <strong>of</strong> decentralized <strong>control</strong><br />

can be revealed. To apply a simple, but accurate model <strong>based</strong> <strong>control</strong>ler in practise, one<br />

would like to quantify the errors made by undermodelling, to be able to do robust analysis<br />

<strong>of</strong> the <strong>control</strong> scheme. The modelling approach taken here aims at a model from which<br />

the influences <strong>of</strong> different <strong>system</strong> parameters, like masses, inertias, velocity, gravity can be<br />

clearly separated.<br />

2.2 Kinematics<br />

After stating some notational issues, the fundamental formulas <strong>of</strong> mechanics to describe the<br />

kinematics will be introduced. Then the Stewart platform will be defined in order to derive<br />

a model <strong>of</strong> its specific kinematics. After some model analysis with the <strong>control</strong> objective in<br />

mind, this part will be concluded by a resume.<br />

2.2.1 Notation and definitions<br />

Capital symbols, X are used for matrices, x for vectors, x for scalars. With some scalar<br />

(energy) functions X is used. x y denotes the vector product which can also be written<br />

as ~ Xy = ( ~ Y ) T x where the vector product matrix, ~ X, is a skew symmetric (X = ;X T )<br />

matrix parametrized by the vector, x T = x1 x2 x3 , such that the result is the vector<br />

product.<br />

~X =<br />

2<br />

4 0 ;x3 x2<br />

x3 0 ;x1<br />

;x2 x1 0<br />

3<br />

5 (2.1)<br />

X Y denotes vector wise product <strong>of</strong> the columns stacked in the matrices.<br />

p<br />

The index xn is used for the normalising operation xn = x= j x j with j x j= xT x.<br />

Pxn denotes the projector to the (hyper)plane with normal vector x n and can be constructed


34 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 2.1: Projection <strong>of</strong> arbitrary vector, s onto the plane with normalised normal vector, x n.<br />

The projection matrix, Pxn, can be composed from the vector product (matrices,<br />

~Xn).<br />

from the vector product matrix<br />

Pxn =(I ; xnx T n )=(~ Xn) 4 = ~ Xn ~ X T n = ;( ~ Xn) 2 : (2.2)<br />

Projection matrices have some properties, which can be taken advantage <strong>of</strong>, like<br />

Pxn = P T xn = P k xn� (2.3)<br />

with k>0 and integer. In some cases it will be appropriate to use such a projection plane<br />

for construction <strong>of</strong> a frame as done in Fig. 2.1.<br />

Motion can be described w.r.t. various frames. A matrix or vector described in some<br />

frame can have a superscript referring to this frame. For the inertial frame, G, or ground<br />

coordinates, the index x g will be used. As a function <strong>of</strong> the moving end-effector or platform<br />

frame, M, vectors will be denoted x m . If a (rotation) matrix maps a vector into another<br />

frame it will be denoted as g R m if R maps from M to G.<br />

The subscript index like ai will be used to refer to the i th -actuator if also non actuator<br />

dependent variables appear in the equation.<br />

2.2.2 Calculating velocity and acceleration by differentiation<br />

An important part <strong>of</strong> kinematics is the ability to derive velocity and acceleration vectors <strong>of</strong><br />

any part <strong>of</strong> a (partly) moving construction. The basic formulas to do this will be derived in<br />

this section.<br />

The <strong>motion</strong> <strong>of</strong> a point (mass particle, joint, etc.) is usually most conveniently and invariantly<br />

defined w.r.t. the body frame to whom it’s connected. The <strong>motion</strong> <strong>of</strong> a frame put<br />

in another frame generally consists <strong>of</strong> a translation, which can be described by a vector, t,<br />

and a rotation for which a matrix, R, can be used. The orientation <strong>of</strong> a whole frame can<br />

be described by a rotation matrix. A rotation matrix consists <strong>of</strong> perpendicular unit vectors<br />

which describe the basis <strong>of</strong> the frame into the other frame. As a result a rotation matrix, R


2.2 Kinematics 35<br />

has the following property:<br />

R T R = I (2.4)<br />

Any real 3 3-matrix with this property and the property det(R) = 1 is a rotation matrix,<br />

which set is <strong>of</strong>ten referred as the SO(3)-group [77]. SO(3) stands for the Special<br />

(det(R) =1) Orthonormal (RT R = I) group <strong>of</strong> matrices <strong>of</strong> size 3x3. With det(R) =;1<br />

also the mirror operation is included (transformation <strong>of</strong> right hand frames to left hand frames<br />

and vice versa). The position <strong>of</strong> a point pm in frame M can now be described in frame G<br />

by:<br />

p g = t g + g R m p m<br />

(2.5)<br />

To describe the velocity <strong>of</strong> this point in the other frame one can simply differentiate this<br />

equation. Some properties <strong>of</strong> the time derivative <strong>of</strong> the rotation matrix can be derived by<br />

differentiating (2.4). This results in skew symmetric matrices which can be parametrized by<br />

the vector product matrix <strong>of</strong> the (thereby defined) angular velocity ! m .<br />

with<br />

~ m =<br />

R T _ R = ; _ R T R = ~ m<br />

2<br />

4 0 ;!3 !2<br />

!3 0 ;!1<br />

;!2 !1 0<br />

(2.6)<br />

3<br />

5 (2.7)<br />

Now, for the vector, which is rigidly attached to the frame M, _p m = 0. Hence, by differentiating<br />

(2.5) and using (2.6), one obtains<br />

_p g = _ t g + g R m ~ m p m = _ t g + ~ g p g = _ t g +! g<br />

p g<br />

(2.8)<br />

where the change <strong>of</strong> frame for the matrix ~ ,isgivenby ~ g = g R m ~ mm R g and using<br />

the inverse rotation given by m R g =( g R m ) T with (2.4).<br />

By differentiating (2.8) the acceleration <strong>of</strong> the point, still rigidly attached to the frame<br />

M(_p m = 0), can be calculated, using ~ P m ! m = ~ m p m :<br />

p g = t g + g R m ( ~ P m ) T _! m + g R m ( ~ m ) 2 p m<br />

= t g + ( ~ P g ) T _! g +( ~ g ) 2 p g<br />

= t g ; p g _! g +! g<br />

(! g<br />

p g ) (2.9)<br />

In other cases, the point considered can already be moving in the frame M (p m ,v m m =<br />

_p m ,a m m = pm ). Here, by differentiation <strong>of</strong><br />

_p g = _ t g + g R m ( ~ ) m p m + g R m v m p<br />

the coriolis acceleration appears as the third term in<br />

p g = a g p + g R m a m m +2g R m ~ m v m m<br />

= a g p +ag m +2~ g v g m =a g p +ag m +2(!g<br />

(2.10)<br />

v g m ): (2.11)<br />

Where p is a point connected to M momentarily at the same position as p. Its acceleration,<br />

a g p , is given by (2.9).


36 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 2.2: Physical structure <strong>of</strong> the euler angle representation.<br />

2.2.3 Parametrising orientation by euler angles<br />

In stating the equations <strong>of</strong> <strong>motion</strong>, the state which describes the orientation, usually only<br />

appears in the rotation matrix. It is possible to parametrise the rotation matrix in several<br />

ways.<br />

Although the rotation matrix consists <strong>of</strong> nine entries, its properties, given by (2.4), put<br />

constraints on these entries. The orientation, i.e. the SO(3)-group, can locally be represented<br />

by 3 parameters. It can, however, not be covered by a single coordinate chart <strong>of</strong> this<br />

size [77]. Several parametrizations are used in literature. The euler angle description consists<br />

<strong>of</strong> three subsequent simple planar rotations around momentary axes. The euler angles<br />

are not convenient for use in modelling the Stewart platform but are merely discussed for<br />

reference since it is a popular description in the aerospace community.<br />

An <strong>of</strong>ten used example <strong>of</strong> euler angle rotation is physically represented in Fig. 2.2. Both<br />

three unit vectors nG�M <strong>of</strong> the ground frame, G, and moving frame, M, are drawn. A simple<br />

x�y�z<br />

planar rotation around the x-axis has the following structure<br />

Tx =<br />

2<br />

4<br />

1 0 0<br />

0 cos( ) ; sin( )<br />

0 sin( ) cos( )<br />

A vector, x m , quantified in the axes <strong>of</strong> M can be represented in G by<br />

x g = Rx m = Tz�Ty�Tx�xm 3<br />

5 (2.12)<br />

(2.13)<br />

Since the subsequent rotations are not commutative, it is important to define the order <strong>of</strong><br />

rotation. The order defined in (2.13) is the usual order applied in the aerospace community.


2.2 Kinematics 37<br />

This means that the airplane, at first positioned so that the body axis is parallel to the fixed<br />

inertial frame, G, has to be rotated as follows:<br />

- First rotate over an angle , the yaw angle, around the inertial n g z-axis. In taxiing, this<br />

typically will be the main angle <strong>of</strong> rotation.<br />

- Then pitch over an angle around the intermediate y � -axis. This will usually be the main<br />

angle to be changed in take-<strong>of</strong>f. Note, e.g. from Fig. 2.2 that this axis is in general not<br />

an axis <strong>of</strong> the inertial or the moving frame.<br />

- Finally roll around the airplanes’ longitudinal axis, n m x , over an angle to obtain its actual<br />

orientation.<br />

The fully parametrized rotation matrix is then as follows:<br />

R =<br />

2<br />

4<br />

c c c s s ; s c s s + c s c<br />

s c c c + s s s s s c ; c s<br />

;s c s c c<br />

3<br />

5 (2.14)<br />

In this equation s = sin( ), c = cos( ) and so on. The three euler angles have the<br />

disadvantage <strong>of</strong> a highly non-linear appearance in both the rotation matrix and the euler<br />

angle velocity to angular velocity transformation. The latter can even become singular.<br />

Since, although any rotation matrix can be described by this parametrization, in some states<br />

(consider = =2 in Fig. 2.2) this chart does not allow three independent changes <strong>of</strong><br />

orientation, i.e. the kinematical relationship between =[ ] T and the angular velocity<br />

in the body frame, ! m , given by<br />

! m =<br />

2<br />

4 p<br />

q<br />

r<br />

2<br />

4 1<br />

0<br />

0<br />

3<br />

5 = ( ) _ =<br />

3 2<br />

5 _ T<br />

+ Tx0 2<br />

4<br />

4 0<br />

1<br />

0<br />

1 0 ;s<br />

0 c c s<br />

0 ;s c c<br />

3<br />

5 _ T<br />

+ Tx0T T y0 2<br />

4 0<br />

0<br />

1<br />

32<br />

5<br />

4 _<br />

_<br />

_<br />

3<br />

5 =<br />

3<br />

5 _ (2.15)<br />

becomes singular. p, q and r are the components <strong>of</strong> the angular velocity, ! m , measured in<br />

moving body frame. In an airplane this will usually be the output <strong>of</strong> the gyros.<br />

In some experiments, a rotation matrix can be identified. With ; =2 < � � < =2,<br />

the euler angles can be calculated from such rotation matrix as follows:<br />

sin( ) = ;R(3� 1) (2.16)<br />

sin( ) = R(2� 1)= cos( )<br />

sin( ) = R(3� 2)= cos( )<br />

As already said, the euler angle description <strong>of</strong> rotation is not convenient in describing the<br />

model <strong>of</strong> the Stewart platform. This is mainly because evaluation <strong>of</strong> the high number <strong>of</strong> sine<br />

and cosine functions is both analytically and numerically relatively troublesome and this can<br />

fully be avoided in modelling the <strong>motion</strong> <strong>system</strong> mechanics by using euler parameters.


38 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

G<br />

nµ<br />

n<br />

g<br />

z<br />

n<br />

n<br />

g<br />

x<br />

g<br />

y<br />

s m<br />

µ<br />

s g<br />

nµ x sm<br />

µ<br />

s<br />

s<br />

g<br />

m<br />

2sin(1/2 µ )<br />

( nµ x sm)xnµ<br />

Fig. 2.3: Visualising the euler parameter representation by a rotation around the unit vector,<br />

n , over an angle, . The lower right schematic drawing is the plane orthogonal<br />

to n through the points sg and sm.<br />

2.2.4 Parametrising orientation by euler parameters<br />

An attractive alternative <strong>of</strong> euler angles are the unit quaternions [130] called euler parameters.<br />

No singularities and, even more important, numerically convenient relationships with<br />

both the rotation matrix and the angular velocity and as with the euler angles, an interpretation<br />

which can be visualised. This at the cost <strong>of</strong> an additional fourth parameter and<br />

consequently an extra constraint equation.<br />

The orientation <strong>of</strong> one frame w.r.t. another frame can always be described by a single<br />

rotation over an angle , about a unique axis with direction n . The four euler parameters,<br />

T T<br />

, parametrise and n as follows:<br />

e = 0<br />

The constraint equation,<br />

= 1 2 3<br />

0 = cos(1=2 ) (2.17)<br />

T = sin(1=2 )n (2.18)<br />

T<br />

e e =1� (2.19)<br />

is a direct result from this definition. By not using the angle <strong>of</strong> rotation itself as a coordinate<br />

but the sine and cosine instead, these geometric functions do not have to be evaluated anymore.<br />

And as a result we will see in the subsequent sections that in a full mechanical model<br />

<strong>of</strong> the Stewart platform not one geometric function will appear.


2.2 Kinematics 39<br />

With Fig. 2.3 the rotation <strong>of</strong> a vector, sm, in the body frame, M, to its direction in the<br />

inertial frame, G, given as sg, can be described using the unit vector, n , along the axis <strong>of</strong><br />

rotation and the angle <strong>of</strong> rotation, .<br />

The following derivation starts with constructing s g from sm adding the two component<br />

vectors in the plane given in Fig. 2.3 and is directed towards the specification <strong>of</strong> the rotation<br />

matrix.<br />

sg = sm ; (1 ; cos( ))(n sm) n + sin( )(n sm)<br />

= sm + 2 sin 2 (1=2 )n (n sm) +2cos(1=2 )sin(1=2 )(n sm)<br />

= I +2~~ +2 0~ sm<br />

= 2<br />

=<br />

2<br />

2<br />

4 2<br />

0<br />

2<br />

+ 1 ; 1=2 1 2 ; 0 3 1 3 + 0 2<br />

2<br />

1 2 + 0 3 0 + 2 2 ; 1=2 2 3 ; 0 1<br />

1 3 ; 0 2 2 3 + 0 1 ; 1=2<br />

4 ; 1 0 ; 3 2<br />

; 2 3 0 ; 1<br />

; 3 ; 2 1 0<br />

3 2<br />

5<br />

2<br />

0<br />

+ 2<br />

3<br />

3<br />

5 sm<br />

4 ; 1 0 3 ; 2<br />

; 2 ; 3 0 1<br />

; 3 2 ; 1 0<br />

3T<br />

5 sm<br />

= G( e)L T ( e)sm = R( e)sm (2.20)<br />

So, rotation appears to be equal to two simple subsequent transformations, G and L,<br />

which are linear in the euler parameters. By Nikravesh et al. [107] it is shown that this<br />

parametrization and using _ RR T = ~! g leads to the following relation to calculate _ e or _<br />

from ! and e,<br />

_e = 1<br />

2 GT ( e)! g<br />

In only constructing the variable, _, a reduced version is given by<br />

(2.21)<br />

_ = 1<br />

2 GT s ( e)! g � (2.22)<br />

where the matrix, Gs, is given by the last three columns <strong>of</strong> G. This equation can be used to<br />

get from the angular velocity to the rotation matrix with help <strong>of</strong> an integration routine and<br />

initial conditions on . Further, the fairly simple relations like<br />

! g =2G( e)_ e� (2.23)<br />

! m =2L_ e and e =1=2G T _! g can be derived.<br />

With angles ; < < , can be used as the (orientation) state from which 0 =<br />

p 1 ; T is solved. In that case, a three parameter setting is used at the cost <strong>of</strong> possible<br />

singularities. The nonsingular envelope is, however, twice as large as in case <strong>of</strong> using euler<br />

angles.


40 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 2.4: A 2-d.o.f. serial <strong>system</strong> in a singular configuration<br />

It is possible to calculate euler parameters from any rotation matrix where j j< .<br />

0 = 1p<br />

(1 + tr(R))<br />

2<br />

(2.24)<br />

1 = (R(3� 2) ; R(2� 3))=(4 0) (2.25)<br />

2 = (R(1� 3) ; R(3� 1))=(4 0) (2.26)<br />

3 = (R(2� 1) ; R(1� 2))=(4 0) (2.27)<br />

where tr(R) = P i=3<br />

i=1 R(i� i) is the trace <strong>of</strong> the rotation matrix. Only in case <strong>of</strong> 0 = 0<br />

another solution strategy has to be applied. With the earlier given relationships, (2.14) and<br />

(2.16) between R and and vice versa, euler parameters can be calculated from euler angles<br />

and vice versa.<br />

2.2.5 Jacobian matrices<br />

Connected parts <strong>of</strong> a mechanical <strong>system</strong> will result in less freedom <strong>of</strong> <strong>motion</strong> reflected by<br />

constraint equations which can be explicit as in y = f(x) or implicit as in g(x� y) = 0.<br />

Such equations also constrain the time derivatives <strong>of</strong> these coordinates, i.e. the velocities<br />

and accelerations. In describing a mechanical <strong>system</strong>, e.g. stating its equations <strong>of</strong> <strong>motion</strong>,<br />

it is <strong>of</strong>ten convenient to state the <strong>motion</strong> <strong>of</strong> the <strong>system</strong> as a function <strong>of</strong> a limited number <strong>of</strong><br />

(generalised) variables (coordinates and/or velocities).<br />

If some variations or velocities can be described as a product <strong>of</strong> a (position-dependent)<br />

matrix and a vector <strong>of</strong> other variations this matrix will be called a (semi-)Jacobian matrix.<br />

The jacobian matrix between two sets <strong>of</strong> variables usually comes out naturally by a time<br />

differentiated version <strong>of</strong> an equation in which one <strong>of</strong> the sets is explicitly stated as in y(t) =


2.2 Kinematics 41<br />

f(x(t)),<br />

_y(t) = @f<br />

@x (x(t)) _x(t) =Jy�x(x(t)) _x(t) (2.28)<br />

In mechanical <strong>system</strong>s such mappings <strong>of</strong>ten result in an indirect way from differentiation<br />

<strong>of</strong> the rotation matrix and using ! as velocity term. The angular velocity, !, however, is no<br />

time derivative <strong>of</strong> any representation <strong>of</strong> attitude. E.g. given (2.8),<br />

_p g = I g R m ( ~ P m ) T<br />

_t g<br />

! m = Jp g �x _x (2.29)<br />

Also matrices like Jp g �x will be termed jacobian, although this is not precise i.e. not a result<br />

<strong>of</strong> (2.28), but just states one set <strong>of</strong> velocities as a function <strong>of</strong> another set.<br />

2.2.6 Parallel, serial and singular configurations<br />

In kinematics, jacobian matrices <strong>of</strong>ten play an important role in relating input/actuator and<br />

output/ end-effector coordinates. With parallel manipulators, it will be shown using some<br />

examples, that these relations differ from those constructed from a serial manipulator. In<br />

Fig. 2.4, a 2 degree-<strong>of</strong>-freedom serial joint robot is depicted. The end-effector coordinates<br />

x and y can in the serial configuration explicitly be stated as functions <strong>of</strong> actuated rotating<br />

joints angles q1 and q2 as<br />

And by differentiation<br />

x = l1 cos(q1) +l2 cos(q1 + q2) (2.30)<br />

y = l1 sin(q1) +l2 sin(q1 + q2) (2.31)<br />

_x<br />

_y = ;l1 sin(q1) ; l2 sin(q1 + q2) ;l2 sin(q1 + q2)<br />

l1 cos(q1) +l2 cos(q1 + q2) l2 cos(q1 + q2)<br />

_q1<br />

_q2<br />

= Jx�qq (2.32)<br />

At q2 =0 , the jacobian <strong>of</strong> this configuration becomes singular. At such configurations,<br />

physically, the actuators can only generate one d.o.f. <strong>motion</strong>/velocity <strong>of</strong> the end-effector<br />

locally. Further, it can be noted that, for each nonsingular configuration, two actuator positions<br />

generate the same end-effector state.<br />

The dual parallel configuration <strong>of</strong> a 2 d.o.f. robot is given in Fig. 2.5. In fully parallel<br />

<strong>system</strong>s, the sliding actuator lengths can explicitly be stated as functions <strong>of</strong> the end-effector<br />

coordinates. With xT =[xy] and li(x) =x ; bi<br />

q<br />

klik = lT i li<br />

(2.33)<br />

and by differentiation<br />

k _ lik = l T n�i _x (2.34)<br />

Thus J T l12�x = [ln�1 ln�2]. If the unit actuator direction vectors, ln�i are in parallel, the<br />

configuration becomes singular. In this case, however, the end-effector can still move but


42 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 2.5: A 2-d.o.f. parallel <strong>system</strong> in a singular configuration<br />

Fig. 2.6: A 3-d.o.f. <strong>system</strong> with serial and parallel singular configurations<br />

not be forced in 2 d.o.f. and from the speed <strong>of</strong> the actuators, the end-effector velocity can<br />

not be determined. In this <strong>system</strong>, in a nonsingular configuration, there are two end-effector<br />

states for each actuator position.<br />

In general, a rigid body mechanical <strong>system</strong>, will have combinations <strong>of</strong> serial and parallel<br />

connections as in Fig. 2.6. In this example, the actuator coordinates, q, are no explicit functions<br />

<strong>of</strong> the end-effector coordinates, x and v.v. Also both kind <strong>of</strong> singular configurations<br />

can usually occur. An important observation in the following sections will be the fact that<br />

the Stewart platform <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> is a purely parallel <strong>system</strong>.<br />

The parallel manipulator construction <strong>of</strong> the Stewart platform is first defined. To derive<br />

the equations <strong>of</strong> <strong>motion</strong>, the velocity and accelerations should be described w.r.t. a<br />

limited number <strong>of</strong> generalized variations. This defines the kinematics after which the semiequilibrium<br />

equations <strong>of</strong> the active and inertial forces can be stated.<br />

φ


2.2 Kinematics 43<br />

2.2.7 Stewart platform definitions and assumptions<br />

The Stewart platform (Fig. 2.7) consists <strong>of</strong> an end-effector body with mass, m c, and 3x3<br />

inertia matrix I m c w.r.t. the end-effector frame, M, connected to the centre <strong>of</strong> gravity (c.o.g.)<br />

<strong>of</strong> the body which has varying coordinates, c, in the inertial frame, G. The end-effector body<br />

or platform is connected by six parallel actuators at the moving upper gimbal points with<br />

fixed coordinates, ai, in the moving frame, M, tobi in the inertial frame, G. The length <strong>of</strong><br />

the six actuators can be varied. In describing a specific actuator the subscript i for the i th<br />

actuator will be left away.<br />

The platform position is defined as<br />

sx =<br />

c<br />

s<br />

(2.35)<br />

With the last three euler parameters, ,todefine the rotation matrix, T ( ), asR in (2.20),<br />

from the moving (platform) frame, M, to inertial (ground) frame, G and a scaling factor<br />

s = 2kakmax equal to the maximum distance <strong>of</strong> the c.o.g. to any <strong>of</strong> the upper gimbal<br />

points. The platform speed is not defined as the derivative <strong>of</strong> platform position but as<br />

_x =<br />

_c<br />

! � (2.36)<br />

where ! is the angular velocity <strong>of</strong> the platform. The platform (generalised) speed and<br />

position are related through the jacobian, J sx�x, using (2.22)<br />

sx _ =<br />

_c<br />

s_<br />

= I 0<br />

0 sG T s =2<br />

_c<br />

! = Jsx�x _x (2.37)<br />

An actuator (Fig. 2.8) will be modelled as 2 bodies, the rotating cylinder and the moving<br />

piston. The rotating body, with mass mb and a constant distance <strong>of</strong> rb <strong>of</strong> the c.o.g., bc, to<br />

the connection <strong>of</strong> a 2-d.o.f.-rotational gimbal joint to the inertial frame at b. The moving<br />

actuator body, i.e. the piston, with mass ma with a constant distance <strong>of</strong> ra <strong>of</strong> the c.o.g.,<br />

ac, is connected with a 3-d.o.f.-rotational gimbal joint to the platform at a. Witha1d.o.f.<br />

<strong>control</strong>led sliding joint between these two bodies the length <strong>of</strong> the actuator can be<br />

varied.<br />

It is assumed that the inertia <strong>of</strong> the actuator bodies can be neglected around the actuator<br />

axis and to be uniform perpendicular to this axis. i a is the inertia <strong>of</strong> the moving actuator<br />

body at ac and any axis perpendicular to the actuator. i b is the inertia <strong>of</strong> the rotating part<br />

<strong>of</strong> the actuator along any axis perpendicular to the actuator w.r.t. the connection point to the<br />

inertial frame (b) .<br />

With these assumptions also the case, <strong>of</strong>ten seen in practise, in which the moving part <strong>of</strong><br />

the actuator not only slides at the connection with the rotating part but also rotates around<br />

the l-axis , and has only a 2-d.o.f. rotation gimbal in connection to the platform, results in<br />

the same dynamics. Finally, the c.o.g.’s <strong>of</strong> the actuator parts are assumed to lie on the line<br />

connecting the upper and lower gimbal point.


44 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 2.7: A schematic view <strong>of</strong> the Stewart platform<br />

2.2.8 Stewart platform kinematics<br />

The kinematics <strong>of</strong> the Stewart platform will be described by first defining the transformation<br />

<strong>of</strong> the platform to actuator coordinates. Then by differentiation also velocity and acceleration<br />

<strong>of</strong> all relevant points can be calculated as a function <strong>of</strong> the platform <strong>motion</strong>, whose<br />

velocities will be taken as the generalized speeds.<br />

Almost all vectors can be conveniently described in the inertial frame. Apart from a m i<br />

whose time derivative in the moving frame is 0.<br />

As shown in Fig. 2.7, the vector, li, between the two attachment points <strong>of</strong> an actuator<br />

can be described by<br />

li =c + T a m i ; bi (2.38)<br />

Now the squared length <strong>of</strong> the actuator, j li j 2 = l T i li, and the unit vector in direction <strong>of</strong><br />

the actuator, ln�i = li<br />

jlij can be calculated from the platform variables c g and the orientation<br />

matrix T = g R m which will be the only rotation matrix used.<br />

The velocity <strong>of</strong> the length <strong>of</strong> the actuators can be calculated by projection <strong>of</strong> the velocity<br />

<strong>of</strong> the upper gimbal attachment point, v a, in the direction <strong>of</strong> the actuator, since<br />

d<br />

dt j li j= d<br />

dt<br />

q l T i li = lT i va<br />

j li j = lT n va: (2.39)


2.2 Kinematics 45<br />

Using (2.8), the velocity <strong>of</strong> the upper gimbal points is given by<br />

vai = _c +! T a m i � (2.40)<br />

where ! is the angular velocity <strong>of</strong> the moving frame in the inertial frame. By projection <strong>of</strong><br />

this velocity with (2.39) and filling in (2.40), the velocity <strong>of</strong> the actuator is described as a<br />

function <strong>of</strong> platform velocities and poses.<br />

_li = l T n�ivai = lT n�i _c + l T n�i (! T ami ) (2.41)<br />

With some reordering and written as matrix equation (ln�i, am i and vai stacked in Ln, Am and Va) for all the actuators the jacobian between the actuator and platform velocities is<br />

defined.<br />

_<br />

l = L T n _c +(TA m Ln) T ! = Jl�x _x = L T n Va (2.42)<br />

The semi-jacobian matrix, Jl�x, is one <strong>of</strong> the most important variables in the Stewart platform,<br />

relating the platform coordinates to be <strong>control</strong>led and used as basic model coordinates,<br />

and the actuator lengths, which can be measured. Further, in transposed form, J l�x relates<br />

the actuator input forces with the platform forces as will be shown in deriving the platform<br />

dynamics. Note that with defining _x T =[_c T ! T ], x does not exist.<br />

2.2.9 Interpretation and use <strong>of</strong> the jacobian matrix, Jl�x<br />

If the <strong>system</strong> has to be <strong>control</strong>led by the actuators, the jacobian specifies how the <strong>control</strong><br />

inputs, the actuator forces, influence the platform (accelerations), which are, especially in<br />

<strong>flight</strong> simulation applications, <strong>of</strong>ten the variables to be <strong>control</strong>led. Further, the measurable<br />

outputs are <strong>of</strong>ten only the actuator lengths. The derivatives (actuator speed) <strong>of</strong> these outputs<br />

are given by the product <strong>of</strong> the jacobian and the platform speed.<br />

There are two interpretations to the jacobian. In the force interpretation the rows <strong>of</strong> J l�x<br />

give the (generalized) forces in the platform coordinates given a unit force in an actuator. In<br />

the velocity interpretation the columns <strong>of</strong> Jl�x specify the velocity <strong>of</strong> the actuators required<br />

to have unit velocity <strong>of</strong> the platform.<br />

In model <strong>based</strong> <strong>control</strong>, the inverse information is <strong>of</strong> interest. The measured variations<br />

<strong>of</strong> the actuator have to be put in platform variations to calculate corrections in a model<br />

specified in platform coordinates. Each column <strong>of</strong> the inverse jacobian, J ;1<br />

l�x , specifies what<br />

velocity (angular velocity included) <strong>of</strong> the platform is necessary to have elongation <strong>of</strong> just<br />

one actuator while the others only rotate. Given measurable actuator velocities, the platform<br />

velocities can be calculated.<br />

The correction forces in a model <strong>based</strong> <strong>control</strong>ler are also most conveniently calculated<br />

as a function <strong>of</strong> platform coordinates. Each row <strong>of</strong> J ;1<br />

l�x specifies the forces necessary in the<br />

actuators to have unit force correction in platform coordinates.<br />

The inverse jacobian appears in feedback linearising structures (like computed torque,<br />

etc.), which will be dealt with in the forthcoming chapters.<br />

Another problem <strong>of</strong> a parallel manipulator, with only the link position measured, are<br />

the forward kinematics. It is not known how to analytically calculate the platform position<br />

(without decision making about set <strong>of</strong> possible solutions) from link measurements.


46 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

r b<br />

l<br />

bc<br />

ac<br />

b<br />

Lower Gimbal<br />

ra<br />

Upper Gimbal<br />

Fig. 2.8: Stewart platform actuator link construction<br />

The jacobian Jl�x(sx) is an important element in the jacobian Jsx�l(sx) from actuator<br />

variation to the platform position variation.<br />

Jsx�l(sx) =J ;1<br />

;1<br />

l�sx (sx) =Jsx�x(sx)J l�x<br />

(sx)� (2.43)<br />

where Jsx�x is given by (2.37).<br />

This jacobian provides a way to apply a Newton-Raphson iteration to calculate the solution<br />

provided one starts in a point sufficiently close to the solution and away from jacobian<br />

singularities.<br />

sxj+1 = sxj + J ;1<br />

l�sxj (lmeasured ; lj) (2.44)<br />

The condition number <strong>of</strong> Jl�x also provides a measure for the <strong>control</strong>lability <strong>of</strong> the platform<br />

from the actuators which becomes un<strong>control</strong>lable at singularities <strong>of</strong> this matrix.<br />

Further, most <strong>of</strong> the constraints <strong>of</strong> the platform are caused by the characteristics <strong>of</strong> the<br />

actuators like limited stroke, maximum speed and force. The jacobian plays an important<br />

role in translating these limitations into platform coordinates. E.g. given a maximum velocity,<br />

j v jmax, in extending an actuator either direction, the maximum velocity in the j’th<br />

platform coordinate is given by the 1-norm <strong>of</strong> the respective row <strong>of</strong> J ;1<br />

l�x times the maximum<br />

actuator velocity.<br />

_xi�max = kJ ;1<br />

l�x (i� )k1 j v jmax=<br />

6X<br />

j=1<br />

j J ;1<br />

l�x (i� j) jj v jmax (2.45)<br />

a


2.2 Kinematics 47<br />

So, the maximal platform velocity along a specific coordinate can be calculated by adding<br />

all absolute values <strong>of</strong> the specific row <strong>of</strong> the inverse jacobian and multiplying with the<br />

maximum actuator velocity. In this case, every other platform coordinate is left unspecified.<br />

In the more interesting case, with the other platform coordinates constraint to zero<br />

_xi�max =j v jmax =kJl�x(i� )k1 =j v jmax = max(j<br />

Jl�x(i� j) j)� (2.46)<br />

j<br />

e.g. the maximum pure surge speed can be calculated by dividing the maximum actuator<br />

velocity by the largest number in the first column <strong>of</strong> the Jacobian. For position and acceleration,<br />

these formulas can be used as an approximation.<br />

2.2.10 Velocity and acceleration <strong>of</strong> the actuator joints<br />

If the inertia <strong>of</strong> the actuator joints is important or gimbal forces have to be calculated, the<br />

kinematical formulas have to be expanded with those to derive the velocities and accelerations<br />

<strong>of</strong> these joints.<br />

The jacobian between the platform and the upper gimbal point velocity is defined by<br />

vai = I T ( ~ A m i )T T T _x = Jai�x _x (2.47)<br />

To determine the inertial forces <strong>of</strong> the actuators, the jacobians, from gimbal point to the<br />

c.o.g.’s <strong>of</strong> the actuators, are also important. The angular velocity <strong>of</strong> the actuator perpendicular<br />

to the actuator, !a,isdefined by<br />

va<br />

!a = ln<br />

(2.48)<br />

j l j<br />

Now the velocities <strong>of</strong> the c.o.g.’s <strong>of</strong> the actuator bodies v ac and vbc can be stated as a function<br />

<strong>of</strong> va:<br />

and<br />

vac =va +!a<br />

(;raln) =(I ; ra<br />

j l j Pln )va = Jac�ava� (2.49)<br />

vbc =!a rbln = rb<br />

j l j Plnva = Jbc�ava� (2.50)<br />

as becomes more clear by looking at Fig. 2.9. The acceleration <strong>of</strong> the actuators can be<br />

calculated by differentiating (2.42),<br />

l = Jl�xx + _<br />

Jl�x _x = L T n _ Va + _ L T n Va: (2.51)<br />

The derivative <strong>of</strong> the unit vectors, ln�i, <strong>of</strong> each actuator, i, can be calculated with:<br />

_<br />

ln = d l<br />

dt j l j =<br />

l<br />

_ j l j;ld dt j l j<br />

j l j2 = (I ; lnlT n )<br />

va =<br />

j l j<br />

1<br />

j l j Plnva<br />

(2.52)


48 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

rb ln<br />

vbc<br />

ω a<br />

bc<br />

ω a<br />

ac<br />

Lower Gimbal<br />

va<br />

ω a<br />

vac<br />

ra ln ra l n<br />

x<br />

ω a<br />

ln .<br />

ln<br />

l<br />

Upper Gimbal<br />

Fig. 2.9: Stewart platform actuator construction <strong>of</strong> link velocities. With the upper gimbal<br />

velocity, va, which can be derived from the platform velocity by (2.40), all other<br />

actuator velocities can be constructed. The actuator speed, _ l, is the projection <strong>of</strong><br />

va along the actuator ln (2.41). The figure is best read by considering v a in the<br />

plane <strong>of</strong> the picture. Then, only the actuator angular velocity, ! a, given by (2.48),<br />

is out <strong>of</strong> the plane. In fact, it is orthogonal to the plane.


2.2 Kinematics 49<br />

The acceleration <strong>of</strong> the actuator length consists <strong>of</strong> a term which is the projection <strong>of</strong> the<br />

acceleration <strong>of</strong> the upper gimbal in the direction <strong>of</strong> the actuator and a positive quadratic<br />

term, which is the centripetal acceleration <strong>of</strong> the actuator.<br />

l = l T n _va +v T 1<br />

a ( Pln )va<br />

(2.53)<br />

j l j<br />

So the acceleration <strong>of</strong> the actuator length is always positive if the platform is moving with<br />

constant translational speed in any direction and constant orientation since _va =0in that<br />

case. The acceleration <strong>of</strong> the upper gimbal can be derived directly with (2.9) and (2.40)<br />

_vai = c + _! ai +! (! ai) =Jai�xx; j! j 2 P!ai� (2.54)<br />

where the projection matrix, P! =(I +!n! T n ), projects the upper gimbal vector, ai, on the<br />

plane orthogonal to the normalised angular velocity direction <strong>of</strong> the platform, ! n.<br />

The acceleration <strong>of</strong> the c.o.g. <strong>of</strong> the moving actuator part also generates inertial forces<br />

and can be written as a function <strong>of</strong> platform <strong>motion</strong>.<br />

_vac = d<br />

dt vac = d<br />

dt (Jac�ava) =Jac�a _va + _ Jac�ava<br />

So this jacobian needs to be differentiated. With (2.49) and (2.52),<br />

Now with j Plnva j 2 =v T a Plnva,<br />

Jac�a<br />

_ = d ra<br />

(I ;<br />

dt j l j<br />

=<br />

ra<br />

j l j2 vT ra<br />

a lnPln +<br />

j l j2 (Plnval T n + lnv T a Pln )<br />

(2.55)<br />

Pln ) (2.56)<br />

Jac�ava<br />

_ = ra<br />

j l j2 (j Plnva j 2 ln +2(v T a ln)Plnva)� (2.57)<br />

which shows a quadratic centripetal term in actuator direction and a coriolis term orthogonal<br />

to the actuator. Clearly, the piston acceleration, _vac, only depends on the velocity, va and<br />

acceleration, _va, <strong>of</strong> the upper gimbal, given the positional coordinates.<br />

The <strong>motion</strong> <strong>of</strong> the actuators has now been explicitly stated as a function <strong>of</strong> the platform<br />

coordinates and its derivatives. The kinematic relations thus provide means to state the<br />

<strong>motion</strong> <strong>of</strong> the <strong>system</strong> as a function <strong>of</strong> a limited number <strong>of</strong> variables. Together with the<br />

dynamics i.e. semi-equilibria <strong>of</strong> active and inertial forces stated in Section 2.4, the equations<br />

<strong>of</strong> <strong>motion</strong> result.<br />

2.2.11 The Simona <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> kinematics<br />

In this thesis, any <strong>system</strong> connected to the environment through six parallel joints with five<br />

(passive) rotational and one (active) translational d.o.f. is considered a Stewart platform.<br />

The structure <strong>of</strong> the Simona <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> design is much more specific. It<br />

is schematically depicted from above in Fig. 2.10 and the measures are given in Table 2.1.


50 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 2.10: The Simona <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> kinematics represented schematically<br />

in a top view in an ’all the actuators down’ position<br />

Both the upper as the lower gimbal points virtually form a platform, which looks like two<br />

blunt equilateral triangles rotated 180 degrees with respect to each other.<br />

All the upper gimbal points as well as the lower lie on a circle. The circles lie in parallel<br />

planes if the actuator lengths are equal. All the minimal and maximal actuator lengths are<br />

equal. The actuators and gimbal points, which are numbered 1 to 6, are pairwise distributed<br />

clockwise over the circle. (6,1), (2,3) and (4,5) for the moving upper and (1,2), (3,4) and<br />

(5,6) for the inertial lower platform. The distance between the two elements <strong>of</strong> the pairs are<br />

also specified in the table. The upper pair (6,1) is connected on the positive x-side <strong>of</strong> the<br />

<strong>simulator</strong>. The lower pair (3,4) at the negative x-side <strong>of</strong> the ground.<br />

In most <strong>of</strong> the derivations, the <strong>motion</strong> <strong>system</strong> kinematics is assumed to belong to the<br />

most general class <strong>of</strong> Stewart platforms. This leaves room to adjust easily to a structure<br />

which differs from its design.<br />

This concludes the kinematical modelling, which is used to derive the equations <strong>of</strong> <strong>motion</strong><br />

in modelling the mechanical <strong>system</strong> in Section 2.4. The nest section finishes the treatment<br />

<strong>of</strong> the kinematics with two issues, which will be specifically important for the model<br />

<strong>based</strong> <strong>control</strong> structure to be used in Chapter 5.<br />

2.3 Analysis <strong>of</strong> the Stewart platform kinematics<br />

A vast majority <strong>of</strong> the literature on parallel robotic <strong>system</strong>s is devoted to their kinematics.<br />

This is an important subject because in the design <strong>of</strong> the kinematics <strong>of</strong> a robotic <strong>system</strong>,


2.3 Analysis <strong>of</strong> the Stewart platform kinematics 51<br />

Table 2.1: The Simona <strong>system</strong> parameters<br />

variable value description<br />

ra 1:60 m upper gimbal radius<br />

rb 1:65 m lower gimbal radius<br />

du 0:20 m upper gimbal spacing<br />

dl 0:60 m lower gimbal spacing<br />

lmin 2:08 m minimal actuator length<br />

qmax 1:25 m actuator stroke<br />

qop 1:15 m operational actuator stroke<br />

_lmax 1 m=s maximum actuator speed<br />

its manipulability, its workspace, etc. is to be fixed. This problem in combination with the<br />

application to <strong>flight</strong> simulation has been studied by Advani [7].<br />

In this section, two new solutions towards kinematical problems will be presented,<br />

which have a direct relation with the ability to perform real time model <strong>based</strong> feedback<br />

safely on a specific kinematic design <strong>of</strong> a Stewart platform <strong>motion</strong> <strong>system</strong>.<br />

- Given a specific design <strong>of</strong> the manipulator, one has to ensure the <strong>system</strong> can not be driven<br />

into a singular configuration.<br />

- Further, model coordinates have to be calculated from the measurements taken in real<br />

time (e.g. a finite number <strong>of</strong> calculation steps) without the chance to have divergence<br />

between actual and calculated coordinates.<br />

In parallel robots such as most <strong>flight</strong> simulation <strong>motion</strong> platforms, the position <strong>of</strong> the<br />

<strong>system</strong> is usually indirectly measured by the length <strong>of</strong> the actuators. The forward kinematical<br />

problem <strong>of</strong> calculating the platform coordinates given the actuator lengths <strong>of</strong> the Stewart<br />

platform is seen to be solved in roughly two ways in literature.<br />

Using analytic techniques, the problem can be transformed to a set <strong>of</strong> combined polynomial<br />

equations whose roots have to be found to solve the forward kinematics e.g. [57].<br />

Although these equations can provide insight into the structure <strong>of</strong> the problem, closed form<br />

solutions are only seen to be presented for special classes <strong>of</strong> platforms e.g. Merlet [97],<br />

which still is an active area <strong>of</strong> research [68]. Solving the roots <strong>of</strong> the equations, still leaves<br />

the problem <strong>of</strong> choosing the actual pose. The complexity <strong>of</strong> this problem is reflected by the<br />

fact that it has been shown that the general version <strong>of</strong> the Stewart platform can have up to<br />

40 real solutions [34].<br />

Secondly, for long, the forward kinematics has been tackled numerically by performing<br />

the Newton-Raphson (NR) iteration scheme [35] already given in (2.44).<br />

sxk+1 = sxk + J ;1<br />

l�sx (sxk)(lmeasured ; lk) (2.58)


52 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

As the actuator lengths, l, are explicit functions <strong>of</strong> the platform coordinates, sx, the<br />

jacobian, Jl�sx, is a function <strong>of</strong> platform coordinates.<br />

Jl�sx(x)(i� j) = @li(sx)<br />

@sxj<br />

(2.59)<br />

This latter method is preferred here as it is less involved to be implemented in a realtime<br />

model <strong>based</strong> <strong>control</strong>ler where inversion <strong>of</strong> the jacobian is already part <strong>of</strong> the <strong>control</strong><br />

structure. Convergence or convergence to the actual physical pose in this scheme is not<br />

guaranteed in general. As the forward kinematics <strong>of</strong> a Stewart platform have more than one<br />

solution [57], and since singular points <strong>of</strong> the jacobian for unconstrained actuator lengths<br />

exist [87], the iteration scheme does not converge globally to the right solution. This problem<br />

has only recently been considered in literature [21], [70].<br />

Chetelat [21] considers the problem <strong>of</strong> convergence in a general way by considering reconstruction<br />

<strong>of</strong> coordinates <strong>of</strong> an implicit (matrix) function from which the image is known<br />

e.g. a Stewart platform pose from length measurements. He points out that if one considers<br />

functions, <strong>of</strong> possibly interconnected vector equations, <strong>of</strong> polynomial degree


2.3 Analysis <strong>of</strong> the Stewart platform kinematics 53<br />

and minimum gain <strong>of</strong> J and again the Lipschitz condition. These can be calculated for<br />

Stewart platforms. Exclusion <strong>of</strong> singular points <strong>of</strong> J is necessary to calculate a radius <strong>of</strong> the<br />

neighbourhood in which convergence is guaranteed.<br />

From this radius, the maximum gain <strong>of</strong> J and the maximum speed <strong>of</strong> the actuator, a<br />

sufficient update frequency <strong>of</strong> the iteration can be calculated above which (quadratic) convergence<br />

is guaranteed.<br />

The parameters <strong>of</strong> the new Simona research <strong>simulator</strong> are used as an example which<br />

shows that reasonable results can be obtained although the conditions derived are rather<br />

conservative.<br />

2.3.1 Convergence NR-iteration<br />

A weak version <strong>of</strong> the Newton-Kantovorich theorem [109] given by Stoer [141] will be<br />

used. From this theorem convergence <strong>of</strong> the NR-iteration can be inferred. It is stated as<br />

follows.<br />

Theorem 2.1 Given: a set D IR n , a convex set Do with exterior Do D and a function<br />

f : D ! IR n which is continuous on D and differentiable with derivative Df(x) on Do.<br />

If positive constants r, , , and h can be found for x o 2 Do such that Sr(xo) =<br />

fx jkx ; xok 0<br />

B) limk!1 xk = exists, 2 Sr(xo), and f( )=0<br />

C)<br />

8 k 0� kxk ; k<br />

h 2k;1<br />

1 ; h 2k<br />

With 0


54 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

The pro<strong>of</strong> <strong>of</strong> this theorem is given in [141].<br />

Roughly speaking, this theorem states that a solution can be found in the NR-iteration<br />

(B) if the differential Df (x) does not vary too much (a)), is far enough from singularities<br />

(b)) and eventually does not jump too close to the boundary <strong>of</strong> the defined neighbourhood,<br />

at the first iteration (c)).<br />

It also states that the iteration will not go out <strong>of</strong> a specified neighbourhood (A) and<br />

converges at a certain speed (C). Conditions for a NR-iteration towards the Stewart platform<br />

coordinates will be derived.<br />

Chetelat [21] uses the fact that if one ensures that the function, f (x), is not <strong>of</strong> polynomial<br />

degree higher than two, then the jacobian is Lipschitz and in this case the Lipschitz constant<br />

can be calculated using the infinity norm.<br />

2.3.2 NR-convergence Stewart platform<br />

Having stated the kinematical structure <strong>of</strong> the Stewart platform and given Theorem 2.1 on<br />

convergence <strong>of</strong> a NR-iteration, it is now possible to investigate under what conditions the<br />

specific NR-iteration <strong>of</strong> (2.58) will converge to the physical platform pose.<br />

First an appropriate definition <strong>of</strong> the coordinates has to be given. The defined jacobian,<br />

Jl�x, is the description <strong>of</strong> platform translational and angular speed/variation to actuator<br />

length variations. To go from orientational parameter variations to angular speed depends<br />

on the parametrization used.<br />

In this case the last three euler parameters, are used to parametrize orientation, sx T =<br />

[cT s T ] T , where s is a scaling factor which can be used to get less conservative results<br />

in specifying a variation <strong>of</strong> x since c and = sin( 1<br />

2 )n have different dimension (m and<br />

rad). The scaling factor s =2kakmax will shown to be appropriate in the sequel.<br />

It will be assumed that after each iteration will be reset to zero. In that case the<br />

relation between the angular velocity and the euler parameters is very simple, ! j =0= 2I _.<br />

Off course the rotation matrix in the iteration is now the multiplication <strong>of</strong> all the rotation<br />

matrices calculated, Tk+1 = T ( k+1)Tk. The continuous function, f, in Theorem 2.1, from<br />

which the platform coordinates have to be found can now be given by k l i k;klik = fi(sx)<br />

where kl i k is the measured length <strong>of</strong> the i th actuator (fixed value for each iteration) and klik<br />

is the length <strong>of</strong> the i th actuator given sx (li given by (2.38)). Since the measured length is<br />

fixed, the derivative function with scaled euler parameters, is only slightly different from<br />

the jacobian Jl�x given earlier in (2.42).<br />

Df (sx) T = Jl�sx(sx) T =<br />

Ln<br />

2(TA m Ln)=s<br />

(2.62)<br />

Since it is a function <strong>of</strong> unit direction lengths, it is only defined for k lk 6=0.<br />

To derive the conditions (a,b,c)-st(ewart)pl(atform) for convergence <strong>of</strong> the NR-iteration<br />

defined in Theorem 2.1 for the Stewart platform, the constants ( � � ) can be specified<br />

using the kinematics.<br />

a-stpl) In this case the 2-norm <strong>of</strong> the matrix is taken which is equal to the largest singular<br />

value, and in the formulas d means an finitely small difference.


2.3 Analysis <strong>of</strong> the Stewart platform kinematics 55<br />

kJl�sx(sx1) ; Jl�sx(sx2)k < stplksx1 ; sx2k (2.63)<br />

kJl�sx(sx1) ; Jl�sx(sx2)k =<br />

dLn<br />

2d(TA m Ln)=s<br />

= (dJ) (2.64)<br />

The Frobenius (semi)-norm is an upper bound on this norm and is advantageous in<br />

case <strong>of</strong> the Stewart platform since specific bounds can be derived on the matrix elements<br />

as will be shown later on.<br />

kdJkF =<br />

vu<br />

u<br />

t 6X<br />

i=1<br />

(dJ) kdJkF (2.65)<br />

(kdln�ik 2 + k2d(T ai ln�i)=sk 2 (2.66)<br />

To derive a constant, stpl, the separate elements <strong>of</strong> the last equation will be described<br />

as a function <strong>of</strong> sx. This will be done in the Section 2.3.3.<br />

b-stpl) To state the second condition from which a constant number stpl has to be calculated<br />

also the 2-norm is used which can be upper bounded by one over the minimal<br />

singular value, min, <strong>of</strong> the jacobian at some pose, xo. By using the maximal<br />

variation <strong>of</strong> the jacobian which has been calculated for the previous condition, the<br />

constant, , becomes an upper bound for the maximum gain <strong>of</strong> the inverse jacobian<br />

over a volume <strong>of</strong> poses, sx.<br />

kJ ;1<br />

l�sx (sx)k =( min(Jl�sx(sx))) ;1<br />

(2.67)<br />

(max( min(Jl�sx(sxo)) ;kdJkF � 0)) ;1 = stpl (2.68)<br />

c-stpl) To calculate stpl, also use can be made <strong>of</strong> the singular values. Since any point<br />

in the work space is a possible initial start <strong>of</strong> the iteration, the maximum condition<br />

number <strong>of</strong> the scaled jacobian over the work space can be taken as the constant, stpl.<br />

kJ ;1<br />

l�sx f (sx)k = min (Jl�sx)kdsxk = stpl (2.69)<br />

In the next part it will be shown that indeed the jacobian <strong>of</strong> a Stewart platform is Lipschitz,<br />

as a-stpl) requires, as long as the actuators <strong>of</strong> the platform have minimal stroke strictly<br />

larger than 0.


56 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

2.3.3 Lipschitz condition on jacobian<br />

By constructively analyzing the kinematics <strong>of</strong> the Stewart platform the following theorem<br />

can be derived.<br />

Theorem 2.2 The Stewart platform jacobian, Jl�sx, is lipschitz i.e. a can be found such<br />

that<br />

kJl�sx(sx1) ; Jl�sx(sx2)k ksx1 ; sx2k 8 sx1� sx2 2 Do (2.70)<br />

if kli(sx)k > >0 8 sx 2 D� i 2f1�::: �6g: (2.71)<br />

Note that the requirement <strong>of</strong> an actuator length larger than zero is an implicit constraint<br />

on the set <strong>of</strong> platform poses.<br />

To prove Theorem 2.2, it is observed that jacobian, J l�sx, in (2.62) consists <strong>of</strong> two<br />

elements: the unit actuator direction, ln, and the vector product, (ln T a), for each actuator.<br />

By bounding the difference between the unit actuator directions and vector products <strong>of</strong> two<br />

jacobians with a function linear in kdsxk, Theorem 2.2 can be proven and a can actually<br />

be constructed.<br />

In two steps dln will be bounded first.<br />

1. Difference <strong>of</strong> an upper gimbal connection coordinate, dx a, will be bounded given the<br />

finite difference between two platform coordinates dsx.<br />

2. Then the difference between two actuator unit vectors, d ln, givendxa, will be considered.<br />

See Fig. 2.11 in which the <strong>motion</strong> <strong>of</strong> actuator as a function <strong>of</strong> the moving<br />

gimbal point xa is schematically depicted.<br />

With<br />

the following bound directly follows<br />

xa�i =c + T a m i<br />

(2.72)<br />

kdxa�ik kdck +2kakkd k p 2kdsxk (2.73)<br />

as the <strong>motion</strong> <strong>of</strong> a point due to rotation can be bounded easily with the reseted euler parameter<br />

description as<br />

kd(T a)k =2kakkd k (2.74)<br />

As from Fig. 2.3, it can be observed that the absolute change <strong>of</strong> a vector due to rotation has<br />

length 2 j sin(1=2 ) j=j 2 j.<br />

The second bound takes into account that as the upper gimbal moves within a ball<br />

(Fig. 2.11), the maximum change <strong>of</strong> the actuator direction is achieved if the new l, just


2.3 Analysis <strong>of</strong> the Stewart platform kinematics 57<br />

l n<br />

l<br />

Φ<br />

dl n<br />

dx a<br />

Fig. 2.11: Change <strong>of</strong> the unit vector in the direction <strong>of</strong> the actuator, ln, as a function <strong>of</strong><br />

change <strong>of</strong> the upper gimbal vector, dx a.<br />

touches the ball. In that case (ln + dln) ? dxa. With some geometry, sin( )=kdxak=klk<br />

and<br />

cos( )=<br />

s<br />

b<br />

1 ; kdxak<br />

klk<br />

the following monotonous upper bounding function can be derived for d ln.<br />

r<br />

1 ; cos( ) p kdxak<br />

kdlnkmax =2<br />

2<br />

2<br />

klk<br />

2<br />

a<br />

(2.75)<br />

(2.76)<br />

Taking into account (2.73) gives kdlnk 2kdsxk=klk. So, the actuator (minimum) length<br />

directly influences the bound, which can be put on kd lnk. If the length <strong>of</strong> the actuator is not<br />

strictly larger than zero, one can show that arbitrary small variations <strong>of</strong> the pose can result<br />

in nondecreasing variations <strong>of</strong> dln i.e. not satisfying the Lipschitz condition.<br />

Now, bounding the vector product is also possible. In general ka bk kakkbk and<br />

(a +c) (b + d) = (a b) +(a d) +(c b) +(c d). Further, rotation does not<br />

change the 2-norm, kT ak = kak. Change <strong>of</strong> the moving gimbal due to rotation is bounded<br />

by kd(T a)k =2kakkd k = kds k. Now,<br />

kd(ln T a)=kakmaxk<br />

kdln T ank + kln d(T an)k + kdln d(T an)k


58 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

p 2<br />

klk kdxak + kds k +<br />

2kds kkdsxk<br />

: (2.77)<br />

klk<br />

So in this way every element <strong>of</strong> the jacobian matrix is explicitly bounded by the variation<br />

<strong>of</strong> the platform pose and the (explicitly platform pose dependent) actuator length. If a<br />

minimal actuator length is given, a specific stpl can be calculated.<br />

Adding the bounds, for each <strong>of</strong> the actuators in the F-norm <strong>of</strong> the jacobian, results in an<br />

explicit number for stpl. Assuming small platform pose variations i.e. second order effects<br />

in the last equation are relatively small e.g. kdsxk


2.3 Analysis <strong>of</strong> the Stewart platform kinematics 59<br />

Algorithm<br />

- Choose an expected minimal singular value <strong>of</strong> the jacobians, over the grid points.<br />

- Take a grid such that the boxes with radius rs�min around the grid points fill the whole<br />

work space.<br />

- Calculate the minimal singular value <strong>of</strong> the jacobian at every grid point.<br />

- If the minimal singular value is larger than the expected value, there are no singular points<br />

in the work space, the algorithm finishes. If not, start another iteration choosing a<br />

smaller expected minimal singular value.<br />

Lemma 2.3 If the work space <strong>of</strong> a robotic manipulator having a bounded Lipschitz constant,<br />

does not have any singular point this non singularity will be detected by Algorithm 7.1<br />

in a finite number <strong>of</strong> iterations.<br />

Of course the boundary <strong>of</strong> the work space (in six dimensions!) should be known which is a<br />

stand alone problem (treated by Haug et al. [86]). To calculate an upper bound for the gain<br />

<strong>of</strong> the inverse jacobian ( stpl) afiner grid should be taken. (This will increase calculation<br />

time tremendously, e.g. n 6 grid points extra.)<br />

2.3.5 Sufficient update frequency<br />

The bounds derived are rather conservative in most cases. However, by calculating a convergence<br />

radius for the NR-iteration in the practical example <strong>of</strong> the new Simona research<br />

<strong>simulator</strong> shows that it is possible to guarantee that this iteration can be used if a reasonable<br />

frequency is used to update the platform pose. Further, the singular points can really be<br />

excluded from the work space in this case.<br />

To satisfy NR-iteration convergence the following constants were obtained using the<br />

kinematics <strong>of</strong> the Simona <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> given in Table 2.1.<br />

a-simona Using (2.78), assuming a minimal length <strong>of</strong> the actuator <strong>of</strong> l min =2:08m, results<br />

in simona =5:2.<br />

b-simona With gridding a minimal radius <strong>of</strong> rs�min = :09 is found. (About 200000 grid<br />

points had to be calculated to exclude singularity from the work space). The smallest<br />

singular value found is 0:75 and with finer gridding a simona =2can be guaranteed.<br />

c-simona The first value <strong>of</strong> f(sxo) in any point can not be larger than f(sxo) (J)kdsxk<br />

Together with the smallest gain this gives simona = kdsxk 3:5kdsxk<br />

Now with a bound kdsxk 0:02, hsimona =( )=2 =(3:5 0:02 2 5:2)=2 =0:23 and<br />

rsimona = =(1 ; h) =(3:5 0:02)=(1 ; 0:23) = 0:11. Within the every ball with radius<br />

rsimona, [a-simona] and [b-simona] should be guaranteed which is the case in the work<br />

space given the operational stroke (not including actuator cushioning part). To guarantee<br />

convergence in the whole work space also non-singularity, etc. should be guaranteed further


60 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

outside the work space which needs lots <strong>of</strong> calculation (with an extra stroke <strong>of</strong> 0.15m,<br />

singularities can be obtained so rs�min becomes very small).<br />

With a bound on the maximal speed <strong>of</strong> the actuator it is possible to calculate a minimal<br />

update frequency which guarantees kdsxk < 0:02. Givenk _ lk < 1m=s, now<br />

simona t _ lmax < kdsxk (2.80)<br />

and t 100 Hz convergence <strong>of</strong> the<br />

NR-iteration is attained.<br />

Lemma 2.4 The Stewart platform with the Simona <strong>motion</strong> <strong>system</strong> parameters has no singular<br />

points in the work space and the NR-iteration with this platform will converge to the<br />

right platform pose if the update frequency is larger than 100 Hz.<br />

Summarising, to satisfy a general convergence theorem on Newton-Raphson iteration,<br />

one <strong>of</strong> the requirements is the Lipschitz condition on the derivative function, which was<br />

shown to be satisfied for Stewart platforms. Variations <strong>of</strong> the jacobian can be bounded by<br />

the variations <strong>of</strong> the platform coordinates.<br />

Next to this requirement, the jacobian should not be singular in any point <strong>of</strong> the work<br />

space. With the Lipschitz condition on the jacobian and gridding, volumes can be excluded<br />

from singularities.<br />

Although only sufficient (and thereby conservative) conditions could be derived for convergence,<br />

it was shown to be possible to obtain an important convergence result in the practical<br />

example <strong>of</strong> Simona <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>. It will be one <strong>of</strong> the requirements<br />

for safe application <strong>of</strong> model <strong>based</strong> feedback.<br />

2.4 <strong>Model</strong>ling the mechanical part <strong>of</strong> the <strong>system</strong> dynamics<br />

After the thorough analysis <strong>of</strong> the kinematics <strong>of</strong> the parallel driven Stewart platforms in the<br />

previous section, this section will focus on the dynamics <strong>of</strong> such <strong>system</strong>s. The aim <strong>of</strong> this<br />

section is to show how to derive a limited number <strong>of</strong> differential, and possibly also some<br />

algebraic, equations, which describe the <strong>motion</strong> <strong>of</strong> mechanical <strong>system</strong>s such as the <strong>motion</strong><br />

<strong>system</strong> <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> as a function <strong>of</strong> the forces acting upon this <strong>system</strong>. It will<br />

be assumed that these <strong>system</strong>s consist <strong>of</strong> rigid bodies. In the extensions these bodies are<br />

possibly interconnected by springs and dampers.<br />

2.4.1 General theory in modelling the dynamics <strong>of</strong> mechanical <strong>system</strong>s<br />

Most theory described in this section can be found in many textbooks on mechanics and<br />

robotics, rigid multi body <strong>system</strong>s like [64, 81, 102, 106, 120, 129, 134]. Based on this<br />

general theory, a rigid body model <strong>of</strong> the Stewart platform, possibly with actuator joint<br />

inertia or linear ground flexibility, will be derived.<br />

All mechanics treated here are <strong>based</strong> on the assumption <strong>of</strong> a semi-equilibrium given by<br />

Newton’s second law f ; mp =0in an inertial frame for any mass particle. To describe the


2.4 <strong>Model</strong>ling the mechanical part <strong>of</strong> the <strong>system</strong> dynamics 61<br />

acceleration, p, <strong>of</strong> all parts in a <strong>system</strong> as a function <strong>of</strong> a limited number <strong>of</strong> variables some<br />

kinematics has been introduced.<br />

Having defined the kinematics <strong>of</strong> a mechanical <strong>system</strong>, its dynamics can be specified. In<br />

the equations <strong>of</strong> <strong>motion</strong> the semi-equilibria are described in a compact form as a function <strong>of</strong><br />

generalized velocities or coordinates. The integrals over the mass-particles <strong>of</strong> a body result<br />

in inertia matrices and the active forces are projected along the velocities by a virtual work<br />

argument.<br />

The equations <strong>of</strong> <strong>motion</strong> can be stated in several ways. First the Lagrange equations will<br />

be given. Then it will be shown that from these equations the more simple Newton-euler<br />

equations follow if one rigid body is considered. Finally the method <strong>based</strong> on Kane [64]<br />

will be introduced. This method in which all forces, active and inertial, are projected along<br />

a limited set <strong>of</strong> generalized velocities, will be central in the derivation <strong>of</strong> the <strong>motion</strong> <strong>system</strong><br />

mechanical model.<br />

Considering generalized velocities or coordinates ( _z� z), with the method <strong>of</strong> Lagrange<br />

the difference between the kinetic energy, K(z� _z), and potential energy, P(z), <strong>of</strong> a <strong>system</strong><br />

is called the Lagrangian, L(z� _z): L = K ; P. For a <strong>system</strong>, which can be described by a<br />

minimal number <strong>of</strong> generalized coordinates z, the Lagrange equations are given by:<br />

d<br />

dt<br />

@L<br />

@ _z<br />

; @L<br />

@z<br />

= (2.81)<br />

The driving moments/forces, , are all the working forces (non inertial or conservative) projected<br />

along the variations <strong>of</strong> the generalized coordinates. These are called the generalized<br />

forces.<br />

The kinetic energy can in general be described by<br />

K = 1<br />

2 _z T M (z) _z� (2.82)<br />

where the mass matrix, M (z), is a symmetric positive semi definite matrix.<br />

It can be appropriate to apply a coordinate transformation by introducing the generalized<br />

momenta (assuming a positive definite and thus invertible mass matrix)<br />

m = @L<br />

@ _z = M (z) _z: (2.83)<br />

The implicit second order differential Lagrange equations (2.81) can now be transformed<br />

to a (nonlinear) state space description, i.e. the Hamiltonian equations <strong>of</strong> <strong>motion</strong> following<br />

Van der Schaft [147]. A key role is played by the hamiltonian or total energy function, H,<br />

consisting <strong>of</strong> the addition <strong>of</strong> kinetic and potential energy,<br />

H(m�z) = 1<br />

2 mT M ;1 (z)m + P(z) =m T _z ; L(z� _z): (2.84)<br />

Taking the partial derivative <strong>of</strong> (2.84) to m for the first set <strong>of</strong> state equations and filling in<br />

(2.84) into (2.81) eliminating L for the second, results in<br />

_z = @H<br />

(m� z)<br />

@m<br />

(2.85)<br />

_m = ; @H<br />

(m� z) +<br />

@z<br />

: (2.86)


62 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

In this non energy dissipating <strong>system</strong> the increase in energy can be shown to be equal to the<br />

amount <strong>of</strong> work delivered,<br />

d<br />

dt H = _z T<br />

(2.87)<br />

By observing that the work W, done by forces , does not change if a change <strong>of</strong> coordinates<br />

z, is applied, it is easy to show that projecting the forces along other coordinates is equal to<br />

multiplying by the transpose jacobian.<br />

W = T z1 = T Jz1�z2 z2 =(J T z1�z2 )T z2 (2.88)<br />

Considering such collocated pairs <strong>of</strong> forces and velocities as inputs ( = J T (z)u) and outputs<br />

y = J(z) _z, a hamiltonian <strong>system</strong> results. The Stewart platform rigid body mechanics<br />

will have this structure. The hamiltonian <strong>system</strong> structure does not require that the number<br />

<strong>of</strong> input/output pairs and the number <strong>of</strong> generalized momenta/velocity pairs is equal. This<br />

allows to take into account parasitic resonant effects without losing the structural properties<br />

<strong>of</strong> a hamiltonian <strong>system</strong>.<br />

Further, manipulation <strong>of</strong> the second order differential equation (2.81) can lead to<br />

M (z)z +<br />

Compactly written as<br />

d<br />

dt (M (z) _z ; _z T<br />

@<br />

@zi<br />

M (z) _z + @<br />

P(z) = : (2.89)<br />

@z<br />

M (z)z + C(z� _z) _z + G(z) = : (2.90)<br />

with a mass matrix, M, a non linear coriolis/centripetal matrix, C, and a gravity vector, G.<br />

If a rigid body (with mass, m, and inertia, Ic) is considered at its centre <strong>of</strong> gravity c, the<br />

Newton-euler equations result. The mass matrix is block diagonal in this case.<br />

Kbody = 1<br />

2<br />

_c !<br />

mI<br />

0<br />

0<br />

Ic<br />

_c<br />

!<br />

(2.91)<br />

From the two blocks, two independent equations follow from (2.81). The impulse law:<br />

f = d d<br />

(m (c)) = mIc (2.92)<br />

dt dt<br />

And the impulse moment law with the generalized forces f g c (moments in this case) is<br />

derived similarly from (2.81) taking the time derivative <strong>of</strong> the Lagrangians partial derivative<br />

to the second set <strong>of</strong> (angular) velocities. The Lagrangian, in this case, only consists <strong>of</strong> the<br />

kinetic energy, which does not depend on the position.<br />

f g c<br />

= d<br />

dt (g R m I m c ! m )= g R m I m c _! m + g R m ~ m I m c ! m<br />

= I g c _! g + ~ g I g c ! g = I g c _! g +! g<br />

I g c ! g<br />

(2.93)<br />

In this way, the Newton-euler equations are easily stated for each rigid body in a <strong>system</strong>.<br />

However, extra equations with unknown internal forces result, in using this method to state


2.4 <strong>Model</strong>ling the mechanical part <strong>of</strong> the <strong>system</strong> dynamics 63<br />

the equations <strong>of</strong> <strong>motion</strong> for a multi body <strong>system</strong>. Applying Lagrange, results in taking<br />

partial derivatives <strong>of</strong> complex energy functions if the whole <strong>system</strong> is considered. These are<br />

disadvantages which can be circumvented by using Kane’s method.<br />

With Kane also generalized variations or velocities have to be specified. The general<br />

formula<br />

f + f = 0 (2.94)<br />

states the (semi-)equilibrium <strong>of</strong> the active forces, f, and inertial forces, f , projected along<br />

the directions <strong>of</strong> the generalized velocities. Generalized velocities do not have to be derivatives<br />

<strong>of</strong> some positional coordinate e.g. in principle the angular velocity can be used as<br />

such.<br />

To calculate the over-all inertial forces, as in the Newton-euler approach, the specific<br />

inertial forces generated in the frame <strong>of</strong> each body can be stated. As in the Lagrangian<br />

approach, a minimal number <strong>of</strong> equations results by writing the <strong>motion</strong> <strong>of</strong> the bodies as a<br />

function <strong>of</strong> the generalized velocities and projecting each specific force from its local coordinates<br />

to the generalized ones. Also the active forces can first be stated in an appropriate<br />

frame after which projection follows. The projection in general consists <strong>of</strong> a change <strong>of</strong> coordinates<br />

i.e. a multiplication with a jacobian. This procedure can also be automated [65].<br />

Especially in fast modelling with the aim <strong>of</strong> simulation, this can be useful.<br />

With this method it is possible to start with a strongly simplified <strong>system</strong> by calculating<br />

part <strong>of</strong> the (inertial) forces and separately adding other forces if a more accurate model<br />

has to be taken into account. Many <strong>of</strong> the specific merits <strong>of</strong> the method will become clear<br />

considering the modelling <strong>of</strong> the mechanical part <strong>of</strong> the Stewart platform including actuator<br />

inertia.<br />

The amount <strong>of</strong> generalized coordinates can exceed the number <strong>of</strong> d.o.f. <strong>of</strong> a <strong>system</strong>. In<br />

that case, next to the differential equations given by the semi-equilibria <strong>of</strong> the forces (2.94)<br />

constraint equations <strong>of</strong> velocities and position, generally stated as<br />

A(x� t) _x + b(x� t) =0� (2.95)<br />

should be added. With mechanical <strong>system</strong>s having so called kinematic chains i.e. parallel<br />

manipulators, such additional equations easily appear if the <strong>motion</strong> <strong>of</strong> some parts <strong>of</strong> the<br />

<strong>system</strong> can not explicitly be written as functions <strong>of</strong> the minimal number <strong>of</strong> coordinates<br />

(amounting to the number <strong>of</strong> d.o.f.). This will not be the case with the Stewart platform apart<br />

from the additional coordinate necessary to attain a global description <strong>of</strong> the orientation.<br />

Systems where the constraint equations cannot be integrated to constraint position equations<br />

are called non-holonomic. A parallel manipulator, like the Stewart platform, will occur<br />

to be a holonomic <strong>system</strong>. Since there are kinematic chains, it is <strong>of</strong>ten easier to state the<br />

equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> a parallel manipulator by using constraint equations. In that case the<br />

manipulator is described as a serial <strong>system</strong> (with some <strong>of</strong> the joints disconnected). The parallel<br />

connections are incorporated by adding constraints. A combined differential/algebraic<br />

description results. This kind <strong>of</strong> description causes difficulties (index problems etc., [19])<br />

in simulation and model <strong>based</strong> <strong>control</strong>. With these goals in mind during modelling, it is<br />

more convenient to state the model in explicit differential equations if possible. Starting<br />

with Section 2.4.3 in the second part <strong>of</strong> this section it is shown that this can be done with


64 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

φ<br />

r<br />

x<br />

y<br />

m<br />

Fig. 2.12: Example <strong>system</strong> consisting <strong>of</strong> one mass rotating in the plane on the left. On the<br />

right also a variable length, r, exists to the centre <strong>of</strong> rotation.<br />

the Stewart platform. But first a simplified example will be given in Section 2.4.2 to clarify<br />

the modelling method used.<br />

2.4.2 Example in modelling using Kane’s method<br />

To clarify the method <strong>of</strong> Kane, the dynamics <strong>of</strong> a simple example <strong>system</strong> will be modelled.<br />

The <strong>system</strong> is given in Fig. 2.12. A simple mass, m, can be rotated in the plane around<br />

some point, which is taken as the origin in the inertial frame. It is assumed we do not know<br />

how to define the mass properties in rotation but only know the translational inertial force<br />

equations:<br />

f c =<br />

f x<br />

f y<br />

= ;<br />

m 0<br />

0 m<br />

φ<br />

x<br />

y<br />

r<br />

x<br />

y<br />

m<br />

= ;Mc (2.96)<br />

We, however, want to choose the angular velocity, _ ,defined in Fig. 2.12 as the generalised<br />

velocity for this 1-d.o.f.-<strong>system</strong>. The coordinates, x and y, and their derivatives can be<br />

formulated as a function <strong>of</strong> and _ with<br />

and by differentiation<br />

_c =<br />

c =<br />

_x<br />

_y<br />

x<br />

y<br />

= r cos( )<br />

r sin( )<br />

= ;r sin( )<br />

r cos( )<br />

_ = Jc�<br />

(2.97)<br />

_ : (2.98)<br />

The local inertial forces, f c can be projected along the generalised velocity by<br />

f = J T c� f c (2.99)<br />

Further differentiation <strong>of</strong> (2.98), provides us with the acceleration<br />

c = Jc� + _<br />

Jc�<br />

_ (2.100)


2.4 <strong>Model</strong>ling the mechanical part <strong>of</strong> the <strong>system</strong> dynamics 65<br />

with<br />

_<br />

Jc� =<br />

;r cos( )<br />

;r sin( )<br />

_ : (2.101)<br />

This equation can be filled into (2.96), which can be filled into (2.99) to arrive at the simple<br />

f = ;mr 2 � (2.102)<br />

since J T c� Jc�<br />

_ =0and J T c� Jc� = r2 in this example. The equilibrium equation <strong>of</strong> <strong>motion</strong>,<br />

considering an active force already projected along _ , i.e. a moment fa, will be<br />

f + f = fa ; mr 2 =0 (2.103)<br />

In principle, also _c could have been defined as the generalised velocity. The only constraint<br />

in this choice is the ability to describe the <strong>motion</strong> <strong>of</strong> any body with this set <strong>of</strong> velocities.<br />

In that case (2.96) would give the inertial forces and an additional constraint equation<br />

_c T c =0 (2.104)<br />

would have to be satisfied. I.e. a combined algebraic and differential set <strong>of</strong> equations would<br />

result. Simulating such a <strong>system</strong> is not trivial. E.g. in the example any finitely small step<br />

satisfying (2.104), a velocity orthogonal to c, would enlarge the radius r.<br />

In the case sketched, it is relatively easy to choose a minimal number <strong>of</strong> coordinates<br />

and end up with differential equations only. But already with planar <strong>system</strong>s such as the<br />

manipulator given in Fig. 2.6, it is not clear how to choose three coordinates (minimal local<br />

degrees <strong>of</strong> freedom) which globally describe this <strong>system</strong>.<br />

The method presented seems somewhat overdone if applied on such an example but it<br />

can easily be extended to the more complex cases. Consider for example the <strong>system</strong> <strong>of</strong><br />

Fig. 2.12 with an variable radius (e.g. due to a non rigid spring/damper) connection.<br />

If the force <strong>of</strong> the spring always acts along the radial direction it could be appropriate to<br />

choose the polar generalised velocities, _p, consisting <strong>of</strong> _r and _ .<br />

Now<br />

_c = Jxp _p =<br />

and the inertial forces follow from<br />

and using<br />

cos( ) ;r sin( )<br />

sin( ) r cos( )<br />

_r<br />

_<br />

(2.105)<br />

f p = ;J T xpm(Jxpp + _<br />

Jxp _p) (2.106)<br />

Jxp<br />

_ = ; sin( ) _ ;r cos( ) _ ; _r sin( )<br />

cos( ) _ ;r sin( ) _ +_r cos( )<br />

gives the inertial forces more explicitly as<br />

f xp = ;<br />

m 0<br />

0 mr 2<br />

r ; ;mr _2<br />

2mr _ _r<br />

(2.107)<br />

: (2.108)


66 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

In this case nonlinear velocity forces like the quadratic centripetal force, mr _2 , and the<br />

velocity product coriolis force, 2mr _ _r, naturally show up.<br />

Including an additional body is relatively easy. E.g. the inertia, i tube <strong>of</strong> the tube drawn<br />

in Fig. 2.12, can be taken into account by adding the inertial force, f tube = itube .<br />

In the calculations, the projections using the jacobians <strong>of</strong>ten enable geometric simplifications.<br />

This will also be the case in using this method in modelling the Stewart platform<br />

including the actuator inertial properties.<br />

With this example, the method <strong>of</strong> Kane has been made explicit and will be used in the<br />

next sections to derive a model <strong>of</strong> the mechanical part <strong>of</strong> the Stewart platform <strong>motion</strong> <strong>system</strong><br />

dynamics.<br />

2.4.3 Stewart platform dynamics<br />

The general procedure <strong>of</strong> stating the equations <strong>of</strong> <strong>motion</strong> has now been given. In this part,<br />

this will be applied in modelling the Stewart platform. Kane’s method <strong>of</strong> projecting local<br />

semi-equilibria (equations <strong>of</strong> <strong>motion</strong>) will be used to arrive at a compact description. To<br />

state the local equations, both Lagrange and Newton-Euler are applied wherever either one<br />

is most appropriate.<br />

With the choice <strong>of</strong> platform position/orientation as the generalized coordinates, all equilibria<br />

can be written as explicit functions <strong>of</strong> these coordinates (and its derivatives). The<br />

modelling approach, however, still enables a clear assignment <strong>of</strong> the contribution each part<br />

<strong>of</strong> the <strong>system</strong> has on the over all dynamics.<br />

First a simplified model, with the platform as the only (rigid) body, will be derived.<br />

Then, the influence <strong>of</strong> the actuator inertial forces is quantified.<br />

2.4.4 Basic Stewart platform mechanical model dynamics<br />

The basic structure <strong>of</strong> the Stewart platform model results if one considers the platform<br />

alone, not taking into account the inertial forces <strong>of</strong> the actuators. Since the <strong>system</strong> in this<br />

case consists <strong>of</strong> only one body, the equations <strong>of</strong> <strong>motion</strong> are easily derived with Newton-<br />

Euler taking the velocity <strong>of</strong> the platform coordinates as the generalized speed. With the<br />

impulse law, (2.92), the translational <strong>motion</strong> is described<br />

Lnfa + mcg = mcc� (2.109)<br />

where fa are the active forces generated by the actuators and g is the gravity vector.<br />

And with the impulse moment law, (2.93), the rotational <strong>motion</strong> is described by<br />

TA m Ln fa = Ic _! + ~ Ic!� (2.110)<br />

with Ic = TI m c T T . Combining these two results in the simplified model <strong>of</strong> the Stewart<br />

platform.<br />

Ln<br />

TA m Ln<br />

fa = mcI 0<br />

0 Ic<br />

c<br />

_!<br />

+ 0 0<br />

0 ~ Ic<br />

_c<br />

! ; mcg<br />

0<br />

(2.111)


2.4 <strong>Model</strong>ling the mechanical part <strong>of</strong> the <strong>system</strong> dynamics 67<br />

In short<br />

J T l�x (sx)fa = Mc(sx)x + Cc( _x� sx) _x + Gc� (2.112)<br />

from which the mass matrix, Mc, the coriolis/centripetal effects, Cc, and gravity, G, are<br />

clearly separated and the structure given in (2.90) is visible. Further, the jacobian, J l�x,<br />

already playing an important part in the sections on kinematics, also occurs in the equations<br />

<strong>of</strong> <strong>motion</strong> directing the actuator driving forces.<br />

In many cases, especially with large and heavy <strong>motion</strong> <strong>system</strong>s, these equations will sufficiently<br />

describe the rigid body dynamics <strong>of</strong> the mechanical part <strong>of</strong> the <strong>system</strong>. In practice,<br />

these dynamics will interact foremost with the hydraulic actuators, which will be modelled<br />

in the next chapter. This chapter concludes with some sections on extending the mechanical<br />

model.<br />

2.4.5 Influence <strong>of</strong> the actuator inertial forces<br />

In some cases, the inertial forces (i.e. mass properties) <strong>of</strong> the actuators can not be neglected.<br />

With many heavy <strong>flight</strong> simulation <strong>system</strong>s they can, but with the light weight <strong>motion</strong> <strong>system</strong>s<br />

requiring high performance <strong>motion</strong> this has to be reevaluated. Explicitly taking into<br />

account the actuator inertial forces will complicate the model. However, this model can be<br />

used to attain an appropriate approximate version as will be shown in the next chapter.<br />

Most <strong>of</strong> the assumptions considering the actuators have already been given considering<br />

their kinematical properties in Section 2.2.11. See e.g. Fig. 2.8. The gimbals are assumed to<br />

rotate frictionless. The inertia <strong>of</strong> both actuator parts rotating around the actuator direction<br />

is neglected and further assumed symmetric w.r.t. this axis. The mass properties <strong>of</strong> each<br />

actuator part can therefore be described by two parameters, the mass and the inertia around<br />

an axis orthogonal to the actuator direction. Mass m a and inertia ia are taken for the upper<br />

moving part <strong>of</strong> the actuator (mainly the piston), and (m b, ib) for the lower rotating part, the<br />

cylinder.<br />

The inertial forces <strong>of</strong> the actuators can be split up in three parts: the gravitational forces,<br />

the inertial mass forces and the influence <strong>of</strong> its inertia. These forces will first be projected<br />

on the upper gimbal points. Any force generated at this point is easily projected at the<br />

generalized platform velocities with (2.47) by<br />

f x a = J T a�xf a a : (2.113)<br />

The gravitational forces are easily projected along the platform velocities. The lower part<br />

<strong>of</strong> the actuator using (2.50),<br />

f a bg = J T bc�ambg = mbrb<br />

j l j Plng = Gmb : (2.114)<br />

Clearly, in a position where the gravity vector is directed along the actuator ( ln) this force<br />

will not contribute. The contribution <strong>of</strong> the moving part is in that case maximal at a as is<br />

shown by equivalently using (2.49),<br />

f a ag = J T ac�amag = ma(I ; ra<br />

Pln )g = Gma<br />

(2.115)<br />

j l j


68 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

The inertial force generated by the mass at the c.o.g. <strong>of</strong> the moving part <strong>of</strong> the actuator is<br />

easily described w.r.t. a frame at its c.o.g.<br />

fac = ma _vac<br />

(2.116)<br />

Projection <strong>of</strong> this force at the upper gimbal point using (2.49) and writing _vac as a<br />

function <strong>of</strong> _va using (2.55) and (2.57), results in<br />

f a ma = ma(I ; Pln<br />

+ ma<br />

ra<br />

2<br />

(j l j;ra)<br />

+<br />

j l j2 Pln ) _va ::: (2.117)<br />

j l j 2 j Plnva j 2 ln +2ma<br />

= Ma _va + Cmava<br />

(j l j;ra)ra<br />

j l j3 (l T n va)Plnva<br />

The first term consists <strong>of</strong> a part in the direction <strong>of</strong> the actuator (I ; P ln) where the mass<br />

directly acts on the gimbal point. Perpendicular to this direction the influence gets smaller<br />

with the squared ratio <strong>of</strong> distances to the lower gimbal point. The second and third term are<br />

the quadratic velocity dependent coriolis and centripetal force.<br />

The inertial forces generated by the inertias <strong>of</strong> the lower and upper part <strong>of</strong> the actuator<br />

can be taken together considering Lagrange and observing that their contribution to the<br />

kinetic energy is equivalent. With (2.48)<br />

Kia�ib<br />

= 1<br />

2 !T a !a(ia + ib) (2.118)<br />

= 1<br />

2 vT a<br />

(ia + ib)<br />

j l j2 Plnva = 1<br />

2 vT a M a ia�ibva The derivation <strong>of</strong> the following two equations can be found in Appendix B together with<br />

a proper choice for factoring the quadratic velocity terms. p a is the position <strong>of</strong> the upper<br />

gimbal point.<br />

d<br />

dt (Mia�ib )=; (ia + ib)<br />

j l j3 Kia�ib<br />

pa<br />

(2v T a lnPln + Plnval T n + lnv T a Pln )� (2.119)<br />

= ; (ia + ib)<br />

j l j3 ((l T n va)Plnva + lnv T a Plnva)� (2.120)<br />

With Lagrange (d=dt(M ) _v ; @K=@p g a ), (2.119) and (2.120) the inertial forces at p a result<br />

f a ia�ib = (ia + ib)<br />

j l j 2<br />

Pln _va ; 2(ia + ib)<br />

j l j3 (l T n va)Plnva = Mia�ib _va + Cia�ibva: (2.121)<br />

The contribution to the mass matrix only exists at <strong>motion</strong> perpendicular to the direction <strong>of</strong><br />

the actuator. Next to this, only coriolis and no centripetal terms appear. The coriolis force<br />

is generated as a result <strong>of</strong> the inertia points in the opposite direction as the one generated by


2.4 <strong>Model</strong>ling the mechanical part <strong>of</strong> the <strong>system</strong> dynamics 69<br />

the mass. This is due to the fact that the influence <strong>of</strong> the inertia decreases while that <strong>of</strong> the<br />

mass increases as the actuator gets longer.<br />

All the separate inertial force contributions <strong>of</strong> each actuator represented by f a bg , f a ag ,<br />

f a ma and f a ia�ib together with the active actuator force generated by the hydraulics, result in<br />

total forces at the upper gimbals. Projection to the platform coordinates via the jacobian,<br />

Ja�x, puts these forces into place for the equations <strong>of</strong> <strong>motion</strong>.<br />

2.4.6 The Stewart platform model including actuator inertia<br />

The equation <strong>of</strong> <strong>motion</strong> <strong>of</strong> the Stewart platform including the inertia <strong>of</strong> the actuators can<br />

still be put in form <strong>of</strong><br />

Where Mt, Ct and Gt are given by<br />

Ct _x = Cc _x +<br />

6X<br />

i=1<br />

J T l�x (sx)fa = Mt(sx)x + Ct( _x� sx) _x + Gt(sx) (2.122)<br />

Mt = Mc +<br />

6X<br />

i=1<br />

J T ai�x (Mma�i + Mia�i�ib�i )Jai�x<br />

(2.123)<br />

(J T ai�x (Cma�i + Cia�i�ib�i )Jai�x _x; j! j 2 (Mma�i<br />

Gt = Gc +<br />

6X<br />

i=1<br />

+ Mia�i�ib�i )P!ai)<br />

(2.124)<br />

J T ai�x (Gma�i + Gmb�i ) (2.125)<br />

In this model, the platform coordinates, velocities and acceleration are the only variables<br />

and the effect <strong>of</strong> each term (mass, coriolis, centripetal, gravity, driving forces) and parameter<br />

(mass, inertia, centres <strong>of</strong> gravity, gimbal point) can clearly be distinguished. The model is<br />

well defined in each state except for those in which any <strong>of</strong> the actuator lengths becomes<br />

zero. Note that this puts an implicit limitation to the validity <strong>of</strong> the model.<br />

Adding the inertial influence <strong>of</strong> the actuators to the simplified model did not change the<br />

compact form <strong>of</strong> six coupled second order differential equations. The equations, however,<br />

became much more complex which is not favourable in model <strong>based</strong> <strong>control</strong> in which the<br />

model has to be calculated at high speed.<br />

Ji claims in [61] that the actuator inertial effects can be seen as a change <strong>of</strong> the platform<br />

mass, inertia and c.o.g. This claim should be carefully interpreted as this change is not only<br />

dependent on the position, but also on the direction <strong>of</strong> the <strong>motion</strong> i.e. not even valid at one<br />

operating point. E.g. in case <strong>of</strong> the Simona <strong>flight</strong> <strong>simulator</strong>, the mass <strong>of</strong> the actuators add<br />

more to the mass matrix <strong>of</strong> <strong>simulator</strong> in heave than in the lateral directions <strong>of</strong> surge and<br />

sway.<br />

With the equations given, it is possible to give bounds on the forces not taken into account<br />

if the inertial forces <strong>of</strong> the actuators would be neglected. Although with conventional


70 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

<strong>motion</strong> <strong>system</strong>s this is <strong>of</strong>ten justified, the tendency towards light weight platforms makes<br />

the actuator inertial forces more evidently come into play. Total neglection would result<br />

in a too rough approximation in that case. Approximation by a constant additive term will<br />

shown to be more convenient.<br />

2.4.7 Parasitic mechanical aspects: modelled as linear flexibility<br />

The models presented only take into account the mechanical <strong>system</strong> as the rigid bodies.<br />

Together with the hydraulic actuators, this forms the most relevant part <strong>of</strong> the <strong>system</strong> dynamics.<br />

However, it appeared that two kinds <strong>of</strong> parasitic mechanical effects due to flexibility<br />

could not be neglected.<br />

First, the foundation on which the Simona <strong>motion</strong> <strong>system</strong> was tested in the Central<br />

Workshop was not entirely rigidly attached to the inertial world. Though relatively heavy<br />

(45 tons), the foundation formed by a block <strong>of</strong> concrete appeared to be in <strong>motion</strong> in the<br />

plane orthogonal to z if the <strong>simulator</strong> <strong>motion</strong> contained too much energy in the frequency<br />

area above > 5 Hz [41]. The foundation was not ideal for these tests, but in countries<br />

like the Netherlands, where buildings are <strong>of</strong>ten build on large piles in weak ground, similar<br />

effects can be expected if no special precautions are being taken.<br />

Flexible effects like this, can be modelled by considering additional mass/ spring/ damper<br />

<strong>system</strong>s in connection with the rigid body model. In Fig. 2.13, a representation <strong>of</strong> a foundation<br />

is given, which has two translational and one torsional spring/damper pot connecting<br />

the mass/inertia <strong>of</strong> the foundation to the inertial world. Further, the <strong>motion</strong> <strong>system</strong> introduces<br />

forces through the lower gimbal points.<br />

If the foundation is given, the three degrees <strong>of</strong> freedom: x f along the surge (x-)direction,<br />

yf along the sway (y-)direction and the angle f around the z-axis (yaw-<strong>motion</strong>), three<br />

additional equations <strong>of</strong> <strong>motion</strong> will result. At first instance, the interaction <strong>of</strong> the foundation<br />

with the inertial world will be modelled by decoupled spring/damper connections with<br />

spring stiffnesses cxf, cyf, c f, and damping coefficients bxf , byf, b f, which are used for<br />

the respective directions. Now, the equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> the foundation with mass m f and<br />

inertia If can be modelled by using the additional matrix equation:<br />

2<br />

4 mf 0 0<br />

0 mf 0<br />

0 0 If<br />

32<br />

5<br />

4 xf<br />

yf<br />

f<br />

3<br />

5 +<br />

2<br />

4 bxf 0 0<br />

0 byf 0<br />

0 0 b f<br />

2<br />

4 cxf 0 0<br />

0 cyf 0<br />

0 0 c f<br />

32<br />

5<br />

32<br />

5<br />

4 _xf<br />

_yf<br />

_ f<br />

4 xf<br />

yf<br />

f<br />

3<br />

5 + (2.126)<br />

3<br />

5 = J T l�xf fa<br />

The jacobian <strong>of</strong> the general (six) foundation body parameters, x fg to actuator lengths<br />

has a similar structure as Jl�x.<br />

Jl�xfg = ;LT n ;(B Ln) T (2.127)<br />

In case only the aforementioned three degrees <strong>of</strong> freedom in the plane are possible a second<br />

projection is required through the selection matrix P f with unit vectors selecting x, y and


2.4 <strong>Model</strong>ling the mechanical part <strong>of</strong> the <strong>system</strong> dynamics 71<br />

Fig. 2.13: Physical representation <strong>of</strong> the model used for the flexible attached foundation<br />

.<br />

P T f =<br />

2<br />

4<br />

1 0 0 0 0 0<br />

0 1 0 0 0 0<br />

0 0 0 0 0 1<br />

Jl�xf = Jl�xfgPf (2.129)<br />

3<br />

5 (2.128)<br />

Finally, the equation for the platform velocity has to be adapted since b will start to vary.<br />

_<br />

l = L T n (Va ; Vb) (2.130)<br />

With the lower gimbal point velocities, _vb stacked in Vb. If the model with the platform<br />

alone (or a constant mass matrix with respect to the moving <strong>simulator</strong>) is taken, the model<br />

equation (2.111) does not have to be adapted. All together these equations form a model <strong>of</strong><br />

the <strong>system</strong> with a platform attached to a non-inertial ground, which has the same validity as<br />

the simple model (for all <strong>simulator</strong> poses except those for which (klk i =0). This problem<br />

has not seen to be considered in literature apart from cases in space technology [93, 108] in<br />

which the moving platform has very limited manoeuvrability and the actuators are used to<br />

supress vibrations.<br />

At first sight, the foundation, though also moving if the actuators force the <strong>simulator</strong><br />

to move, does not have too much direct interaction with <strong>simulator</strong> <strong>motion</strong>, since the basic<br />

mechanical model equation (2.112) looks the same. Indirectly, the following dependencies<br />

can be observed.


72 2 Mechanics <strong>of</strong> parallel driven <strong>motion</strong> <strong>system</strong>s<br />

- The <strong>motion</strong> <strong>of</strong> the foundation has influence on the actuator velocities. The hydraulic<br />

actuators form a port connection with the mechanical <strong>system</strong> through the active forces<br />

and velocity. Generally, the active force will be influenced by the velocity due to a<br />

coupling in the dynamics <strong>of</strong> the (<strong>control</strong>led) hydraulic <strong>system</strong>. This influence will<br />

therefore affect the <strong>simulator</strong> <strong>motion</strong>. Note that this influence will be much larger if<br />

the actuators are position/velocity <strong>control</strong>led than with force <strong>control</strong> in the relevant<br />

frequency area.<br />

- The platform pose can not be calculated from the actuator lengths anymore without correcting<br />

for the moving foundation.<br />

- The jacobian, Jl�x, depends on the <strong>motion</strong> <strong>of</strong> the foundation since it depends on the unit<br />

direction vectors <strong>of</strong> the actuators. These directions will change as the foundation<br />

moves resulting in change <strong>of</strong> direction <strong>of</strong> the active forces acting on the <strong>simulator</strong>.<br />

This is a second order effect and is mostly compensated for by reconstructing the<br />

Jacobian from the actuator length measurements. The <strong>motion</strong> in the plane <strong>of</strong> the<br />

foundation will not influence the forces necessary to compensate for gravitational<br />

effects <strong>of</strong> the gross moving platform.<br />

- The <strong>motion</strong> <strong>of</strong> the actuator (inertias) is directly influenced by the foundation. The equations<br />

<strong>of</strong> <strong>motion</strong> for these parts have to be adapted and become somewhat more complex.<br />

If the actuator inertia can be neglected or taken as part <strong>of</strong> the constant mass<br />

matrix w.r.t. moving platform, this effect can also be neglected.<br />

Given influences, both in the choice <strong>of</strong> the <strong>control</strong> strategy as in identifying the dynamics<br />

<strong>of</strong> the <strong>system</strong>, the parasitic behaviour <strong>of</strong> the foundation should be taken into account.<br />

The second important parasitic mechanical effect to be analyzed are the structural modes<br />

<strong>of</strong> the <strong>simulator</strong> itself. Especially in optimising the ability to realistically simulate the high<br />

frequency vibrations in a vehicle, the unwanted resonances <strong>of</strong> the <strong>simulator</strong> should not be<br />

hit up to noticeable level. In the design <strong>of</strong> the shuttle <strong>of</strong> the Simona <strong>simulator</strong>, the ratio<br />

<strong>of</strong> stiffness and mass has been optimized [8]. However, finite element models pointed out<br />

that with this <strong>system</strong> one will not be able to extend the lowest structural vibrations above<br />

15-20 Hz<br />

As a rule <strong>of</strong> thumb the flexible modes should only appear in a frequency area well<br />

distinct from the area where the <strong>motion</strong> to be simulated has relevant energy and thus usually<br />

form an upper bound on the achievable bandwidths. Since there are only a limited number<br />

<strong>of</strong> actuators, the moving platform loses functional <strong>control</strong>lability if structural deformation<br />

takes place. In theory, one could well be able to apply a desired acceleration pr<strong>of</strong>ile to the<br />

pilots head but one would in these cases structurally have to settle for vibration in other parts<br />

<strong>of</strong> the platform e.g. the large visual projection screen. Additional active damping devices<br />

can probably be <strong>of</strong> help in making this restriction less tight.<br />

The unneglectable deformations in the moving platform drastically change the structure<br />

<strong>of</strong> the <strong>system</strong>. <strong>Model</strong>ling using the projection methods like Kane’s is still possible if one can<br />

approximate the modes by splitting up the previously rigid bodies in a limited number mass/<br />

spring/ damper <strong>system</strong>s i.e. smaller rigid parts connected by springs and dampers ([146]).<br />

Deriving these equations by hand, however, becomes very tedious and does in general not


2.5 Chapter Resume 73<br />

result in a model from which analysis can take place more easily than from automatically<br />

derived models. In a master’s project as part <strong>of</strong> this research Rijnten [119] modelled the<br />

<strong>system</strong>, including a structural vibration, using the symbolic equation modelling s<strong>of</strong>tware,<br />

Autolev [65]. Though simulation with such models can clarify <strong>system</strong>s behaviour, not much<br />

structural understanding <strong>of</strong> <strong>system</strong>s like a deformed Stewart platform can be gained since<br />

the number <strong>of</strong> variables in the model well goes above a thousand.<br />

Other, more advanced, methods in modelling flexible structures [62, 63, 66, 129] <strong>of</strong>ten<br />

used in space applications with large flexible deformations, usually result in models, which,<br />

due to their complexity, can not directly be used in model <strong>based</strong> <strong>control</strong>. It was decided to<br />

do the actual analysis <strong>of</strong> the structural deformations in the Simona <strong>motion</strong> <strong>system</strong> using the<br />

actual data measured at the <strong>system</strong>. This will be discussed in Chapter 4. Further analysis<br />

and also simulation <strong>of</strong> the models derived will take place after introducing and integrating<br />

the models <strong>of</strong> the hydraulic <strong>system</strong> to be discussed in the next chapter. This will be more<br />

realistic since the integration <strong>of</strong> the hydraulics and the mechanics forms the most relevant<br />

part <strong>of</strong> the dynamics considering the whole <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>.<br />

2.5 Chapter Resume<br />

Since the kinematic model <strong>of</strong> the Stewart platform construction is fully parallel, the platform<br />

(end effector) coordinates and velocities can globally be used to describe the state <strong>of</strong><br />

the mechanical <strong>system</strong>. It appeared most convenient to state orientation applying an euler<br />

parameter (quaternion) like parametrization. Apart from the numerical advantages, an algorithm<br />

could be defined to evaluate the appearance <strong>of</strong> singular points in the work space <strong>of</strong> a<br />

parallel manipulator like the Stewart platform.<br />

With this method, it finally became possible to pro<strong>of</strong> that the SRS kinematical construction<br />

has no singular points. This is not trivial, since it would have been possible to reach<br />

such a point with only 15 % more stroke <strong>of</strong> the actuators. Moreover, it could be proven that<br />

the iterative Newton-Raphson procedure to calculate the platform pose from the actuator<br />

lengths converges anywhere in the SRS work space during any kind <strong>of</strong> possible (velocity<br />

limited) <strong>motion</strong>. It converges sufficiently fast for safe use in a real time digital feedback<br />

loop as will be done in the model <strong>based</strong> <strong>control</strong>ler discussed in Chapter 5.<br />

The dynamics <strong>of</strong> the Stewart platform was stated as a set <strong>of</strong> differential equations without<br />

algebraic constraints resulting from the kinematic chains in the <strong>system</strong>. This is possible<br />

by writing the actuator <strong>motion</strong> explicitly as a function <strong>of</strong> platform <strong>motion</strong>. By using Kane’s<br />

method <strong>of</strong> projecting forces onto the generalized velocities, the contribution <strong>of</strong> each inertial<br />

or active force e.g. stemming from the actuators piston mass, etc., can be quantified separately<br />

using local (simple) Newton-Euler like equations and (by projection) ending up with<br />

a limited number <strong>of</strong> equations <strong>of</strong> <strong>motion</strong>.<br />

The models derived only describe the dynamics <strong>of</strong> the mechanical <strong>system</strong>. The observed<br />

dynamics <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> will shown to be much better approximated by<br />

models which incorporate both the hydraulic and mechanical characteristics and their (two<br />

sided) interaction. This chapter provided for the mechanical modules for such models. The<br />

next chapter will treat the modelling <strong>of</strong> the hydraulic <strong>system</strong> itself and the interconnected<br />

structure <strong>of</strong> hydraulically driven mechanical <strong>system</strong>s.


Chapter 3<br />

Hydraulically driven <strong>motion</strong><br />

<strong>system</strong>s<br />

In the previous chapter, the mechanical <strong>system</strong> was modelled with input actuator forces<br />

from which accelerations, velocities and positions <strong>of</strong> the <strong>motion</strong> <strong>system</strong> and possibly its<br />

parasitic dynamics could be calculated. The energy necessary to move is supplied by the<br />

actuators. In robotics both electrically as hydraulically driven <strong>system</strong>s are seen to be applied<br />

and the mechanical model can be merged easily to models <strong>of</strong> either kind <strong>of</strong> <strong>system</strong>s.<br />

With the larger <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s almost invariably use is made <strong>of</strong> hydraulic<br />

actuators as is also the case with the Simona <strong>motion</strong> <strong>system</strong>. Since both the energy<br />

and the cooling with fluid power technology can be supplied at an exterior hydraulic pump,<br />

a favourable high ratio <strong>of</strong> force delivered over the weight and size <strong>of</strong> the actuator can be<br />

sustained well into the high power area. But even more importantly, the use <strong>of</strong> hydraulics<br />

enables smoothly running (low friction, low wear) long linear actuators, which is necessary<br />

in simulation requiring an unnoticeable change <strong>of</strong> sign in the direction <strong>of</strong> <strong>motion</strong>.<br />

An important aspect <strong>of</strong> the <strong>motion</strong> <strong>system</strong>s under consideration is the fact that the dynamics<br />

<strong>of</strong> the actuators heavily interact with the mechanics. So modelling either one <strong>of</strong><br />

these phenomena does not provide much insight in the dynamics <strong>of</strong> such <strong>system</strong>s. In most<br />

textbooks treating the modelling and <strong>control</strong> <strong>of</strong> mechanical <strong>system</strong>s the dynamics <strong>of</strong> the<br />

actuators are at best considered as parasitic influences to be taken into account at high frequencies<br />

[11, 31, 113]. However, with (heavy) <strong>system</strong>s requiring a high amount <strong>of</strong> power,<br />

especially if driven hydraulically, the interaction between the actuators and mechanical <strong>system</strong><br />

determines the ’rigid’ resonant modes. By using an integrated model, which will be<br />

presented in Section 3.3, the most relevant dynamics is captured as will be shown in the<br />

subsequent Chapter 4 in which evaluation with the experimental data takes place.<br />

The basic integrated model can still have a relative simple structure although the response<br />

<strong>of</strong> such <strong>system</strong>s may already seem quite complex. The basic structure <strong>of</strong> the hydraulically<br />

driven mechanical <strong>system</strong> allows the derivation <strong>of</strong> some important properties<br />

like passivity, discussed in Section 3.3.1, which can be <strong>of</strong> use in <strong>control</strong> and experimental<br />

identification. In this chapter, the dynamics <strong>of</strong> the hydraulic actuators will be modelled with<br />

a bilateral coupling to the mechanics through the energetic pairs <strong>of</strong> force and velocity at the<br />

mechanical side and flow and pressure in the hydraulical part <strong>of</strong> the <strong>system</strong>.<br />

75


76 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 3.1: Schematic drawing <strong>of</strong> a symmetrical hydraulic actuator.<br />

The long stroke hydraulic actuators, which are typical for <strong>flight</strong> <strong>simulator</strong>s, have been<br />

modelled in detail for <strong>motion</strong> <strong>control</strong> design by Van Schothorst in [124]. In Section 3.1, an<br />

overview will be given <strong>of</strong> the parts <strong>of</strong> those models which proved to be essential as either<br />

being an explicit part in the model <strong>based</strong> <strong>control</strong>ler considered in this research or in analysis<br />

exploring the limitations <strong>of</strong> such <strong>control</strong>lers. This last part usually requires an additional<br />

degree <strong>of</strong> detail in modelling discussed in Section 3.2. In this case, the dynamics <strong>of</strong> the<br />

hydraulic servo valve together with the transmission lines had to be taken into account.<br />

For a more fundamental introduction into hydraulic <strong>system</strong>s, one is referred to textbooks<br />

like [99] and [153]. Most <strong>of</strong> the notation used here follows Van Schothorst [124]. The<br />

modelling for <strong>control</strong> approach in hydraulic <strong>system</strong>s can also be found in references like<br />

[48] and [146]. In these references, however, the dynamics <strong>of</strong> the transmission lines was<br />

assumed to be neglectable. In this research, this was not the case due to the long stroke<br />

and high performance requirements, and therefore a modelling for <strong>control</strong> approach with<br />

the interplay <strong>of</strong> transmission lines dynamics and servo valve as the limiting factor, had to be<br />

chosen.<br />

3.1 The basic structure <strong>of</strong> hydraulic actuators<br />

The basic notion <strong>of</strong> the functionality <strong>of</strong> a hydraulic actuator can be observed by looking<br />

at Fig. 3.1, although the actually used actuator looks far more complicated as shown in<br />

Fig. 3.3. In Fig. 3.1, the piston connected to the load can be made to move by letting oil<br />

flow into the compartment <strong>of</strong> the cylinder on the left, o1, and out <strong>of</strong> the compartment on<br />

the right, o2, or v.v. This flow will result in a pressure build up, P o1 ; Po2, which through<br />

the reflected areas, Ap, will force the load to move. In a symmetrical actuator, this area,<br />

Ap, is equal on each side. In <strong>motion</strong>, the oil flows will also have to compensate for the<br />

change in volume in the respective actuator chambers. This effect accounts for the twosided<br />

coupling <strong>of</strong> flow/ pressure and/ or velocity/ force. Usually one also has to account for<br />

some dissipation resulting from leakage flows.<br />

The dynamics due to the hydraulic part <strong>of</strong> the <strong>system</strong> is caused by the compressibility<br />

<strong>of</strong> the oil. The relative decrease in volume, ; _ V=V, causes a pressure rise, P _ , linear with


3.1 The basic structure <strong>of</strong> hydraulic actuators 77<br />

the effective bulk modulus <strong>of</strong> the oil, E.<br />

P _ = ;E _ V<br />

(3.1)<br />

V<br />

By taking into account the oil flow into and out <strong>of</strong> the compartments, the following equations<br />

can be obtained, which describe the pressure dynamics <strong>of</strong> each chamber.<br />

Po1<br />

_ = E<br />

( o1 ; l1 ; Ap _q)<br />

V1<br />

Po2<br />

_ = E<br />

( l2 ; 02 + Ap _q) (3.2)<br />

V2<br />

The sign <strong>of</strong> the oil flows due to leakage, l1, l2, are taken negative w.r.t. the main flows,<br />

o1, o2, usually <strong>control</strong>led by a valve. The position, q, <strong>of</strong> the piston is set to zero at the<br />

point where the volumes <strong>of</strong> the actuator chambers are equal, V m = V1 = V2. This will<br />

typically be at or near half stroke position as drawn and will be referred to as mid position.<br />

As a result<br />

V1 = Vm + Apq<br />

V2 = Vm ; Apq (3.3)<br />

At this point it is most appropriate to apply a change <strong>of</strong> coordinates by taking mean and<br />

difference pressures, flows and oil stiffnesses (Cm and dC), which are defined as follows<br />

Pm m lm Cm<br />

dP d d l dC =<br />

1<br />

2<br />

1<br />

2<br />

1 ;1<br />

The pressure dynamics <strong>of</strong> (3.2) can now be transformed to<br />

Po1 o1 l1 E=V1<br />

Po2 o2 l2 E=V2<br />

_<br />

dP = 2Cm( m ; lm ; Ap _q) +dC=2(d ; d l)<br />

(3.4)<br />

_<br />

Pm = Cm=2(d ; d l) +dC=2( m ; lm ; Ap _q) (3.5)<br />

The change <strong>of</strong> coordinates can be motivated by the fact that only the pressure difference,<br />

dP , directly affects the mechanics through the actuator force, f a = ApdP . Further analysis<br />

reveals that even more advantage can be taken <strong>of</strong> this transformation. First <strong>of</strong> all, dC =0<br />

in the mid position, which fully decouples the two equations. Also in non mid position,<br />

the term d ; d l is small or even zero which makes the second equation not only badly<br />

observable but also almost un<strong>control</strong>lable. So, in most cases a sufficiently accurate model<br />

can be obtained by only taking into account the structure given by<br />

_<br />

dP =2Cm(q)( m ; LlmdP ; Ap _q)� (3.6)<br />

assuming a laminar leakage flow proportional to the pressure difference, lm = LlmdP .<br />

Equation (3.6) provides one <strong>of</strong> the two main equations to describe the basic hydraulically<br />

driven mechanical <strong>system</strong> as given in the block scheme <strong>of</strong> Fig. 3.2.<br />

Removing the dynamics <strong>of</strong> Pm not only simplifies the model but also reduces problems<br />

(e.g. numerical) in model <strong>based</strong> <strong>control</strong> <strong>of</strong> a hydraulic actuator due to the minor effect on


78 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 3.2: Basic block scheme <strong>of</strong> a hydraulic actuator connected to a mechanical <strong>system</strong>.<br />

the input/ output relation <strong>of</strong> this part <strong>of</strong> the dynamics. This model reduction step is <strong>of</strong>ten<br />

performed without consistently showing what is to be neglected i.e. not considering the<br />

change <strong>of</strong> coordinates.<br />

The first order differential equation is nonlinear due to the position dependence <strong>of</strong> the<br />

oil stiffness, using (3.3) and (3.4),<br />

Cm(q) =<br />

EVm<br />

V 2 m ; A2 p<br />

q2 = E<br />

Vm<br />

which attains the minimum at the mid position (q =0).<br />

3.1.1 Leakage<br />

1<br />

1 ; A 2 p q2 =V 2 m<br />

� (3.7)<br />

The leakage flows are considered laminar, which results in flows lineary dependent on the<br />

pressure drops. Assuming two reference pressures: P s as the higher level supply pressure<br />

and Pt as the lower level tank pressure (which can chosen to be the 0-level), the leakage<br />

flows can depend on five different parameters, L i, given in the following equations.<br />

l1 = L1t(Po1 ; Pt) ; L1s(Ps ; Po1) +Ln(Po1 ; Po2) (3.8)<br />

l2 = ;L2t(Po2 ; Pt) +L2s(Ps ; Po2) +Ln(Po1 ; Po2) (3.9)<br />

Not all terms will appear in every construction. E.g. in the double concentric actuator used<br />

for the SRS, schematically given in Fig. 3.3, the terms Ln and L1s do not exist.<br />

Also in this case it is appropriate to consider a change in coordinates<br />

Ltm Lsm<br />

dLt dLs =<br />

1<br />

2<br />

1<br />

2<br />

1 ;1<br />

L1t L1s<br />

L2t L2s<br />

(3.10)


3.1 The basic structure <strong>of</strong> hydraulic actuators 79<br />

The leakage can now be described as function <strong>of</strong> the new coordinates<br />

lm<br />

d l<br />

=<br />

+<br />

Ln +(Lsm + Ltm)=2 dLt + dLs<br />

dLt + dLs 4(Ltm + Lsm)<br />

;dLs<br />

;dLt<br />

;4Lsm ;4Ltm<br />

Ps<br />

Pt<br />

dP<br />

Pm<br />

(3.11)<br />

It can be analyzed at what value the mean pressure, P m, will stabilize, using these equations<br />

together with (3.5) with _ dP = _ Pm =0.<br />

In dropping Pm, only the main leakage parameter, Llm, influencing the resulting basic<br />

hydraulic equation (3.6), is relevant. It can be constructed as<br />

Llm = Ln +(Lsm + Ltm)=2 (3.12)<br />

3.1.2 Basic hydraulically driven mechanical <strong>system</strong> model<br />

In Fig. 3.2, a 2-port block scheme <strong>of</strong> (3.6) is given in connection with a mechanical <strong>system</strong>.<br />

This mechanical <strong>system</strong> with mass, M, viscous/ damping force constant, B p, and possibly<br />

some additional disturbance forces, Fext, can be described by the following equation<br />

Mq = fa ; Bp _q ; Fext<br />

(3.13)<br />

As all the parameters given are strictly positive, the hydraulic <strong>system</strong> can viewed as a velocity<br />

to force feedback with a first order <strong>system</strong>. The loop gain over theqtwo integrators,<br />

which is usually dominating, determines the undamped eigenfrequency 2CmA2 pM ;1 .<br />

Both leakage, Llm, and viscous friction/ damping, Bp will dissipate energy.<br />

It is not advisable to model the physical hydraulic actuator as a fully separate module.<br />

E.g. some parts <strong>of</strong> the hydraulic <strong>system</strong> to be modelled like the mass <strong>of</strong> the piston/ cylinder<br />

and the viscous friction should be made part <strong>of</strong> the mechanical <strong>system</strong> model to be<br />

connected.<br />

In case <strong>of</strong> the multivariable <strong>flight</strong> <strong>simulator</strong>, basically Fig. 3.2 holds, considering vector<br />

paths and matrices. To include platform coordinates, a mapping using the jacobian has to<br />

be applied. The actuator forces fa = ApdP will effect the platform coordinates through<br />

the transpose jacobian J T l�x and the platform velocities will have to be projected along the<br />

actuators by using _q = _ l = Jl�x _x. The undamped eigenfrequencies <strong>of</strong> the hydraulically<br />

driven mechanical <strong>system</strong> can be predicted by calculating the square roots <strong>of</strong> the eigenvalues<br />

<strong>of</strong> the loop matrix gain, p (2CmA2 pJl�xM ;1J T l�x ), assuming equal Ap and Cm for all<br />

actuators.<br />

If this is not the case one should consider taking matrices with the specific actuator<br />

values for Ap and Cm on the diagonal. All these variables, except for A p, depend on the<br />

<strong>simulator</strong> pose. These mass and oil stiffness dependent eigenvalues will be shown to reflect<br />

the most prominent part <strong>of</strong> the dynamics <strong>of</strong> a hydraulically driven <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong><br />

<strong>system</strong>.<br />

The structure given in Fig. 3.1 is a very rough simplification <strong>of</strong> an actual hydraulic<br />

actuator although the model derived will describe the main dynamical effects which can be


80 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 3.3: Schematic drawing double concentric symmetric hydraulic actuator.<br />

observed in practice. However, the way the <strong>control</strong>led oil flow, m (d ), behaves, has not<br />

yet been given and should also be studied to arrive at an appropriate model. This requires<br />

the modelling <strong>of</strong> the servo valve and transmission lines as will be discussed in the next<br />

section.<br />

3.2 Extensions<br />

In Fig. 3.3, a schematic drawing is given <strong>of</strong> the so-called double concentric symmetric<br />

hydraulic actuator, which is used in case <strong>of</strong> the SRS. This construction enables a symmetric<br />

actuator without the disadvantage <strong>of</strong> a piston at both sides. The arrows are drawn in<br />

the direction where the flows and velocity take a positive sign. Note that e.g. the leakage<br />

flow, l2;, is always negative. Next to the cylinder, the picture shows two other parts <strong>of</strong> a<br />

hydraulic servo actuator, which haven’t been discussed yet, the servo valve and the transmission<br />

lines.<br />

3.2.1 Servo valve<br />

The input flows, i1 and i2, given in Fig. 3.3 are determined by the servo valve. <strong>Model</strong>ling<br />

the relevant dynamics between our <strong>control</strong> input signal (voltage), u, to the servo valve and<br />

these oil flows is the subject <strong>of</strong> this section. The valve consists <strong>of</strong> an pilot valve and main<br />

spool as given in Fig. 3.4. The main spool is <strong>control</strong>led electronically using a <strong>control</strong> scheme<br />

given in Fig. 3.5.<br />

With the external connection to two reference oil compartments to the servo valve, the<br />

oil flow can be <strong>control</strong>led by the valve opening. One external reference is the hydraulic<br />

pump supply pressure, Ps, which is 160 Bar higher than the second reference in case <strong>of</strong> the<br />

SRS. This second reference, the tank pressure, Pt, can be considered zero. Both references<br />

are assumed to be constant in the sequel.


3.2 Extensions 81<br />

Fig. 3.4: Schematic drawing <strong>of</strong> a three stage valve. The main spool position, X m, is measured<br />

electronically and <strong>control</strong>led by feedback to the flapper positioning current,<br />

ia. The flapper/ nozzle mechanism internally stabilizes the positioning <strong>of</strong> the pilot<br />

valve with position, Xs, with oil flows ni. This is used to enable large oil flows<br />

at the main spool, which is positioned with oil flows mi. For a valve model,<br />

describing the most relevant high frequent dynamics as given in (3.17), only a<br />

limited number <strong>of</strong> variables given in the figure have to be taken into account.


82 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 3.5: Valve main spool <strong>control</strong> scheme<br />

Considering a turbulent oil flow through the servo valve, the main behaviour <strong>of</strong> this<br />

flow can be described by a nonlinear algebraic relation depending on spool position and the<br />

square root <strong>of</strong> the pressure drop. In the higher frequency area (> 100 Hz in case <strong>of</strong> the SRS<br />

servo valves), the dynamics <strong>of</strong> the spool positioning mechanism have to be considered too.<br />

The spool is an important part <strong>of</strong> the servo valve. By positioning the spool, the valve<br />

opening can be set. The valve opening and pressure drop over the valve determine the<br />

oil flow through the valve to the actuator cylinder compartment itself. If large oil flows are<br />

required (typically larger than 100 l=min), the spool positioning itself requires an additional<br />

hydraulic circuit, in itself an extra servo actuator <strong>control</strong>led by a pilot spool. This multistage<br />

concept is applied in the SRS as flows up to 150 l=min are needed.<br />

A schematic drawing <strong>of</strong> the valve is given in Fig. 3.4. The pilot spool (with position<br />

xs), is positioned by an internal hydromechanical feedback mechanism called the flapper/<br />

nozzle. As already noted, the positioning <strong>of</strong> the main spool is performed by an electronic<br />

feedback given in Fig. 3.5. This feedback uses the main spool position x m, or positioning<br />

error m to the current ia which forces the flapper position, xf , aside.<br />

Over a broad frequency area, usually including the area <strong>of</strong> interest in positioning and/<br />

or forcing a mechanical <strong>system</strong> (servo-valve bandwidth in case <strong>of</strong> the SRS is 150 Hz), the<br />

transfer function from desired to actual spool position, x m, can be considered a static unity<br />

gain.<br />

So the main servo valve phenomena can be captured by considering the turbulent flow<br />

equations <strong>of</strong> the main four-way valve. With symmetric hydraulic actuators, one usually<br />

applies symmetric 4-way valves. The main spool position, x m, together with the pressure<br />

drops determine the flows, i1 and i2, to and from the actuator chambers from P s and to<br />

Pt, as is drawn in Fig. 3.4.<br />

The equation describing a turbulent flow restriction is given by Merrit [99]<br />

=CdAm(x)<br />

s 2(Pin ; Pout)<br />

(3.14)<br />

In this case the discharge coefficient, Cd, and the oil density, , can be considered constant.<br />

The geometric properties, Am(x) <strong>of</strong> the flow restriction depend (almost linearly) on the<br />

spool position. The flow gain evaluated w.r.t. the spool position varies with the square root<br />

<strong>of</strong> the pressure drop and, moreover, this results in a hard non-linear change in gain if the<br />

flow changes sign and Pi1 or Pi2 are not both equal to 1=2Ps.


3.2 Extensions 83<br />

x_s, P_n, f_p, f_v<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

P_n<br />

Step response pilot valve<br />

−0.5<br />

0<br />

f_v<br />

0.001 0.002 0.003 0.004 0.005<br />

time (s)<br />

0.006 0.007 0.008 0.009 0.01<br />

Fig. 3.6: A state step response <strong>of</strong> the model (3.17) <strong>of</strong> the pilot valve with the identified<br />

parameters given in Table 3.1.<br />

As observed earlier, a more compact description can be arrived at at the actuator side in<br />

applying a change <strong>of</strong> coordinates, using difference and mean flows and pressures. This is<br />

also the case with the valve. With the important additional assumptions that<br />

x_s<br />

Pi1 + Pi2 = Ps� (3.15)<br />

and unaltered geometric properties A(xm), and considering normalised pressures and flows,<br />

^P = P=Ps, ^xm = xm=xx�max, ^ = = max, the maximum flow can be found by calculating<br />

or measuring max := f (xm�dPi�nom) =f (xm�max� 0). Now, one equation describes<br />

the oil flow through the four-way valve<br />

q<br />

^<br />

im = f2(^xm) 1 ; sgn(^xm)d ^ Pi� (3.16)<br />

where the mean normalised valve flow is defined by ^ im =( ^ i1 + ^ i2)=2. Again note that,<br />

due to the sign convention taken, this mean valve flow, im, means the average amount <strong>of</strong><br />

oil going into transmission line 1 and out <strong>of</strong> transmission line 2.<br />

With this equation, as proposed in [124] by Van Schorhorst, static geometric nonlinearities<br />

can be captured by the function f 2(^xm), which have been identified experimentally for<br />

all the six SRS servo valves. The sign-function (sgn) accounts for the switch in ports at<br />

the four-way valve. Note that a reduction step by omitting one state in taking away mean<br />

pressure and difference flow at the actuator is also motivated at the input, the valve, since<br />

these variables are not involved in (3.16).<br />

The main limitation in only using (3.16) in describing a valve lies in the high frequency<br />

area, as the spool position, xm, can not be positioned infinitely fast. The positioning mechanism,<br />

as depicted in Fig. 3.4, can be modelled following Van Schothorst [124] by using the


84 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

variable value variable value<br />

c1 4:29e3 c6 1e4<br />

c2 2:96e7 c7 9:93e-1<br />

c3 2:3e10 c8 4e3<br />

c4 1:13e11 c9 2:68e10<br />

c5 1e2 c10 1:27e2<br />

Kp 8 Kd 4:9e-4<br />

Table 3.1: The Simona three stage valve parameters identified by Van Schothorst [124]<br />

state space equation with normalized states given for the pilot valve:<br />

2<br />

6<br />

4<br />

^xf<br />

_^xf<br />

_^xs<br />

d _ ^ P n<br />

3<br />

7<br />

5 =<br />

2<br />

6<br />

4<br />

;c1 ;c2 ;c3 ;c4<br />

1 0 0 0<br />

0 ;1 ;c5 c6<br />

0 c7 1 ;c8<br />

3 2<br />

7<br />

6<br />

5<br />

6<br />

4<br />

_^xf<br />

^xf<br />

^xs<br />

d ^ Pn<br />

3<br />

7<br />

5 +<br />

2<br />

6<br />

4<br />

c9<br />

0<br />

0<br />

0<br />

3<br />

7<br />

5 ia<br />

(3.17)<br />

All the coefficients, c1:::9 are positive and the numerical values, identified on the actual<br />

actuators <strong>of</strong> the SRS by Van Schothorst [124], can be found in Table 3.1. The model takes<br />

into account the force balance and acceleration <strong>of</strong> the flapper, x f , the limited stiffness <strong>of</strong><br />

the oil at Pn and a reduced force balance (velocity only) <strong>of</strong> the pilot spool, x s. The model<br />

incorporates a right half plane zero as the flapper forces the pilot spool to move in negative<br />

direction, which can be compensated only after pressure build up has taken place.<br />

Further, the dynamics <strong>of</strong> the SRS pilot valve has four poles, among which is one lightly<br />

damped resonance, at ca. 900 Hz and one pole at ca. 150 Hz. The step response <strong>of</strong> the pilot<br />

valve is given in Fig. 3.6. Next to each other the flapper velocity (f v), position, xf , and<br />

pressure difference peak, Pn, in resonating at 880 Hz. The pilot spool moves slightly non<br />

minimum phase and resonating along a first order response in approx. 6 ms to its final value.<br />

In this simulation the four states have been scaled to achieve a numerically more favourable<br />

description. xsc = diag(1e ; 7� 1e ; 3� 1� 10): The pilot spool lets oil flow ( m) tomove<br />

the main spool. Considering a dPm 0, this can be modelled by a additional equation for<br />

the third stage adding an integrator.<br />

_^xm = c10f1(^xs) (3.18)<br />

Also the quasi static functions f1(^xs) and constants c10, have all been experimentally identified<br />

[124].<br />

As already noted, the position <strong>of</strong> the main spool is <strong>control</strong>led with an electronic PD<strong>control</strong>ler<br />

(Fig. 3.5) with proportional (K p) plus derivative (Kd) feedback given by<br />

ia = Kp ^xm ; Kd _ ^xm = Kp(u ; ^xm) ; Kdc10f1(^xs) (3.19)<br />

where u is the input signal sent to the valve.<br />

The additional state <strong>of</strong> this <strong>system</strong> in feedback results in a slightly resonating behaviour<br />

( =0:3) at 150 Hz as opposed to the mainly first order like response <strong>of</strong> the pilot valve.


3.2 Extensions 85<br />

x_m, x_s, P_n, f_p, f_v<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

x_s<br />

Step response three stage valve<br />

x_m<br />

−0.2<br />

f_v<br />

−0.4<br />

P_n<br />

−0.6<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

time (s)<br />

Fig. 3.7: <strong>Model</strong> state step response <strong>of</strong> the three stage servo valve.<br />

As can be seen from the state step response <strong>of</strong> the valve in Fig. 3.7, the high frequency<br />

behaviour can hardly be noted in the input/ output relation. Up till 200 Hz, the valve can<br />

also well be approximated by a second order <strong>system</strong> (e.g. via balanced reduction), which<br />

will lack physical interpretation, however. Taking a bode plot <strong>of</strong> the linearised version <strong>of</strong><br />

the valve shows the close correspondence as shown in Fig. 3.8.<br />

In hydraulic servo actuators, the valve characteristics are <strong>of</strong>ten limiting the achievable<br />

performance <strong>of</strong> the <strong>system</strong>. In this case, the high bandwidth valves still introduce a phase<br />

lag <strong>of</strong> approx. 10 at 25 Hz and double this at 50 Hz. The combination with the long<br />

transmission lines, which will be described next, form a risk to stability.<br />

3.2.2 Transmission lines<br />

A disadvantage <strong>of</strong> requiring long stroke actuators is the unavoidable use <strong>of</strong> relatively long<br />

transmission lines. Transmission lines <strong>of</strong>ten show a lightly damped resonating behaviour<br />

and if the transmission lines are longer, the eigenfrequencies tend to be lower which is a<br />

disadvantage in striving towards a smooth high bandwidth <strong>control</strong> by the servo valve <strong>of</strong> the<br />

flow to and from the actuator. Moreover, if actuator pressure <strong>control</strong> (dP o) is intended to<br />

be applied measuring the pressure, dPi, at the valve, the transmission line dynamics are in<br />

between. It will be shown that the pressure at the other side <strong>of</strong> a long transmission line, the<br />

output pressure, Po1�2, <strong>of</strong> the actuator, can deviate considerably from the measured input<br />

pressure, Pi1�2, at frequencies ten times lower than the transmission line eigenfrequencies.<br />

The transmission line dynamics can <strong>of</strong>ten be neglected if properly designed and with<br />

moderate performance requirements. In [48, 146] practical examples can be found <strong>of</strong> hydraulically<br />

driven mechanical <strong>system</strong>s (Brick laying robot, RRR-robot) in which the transmission<br />

lines can be omitted in modelling and <strong>control</strong>. In that case, the bandwidth <strong>of</strong> the<br />

servo-valve can be made reasonably lower than the lowest resonance resulting from the


86 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Amplitude<br />

Phase (deg)<br />

10 1<br />

10 0<br />

10 −1<br />

10 1<br />

10 −2<br />

0<br />

−100<br />

−200<br />

−300<br />

10 1<br />

−400<br />

Bodeplot <strong>of</strong> full and reduced (2nd order) model valve<br />

10 2<br />

10 2<br />

Frequency (Hz)<br />

Fig. 3.8: Bode plot <strong>of</strong> 5 th -order linearised servo valve model and 2 nd -order reduced version.<br />

transmission lines. With [153], the lowest transmission line eigenfrequency, ! t can roughly<br />

be approximated by<br />

Lt<br />

!t =<br />

co 2<br />

10 3<br />

10 3<br />

(3.20)<br />

With line length, Lt =1:4m in case <strong>of</strong> the SRS and the wave propagation velocity, c o<br />

given by<br />

s r<br />

E 109 co = =<br />

1000� (3.21)<br />

850<br />

where E is the bulk modulus <strong>of</strong> the oil and the density, this results in an estimated eigenfrequency<br />

<strong>of</strong> 180Hz.<br />

In [124] it has been shown that these dynamics can not be neglected if one wants to apply<br />

high performance <strong>motion</strong> <strong>control</strong> on long stroke hydraulic actuators typically used in <strong>flight</strong><br />

simulation. The lightly damped resonances <strong>of</strong> the transmission lines together with the phase<br />

lag introduced by the limited bandwidth <strong>of</strong> the servo valve easily result in stability problems.<br />

By taking these phenomena into account in modelling, proper model <strong>based</strong> <strong>control</strong> can be<br />

applied.<br />

The main steps in modelling the transmission lines will now be described, following Van<br />

Schothorst [124] and Yang and Tobler [158]. Details can be found in these references. The<br />

main conclusion will be that transmission lines can be modelled as a (theoretically infinite)<br />

set <strong>of</strong> parallel connected linear second order <strong>system</strong>s. Each second order <strong>system</strong> appoints a


3.2 Extensions 87<br />

φ i<br />

P i<br />

Fig. 3.9: A causal two-port configuration for a transmission line.<br />

specific vibrational mode with increasing eigenfrequency and, in approximating, all modes<br />

<strong>of</strong> interest are easily picked.<br />

In choosing the input flow and output pressure as input and the input pressure and output<br />

flow as output, a proper causal solution will result as discussed by Van Schothorst in<br />

[124]. This two-port representation <strong>of</strong> the transmission line is given in Fig. 3.9. The exact<br />

causal solution to a one dimensional distributed parameter model <strong>of</strong> a uniform rigid fluid<br />

transmission line with laminar flow in the Laplace domain, is given by Yang and Tobler in<br />

[158] by<br />

Pi(s)<br />

o(s)<br />

=<br />

1<br />

cosh(;(s))<br />

1 ;Zc(s) sinh(;(s))<br />

; sinh(;(s))=Zc(s) 1<br />

φ o<br />

oP<br />

Po(s)<br />

i(s)<br />

(3.22)<br />

with the characteristic impedance, Zc(s), and wave propagation operator, ;(s), <strong>of</strong> a single<br />

transmission line, as in [158]. One can approximate this <strong>system</strong> by a finite dimensional state<br />

space model by using the first n terms <strong>of</strong> an exact infinite modal sequence, which can be<br />

derived from (3.22).<br />

This modal approximation technique leads to<br />

Pi(s)<br />

o(s)<br />

=<br />

=<br />

n<br />

k=1<br />

Pik(s)<br />

ok(s)<br />

n 1<br />

k=1<br />

s2 + cks + dk<br />

acks + bck ;azks + bzk<br />

ask ; bsk acks + bck<br />

Po(s)<br />

i(s)<br />

(3.23)<br />

Considering each second order <strong>system</strong>, an even more compact state space representation<br />

can be derived [158], requiring only four parameters for the first mode and two additional<br />

(due to parameter dependencies) for each higher harmonic to be taken into account.<br />

Each harmonic can then separately be described by<br />

_<br />

Pik<br />

_ ok<br />

=<br />

0 (;1) k+1 1k<br />

(;1) k 2k ; 3k<br />

Pbk<br />

ak<br />

+<br />

0 4k<br />

; 2k<br />

1k 4k 0<br />

Pok<br />

ik<br />

:<br />

(3.24)<br />

These parameters still have a clear physical interpretation, as given in [124]. A block scheme<br />

representing this model is given in Fig. 3.10 with for k =1, t j = j j =1�::: �4� t5 =<br />

; 4 2= 1. Only one mode is given in this block scheme, as the structure <strong>of</strong> the others is


88 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 3.10: Basic block scheme <strong>of</strong> the parallel modal transmission line model. Note that each<br />

additional mode, i, can be accounted for by an equivalent scheme connected in<br />

parallel through the sigma signs.<br />

equivalent. The sigma-terms show the connection <strong>of</strong> these additional harmonics, which can<br />

be taken into account by a parallel connection <strong>of</strong> similar block schemes.<br />

To maintain approximate steady state gain for n =1, which will be lost due to simple<br />

neglection <strong>of</strong> all higher order terms, 4 can be set to ; 1 if pressure and flow have been<br />

normalised. The term p 1 2 is equal to the undamped eigenfrequency <strong>of</strong> each mode and<br />

3 is the dissipating term, damping the resonance. In the hydraulic actuators <strong>of</strong> the SRS<br />

the lowest eigenfrequencies <strong>of</strong> the transmission lines could be found in the range <strong>of</strong> 160 ;<br />

220 Hz and the first higher harmonic at a factor three higher. Given the bandwidth <strong>of</strong> the<br />

valve <strong>of</strong> 150Hz, in this research each transmission line was modelled by taking only one<br />

mode into account<br />

Apart from the resonating behaviour, the transmission lines result in a difference between<br />

the pressure measured over the valve, dP i and the pressure dPo, over the actuator<br />

chambers, which actually drive the mechanical <strong>system</strong>. This difference can become a problem<br />

if one wants to apply a well defined force to accelerate with high precision. As will be<br />

shown, these differences occur at much lower frequencies than those where the transmission<br />

line resonances occur.<br />

In [124], it was argued that the state describing the mean pressure in the hydraulic actuator,<br />

Pm, should be retained if transmission lines are to be taken into account and the<br />

model is to be used over the whole stroke including the position dependent stiffness <strong>of</strong> the<br />

hydraulic actuator. But in applying model <strong>based</strong> <strong>control</strong>, it is unfavourable to include an<br />

almost un<strong>control</strong>lable and unobservable state. Therefore it is proposed to use an alternative<br />

model.<br />

There are two separate transmission lines as shown in Fig. 3.3. The model <strong>of</strong> transmission<br />

line 1 has inputs i1 and Po1 and outputs Pi1 and o1. The model <strong>of</strong> transmission line<br />

2 has inputs i2 and Po2 and outputs Pi2 and o2. The transmission line inputs Po1 and<br />

Po2 are constructed from the reduced hydraulic actuator model, (3.6), by<br />

Po1 = 1=2(Ps + dPo)<br />

Po2 = 1=2(Ps ; dPo)� (3.25)


3.3 Integrated <strong>system</strong> 89<br />

which is a similar assumption as given in (3.15). Further, the other connection <strong>of</strong> this two<br />

sided coupling between the transmission lines and the hydraulic actuator is given by<br />

m =( o1 + o2)=2 (3.26)<br />

At the other side <strong>of</strong> the transmission lines, the servo valve is connected. The models <strong>of</strong> the<br />

two can be merged by retaining (3.16) and using i1 = i2 = im and dPi = Pi1 ; Pi2.<br />

By evaluating the differences between the reduced and full order model after connecting<br />

the building blocks, this step can and will be validated. Integrating the set <strong>of</strong> submodels<br />

introduced for the mechanical and hydraulical part <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> will<br />

be the subject <strong>of</strong> the next section.<br />

In this section two relevant extensions, the valve characteristics and transmission line<br />

dynamics, to the basic hydraulically driven mechanical <strong>system</strong> model were introduced. Of<br />

course many more effects could be investigated. One, which should be mentioned here,<br />

but will not be discussed much further, are the hydraulic actuator cushioning areas at both<br />

ends <strong>of</strong> the stroke. By gradually disabling oil flow to and from the actuator, the cushioning<br />

prevents from high accelerations (max. 2g) occurring if an actuator for some reason runs out<br />

<strong>of</strong> stroke. This mechanism, however, also influences the <strong>system</strong> dynamics for at least the<br />

first and last 0:15m stroke <strong>of</strong> all the actuators. Design and evaluation <strong>of</strong> the SRS cushioning<br />

is discussed in [124].<br />

3.3 Integrated <strong>system</strong><br />

Having introduced models <strong>of</strong> the mechanical part <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> in<br />

Chapter 2 and <strong>of</strong> the hydraulic servo actuators which drive the <strong>system</strong> in Section 3.1 and<br />

Section 3.2, the two can be connected. This results in a nonlinear model describing the<br />

hydraulically driven mechanical <strong>system</strong>.<br />

An important <strong>system</strong> analytic tool, passivity analysis, will be used to derive some interesting<br />

properties <strong>of</strong> these non-linear models. E.g. it will be shown that the full multivariable<br />

nonlinear <strong>system</strong> model is passive from input oil flows to input valve pressure differences<br />

irrespective <strong>of</strong> parasitic effects like transmission line resonances and/ or flexible mechanical<br />

modes. As a result proportional (or either other passive) input pressure feedback to valve<br />

oil flow can be applied without destabilising the <strong>system</strong>, as is <strong>of</strong> course already long known<br />

heuristically.<br />

In the sequel, the <strong>system</strong> characteristics due to the hydraulic <strong>system</strong> are best clarified by<br />

first considering a one degree-<strong>of</strong>-freedom actuator driving a single mass. Also the effect <strong>of</strong><br />

a parasitic mass or a non rigidly connected foundation is best understood by using such a<br />

simplified model first. The complexity <strong>of</strong> the fully non-linear six degree-<strong>of</strong>-freedom Stewart<br />

platform model driven by its hydraulic actuators is considerable since it easily includes over<br />

fifty states. Most <strong>of</strong> the <strong>system</strong> properties seen in practise are, however, well understood<br />

by using these kinds <strong>of</strong> models because the physical structure <strong>of</strong> the whole <strong>system</strong> has been<br />

maintained.


90 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

3.3.1 Passive input/output pairs in <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s<br />

Passivity is a strong input/ output <strong>system</strong>s property, which can be derived for a large class<br />

<strong>of</strong> <strong>system</strong>s also inhibiting nonlinearities. It roughly tells that a <strong>system</strong> does not generate<br />

energy by itself. From this property, stability is <strong>of</strong>ten easily derived, also in combination<br />

with feedback.<br />

There exist a large amount <strong>of</strong> <strong>control</strong> strategies <strong>based</strong> on the passivity property e.g.<br />

found in [11, 66, 128, 147]. Mathematically it is defined here following Van der Schaft<br />

[147]. The inner product for the vector signals, f and g in the L n 2 -space, is taken as<br />

Z 1<br />

< f�g>=<br />

0<br />

n<br />

i=1 fi(t)gi(t)dt: (3.27)<br />

Definition 3.1 Let the causal input/ output <strong>system</strong> G map L n 2 ! L n 2 (thus stable). Then G<br />

is passive if there exists some constant such that<br />

� 8 u 2 L n 2 : (3.28)<br />

By introducing the extended L2e-space, the definition can be generalised to a larger signal<br />

set and noncausal/ nonstable <strong>system</strong>s. A causal or nonanticipating <strong>system</strong> means that<br />

G(u)T = G(uT ) 8 u 2 Ln 2e where the subscript T gives the truncation at time T . In the<br />

extended space, integration <strong>of</strong> (3.27) is required only up till time T.<br />

Starting with small sub<strong>system</strong>s from which passivity can be derived, complex <strong>system</strong><br />

passivity can be evaluated by using the fact that feedback <strong>of</strong> a passive <strong>system</strong> with a passive<br />

<strong>system</strong> results in a over all passive <strong>system</strong> and two parallel passive <strong>system</strong>s together form<br />

one passive <strong>system</strong>. This does not hold for passive <strong>system</strong>s in a series connection.<br />

Consider the feedback <strong>system</strong> <strong>of</strong> Fig. 3.11. With the assumption that (e 1� e2) 2 Ln 2 !<br />

(u1� u2) 2 Ln 2 and passive G1 and G2 i.e. < yi� ui > i for i = 1� 2, the feedback<br />

<strong>system</strong> with inputs (e1� e2) and outputs (y1� y2) is also passive since<br />

< y1� e1 > + < y2� e2 > = < y1� e1 +y2 > + < y2� e2 ; y1 >=<br />

< y1� u1 > + < y2� u2 > 1 + 2 8u1�2 2 L n 2 (3.29)<br />

Notice that since this holds for a multivariable <strong>system</strong>, any input/ output combination, for<br />

which a <strong>system</strong> is passive, can be fed back by a passive <strong>system</strong> without influencing the<br />

passivity <strong>of</strong> some passive input/ output combination somewhere else. Passivity <strong>of</strong> a parallel<br />

connection <strong>of</strong> some passive G1 and G2 with output y =y1 +y2 and u =u1 =u2 is proven<br />

by<br />

< y� u>= < y1� u1 > + < y2� u2 > 1 + 2 8 u 2 L n 2 (3.30)<br />

Passivity can be proven easily for some basic <strong>system</strong> structures. For the time varying<br />

positive gain ’<strong>system</strong>’ described by y(t) =k(t)u(t) with k(t) 0, it can be proven by<br />

Z 1 Z 1<br />

y(t)u(t)dt = u(t)k(t)u(t)dt 0 8 u 2 L2 (3.31)<br />

0<br />

0


3.3 Integrated <strong>system</strong> 91<br />

e1<br />

y2<br />

+<br />

- d<br />

- 6<br />

u1<br />

-<br />

G1<br />

G2<br />

u2<br />

Fig. 3.11: A general feedback structure<br />

For the stable linear first order <strong>system</strong> described by _y =(u;ky) with k 0,oru = _y +ky,<br />

it follows from<br />

Z 1 Z 1<br />

y(t)u(t)dt = y(t)ky(t) +<br />

0<br />

0<br />

1<br />

Z 1<br />

2 0<br />

d<br />

dt y2 (t)dt > ; 1<br />

2 y2 (0) 8 u 2 L2 (3.32)<br />

For linear <strong>system</strong>s passivity means that the real part <strong>of</strong> the transfer function in the frequency<br />

domain will be positive. For passive linear single input single output <strong>system</strong>s the phase will<br />

thus vary between =2.<br />

The collocated input torque - output velocity pairs <strong>of</strong> a mechanical <strong>system</strong>, which can be<br />

described in hamiltonian form, are passive for every initial condition. From the definition <strong>of</strong><br />

the Hamiltonian (2.84) and assuming that the potential energy function is chosen as P(z)<br />

0 follows H 0. With the mechanical <strong>system</strong> described by the generalised coordinates,<br />

z, consider an output mapping y = J(z) _z, and the input, u, defined by the dual map,<br />

= J T (z)u. Now the <strong>system</strong> from input u to output y will shown to be passive. Consider<br />

(2.87), then<br />

< y� u>T =< _z� >T =<br />

0<br />

d<br />

H(m(t)� z(t))dt =<br />

dt<br />

H(m(T )� z(T )) ; H(m(0)� z(0)) > ;H(m(0)� z(0)) 8 u 2 L2e (3.33)<br />

Take y = _q, u = fa, z = sx and J(z) = Jl�sx(sx) and this establishes passivity <strong>of</strong> the<br />

rigid body Stewart platform (2.122) for the <strong>system</strong> considered from input actuator force,<br />

fa to output actuator velocity, _q. Moreover, it tells that passivity will be preserved in case<br />

<strong>of</strong> connecting a large class <strong>of</strong> parasitic mass- spring- damper <strong>system</strong>s connected e.g. a non<br />

rigidly attached foundation.<br />

Having established passivity <strong>of</strong> the mechanical <strong>system</strong>, the <strong>system</strong> can be extended to<br />

the hydraulics retaining the passivity property.<br />

-Viscous friction/ damping Every Rayleigh function, R satisfying _z T (@R)=(@ _z)( _z) 0<br />

is a passive <strong>system</strong>. Rayleigh functions <strong>of</strong>ten occur as dissipative feedback over passive<br />

mechanical (flow) <strong>system</strong>s retaining passivity. A simple example <strong>of</strong> a Rayleigh<br />

function is the viscous friction, bi _q i in each <strong>of</strong> the actuators:<br />

Rvact( _x� sx) =J T l�x (sx)diag(bi _qi) =J T l�x (sx)diag(bi)Jl�x(sx) _x: (3.34)<br />

Z T<br />

+<br />

d?<br />

+<br />

y1<br />

e2<br />

-


92 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Amplitude<br />

Phase (deg)<br />

10 2<br />

10 1<br />

10 0<br />

10 −1<br />

10 0<br />

10 −2<br />

100<br />

0<br />

−100<br />

dPo−dPi<br />

−200<br />

10 0<br />

−300<br />

10 1<br />

10 1<br />

Bodeplot Phi_im−>dP_i,o<br />

dPo−dPi<br />

Frequency (Hz)<br />

Fig. 3.12: Bode plot <strong>of</strong> 7 th -order linearised hydraulically driven mechanical <strong>system</strong> model<br />

including transmission lines. Input valve flow, im, outputs, dPi (full), dPo<br />

(dashed), dPo ; dP i (dotted).<br />

Equivalently damping induced through resonances can also be included without destroying<br />

passivity.<br />

-The hydraulic actuator The model <strong>of</strong> the hydraulic part <strong>of</strong> an actuator, given in (3.6)<br />

can be considered as a stable first order <strong>system</strong> with a varying gain C m(q). Since<br />

Cm(q) > 0 each linearisation <strong>of</strong> the hydraulics described by (3.6) results in a passive<br />

<strong>system</strong> for every position q. Further, as shown as a part in the block scheme <strong>of</strong><br />

Fig. 3.2, it is connected to the mechanical <strong>system</strong> as a negative feedback over a passive<br />

<strong>system</strong>. So each linearised hydraulic actuator connected to the nonlinear mechanical<br />

<strong>system</strong> preserves passivity. The passive input/ output <strong>system</strong> <strong>of</strong> the actuator can be<br />

chosen from m to dPo respectively. With the varying gain, the situation is somewhat<br />

different:<br />

< m�dPo > =<br />

+<br />

Z 1<br />

0 Z 1<br />

0<br />

d<br />

dt dPo(t)<br />

10 2<br />

10 2<br />

dPi<br />

dPo<br />

dPi<br />

dPo<br />

dPo(t) Llm ; d<br />

dt<br />

1<br />

4Cm(q(t)<br />

1<br />

4Cm(q(t))<br />

10 3<br />

10 3<br />

dPo(t) dt<br />

dPo(t)dt (3.35)<br />

Due to the time derivative <strong>of</strong> the inverse actuator stiffness, this <strong>system</strong> is not always<br />

passive. With an upper bound on the actuator velocity <strong>of</strong> e.g. 1 m=s and with a<br />

polynomial fit Cm(q) = b2q 2 + b1q + b0 =99:2q 2 ; 9:8q + 104 [124], a leakage<br />

term <strong>of</strong> 0.002 (0.01 is normal) will passify this <strong>system</strong>.<br />

-The transmission lines The structure <strong>of</strong> the transmission lines models in the block scheme<br />

given in Fig. 3.10 is clearly a parallel connection <strong>of</strong> passive <strong>system</strong>s which can be considered<br />

as two stable linear first order <strong>system</strong>s in feedback. The actuator side <strong>of</strong> the


3.3 Integrated <strong>system</strong> 93<br />

10 2<br />

10 0<br />

10 −2<br />

10 2<br />

10 0<br />

10 −2<br />

10 2<br />

10 0<br />

10 0<br />

10 −2<br />

10 1<br />

10 2<br />

Frequency (Hz)<br />

10 3<br />

100<br />

0<br />

−100<br />

100<br />

0<br />

−100<br />

100<br />

0<br />

10 0<br />

−100<br />

10 1<br />

10 2<br />

Frequency (Hz)<br />

Fig. 3.13: Bode plots from input valve flow ( im) to pressure difference (dPi) <strong>of</strong> linearised<br />

hydraulically driven mechanical <strong>system</strong> models including transmission lines in<br />

upper/ mid/ lower position (q = :47� 0� ;:47m) for full (full), reduced (n =6,<br />

dashed) order models and difference (dotted).<br />

port feeds back over the passive input/ output pair dP o and m respectively and at<br />

the valve side the input/ output pair is given by im and dPi, considering the parallel<br />

transmission lines over the pairs Pi1 and i1 and Pi2 and i2.<br />

-The valve port dynamics The valve port dynamics modelled as turbulent flow port can<br />

be considered as a positive time varying gain. If j dP i j 1,<br />

p<br />

1 ; sgn(xm(t))dPi(t) =k(t)xm� (3.36)<br />

im(t) =xm(t)<br />

implies k(t) 0. This is a passive <strong>system</strong> which preserves passivity if used in a feedback<br />

connection with e.g. constant gain from dP i to xm. As shown in Section 3.2.1,<br />

the valve spool positioning <strong>system</strong> cannot be considered a constant gain over an infinitely<br />

large frequency area but inhibits dynamics which is clearly not passive from<br />

input voltage (u) to output spool position (x m). At this point passivity <strong>based</strong> <strong>control</strong><br />

strategies may cause possible (stability) problems.<br />

Summarising, it has been established that the hydraulically driven nonlinear mechanical<br />

<strong>system</strong> is passive with respect to a large number <strong>of</strong> input/ output pairs. Under moderate<br />

conditions, which are typically satisfied in practice, constant feedback <strong>of</strong> the pair (valve<br />

spool position, pressure difference measured at the valve) leaves the passivity properties <strong>of</strong><br />

the nonlinear <strong>system</strong> unaltered. In fact, this kind <strong>of</strong> feedback will dissipate energy, which<br />

drives the <strong>system</strong> with badly damped resonances faster to its stable equilibrium.<br />

10 3


94 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Variable Description Value Unit<br />

n Max. valve flow 25e ;4 [m3 =s]<br />

Ps Supply Pressure 160e5 [N=m2 ]<br />

Fmax = Ps Ap Max. actuator force 40e3 [N ]<br />

_qmax = n=Ap Max. velocity 1 [m=s]<br />

Normalised Operators Description Value Unit<br />

An = Ap _qmax= n Piston area 1 [;]<br />

C1 = E n=(V1Ps) Stiffness chamber 1 101 [1=s]<br />

C2 = E n=(V2Ps)<br />

L1 = L1tPs= n<br />

Stiffness chamber 2<br />

Leakage 1<br />

104<br />

4:32e<br />

[1=s]<br />

;2<br />

L2 =(L2s + L2t)Ps= n Leakage 2 5:06e<br />

[;]<br />

;2<br />

Wpn = wp=Fmax Visc. friction 3:5e<br />

[;]<br />

;4<br />

M<br />

[;]<br />

;1<br />

n = Fmax=(Mp _qmax) Inv. Mass 10 [1/s]<br />

Table 3.2: Parameters taken for a 1-d.o.f. hydraulic mechanical <strong>system</strong>. First the four normalising<br />

quantities are given. Then all the normalised values for <strong>of</strong> the operators<br />

are specified.<br />

Variable Value Variable Value<br />

t11 = 1737 t12 = 1254<br />

t21 = 650 t22 = 1184<br />

t31 = 30 t32 = 54<br />

Table 3.3: Parameters taken for the two transmission lines.<br />

3.3.2 The 1-d.o.f. hydraulically driven mechanical <strong>system</strong><br />

Further insight into the characteristics <strong>of</strong> the hydraulic mechanical models, derived in the<br />

previous sections, can be obtained by filling in realistic values into the parameters, e.g. SRS<br />

design numbers, and performing an evaluation in the frequency (linearised models) and<br />

time domain. First a one-degree-<strong>of</strong>-freedom hydraulically driven mechanical <strong>system</strong> with<br />

transmission line dynamics will be considered. For these models, the difference in the<br />

dynamics <strong>of</strong> the actual driving pressure in the actuator and the measured pressure over the<br />

valve will be evaluated. Further, the implications <strong>of</strong> the reduction step proposed by taking<br />

(3.6) instead <strong>of</strong> (3.2), assuming (3.25) and (3.26), can be quantified.<br />

It is important to normalise the variables such as pressures (typically 10 7 in SI-units)<br />

and flows (typical values <strong>of</strong> 10 ;3 in SI) when building numerical models which contain<br />

hydraulics. In Table 3.2, first the normalising values <strong>of</strong> these variables are given. Then the<br />

normalised values <strong>of</strong> the operators taken for the models consisting <strong>of</strong> (3.13) together with<br />

(3.6) or (3.2) are specified at neutral position (at full stroke q =0:625). With the mass taken<br />

<strong>of</strong> 4000kg and the typical values <strong>of</strong> the hydraulic actuators <strong>of</strong> the SRS, an undamped rigid


3.3 Integrated <strong>system</strong> 95<br />

body mode eigenfrequency, f rm <strong>of</strong> approx. 7Hz results.<br />

frm = 1<br />

2<br />

r<br />

2Cm<br />

= 1<br />

2<br />

r<br />

2 4:1 106 4000<br />

= 1 p<br />

(101 + 104)10 = 7 Hz<br />

2<br />

(3.37)<br />

Mn<br />

This corresponds with the lowest eigenfrequencies found for the SRS. Due to the normalising<br />

velocity <strong>of</strong> one, the normalised inverse mass can also be interpreted as the maximum<br />

acceleration to be attained, which in this case approx. amounts to gravity. The leakage term<br />

related to the damping in the actuator driving direction amounts to L lm =2:35%, which is<br />

equal to the value experimentally found by Van Schothorst [124]. This is also the case for<br />

the viscous friction term.<br />

The model is completed by connecting the transmission lines using (3.24) and assuming<br />

an ideal valve input, i1 = i2 = im. The parameters taken for the transmission lines<br />

correspond to the values derived in [125] adapted with 10 % additional line length to arrive<br />

at the eigenfrequencies found in practice. The three resulting parameters, t 1k:::3k, for the<br />

two lines (k = 1� 2) are given in Table 3.3. Eigenfrequencies <strong>of</strong> the unconnected transmission<br />

lines therefore amount to 169 Hz and 194 Hz with damping <strong>of</strong> t1 = 0:01 and<br />

t2 =0:02 respectively.<br />

Influence <strong>of</strong> the transmission lines on input and output pressure dynamics<br />

Bode plots <strong>of</strong> this <strong>system</strong> from input m to the outputs dPi and dPo are given in Fig. 3.12.<br />

Connection <strong>of</strong> the impedance <strong>of</strong> the hydraulic actuator to the transmission lines shifted<br />

the transmission line eigenfrequencies approx. 10 Hz higher. This shift <strong>of</strong> approx. 5%<br />

is also observed for the lower (rigid body) mode and in that case to lower values for the<br />

eigenfrequency resulting from the load together with the oil spring stiffness. This effect<br />

<strong>of</strong> interaction between the transmission lines and the actuator should not be neglected. It<br />

means e.g. that the effective oil stiffness with the SRS-actuators will be 12% lower than<br />

the theoretically found value in case one identifies this hydraulically driven <strong>system</strong> with a<br />

model structure not accounting for the transmission lines.<br />

The phase <strong>of</strong> the passive input/ output pair ( im�dPi) clearly remains within 90 .It<br />

is important to notice that although the pressures have a one-to-one correspondence in the<br />

lower frequency area, they start to deviate much earlier than the transmission line eigenfrequencies.<br />

This is due to the anti-resonance in the transfer function from im to dPi at<br />

approx. 60 Hz.At20 Hz this amounts to 16% and at 30 Hz already accounts for a difference<br />

<strong>of</strong> 25%. Further, the resonating behaviour <strong>of</strong> the transmission lines is expected to be<br />

much (10 times) heavier observed in the measurable output, dP i, than it will be felt by the<br />

mechanical <strong>system</strong> (through dPo).<br />

Influence <strong>of</strong> the varying actuator stifness and model reduction step<br />

Since the stiffness <strong>of</strong> the oil column depends on the actuator position, the eigenvalues, also<br />

those related to the transmission lines, will shift if the actuator moves. As argued in [124],<br />

this shift will be slightly different if the actuator model is reduced to (3.6). However, the<br />

reduced model describes the <strong>system</strong> reasonably well as can be observed in Fig. 3.13. Bode<br />

plots <strong>of</strong> the <strong>system</strong> from input, im, to the measurable output dPi in an upper/ mid and<br />

lower position are taken for the reduced and nonreduced models. Also the differences are


96 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Amplitude<br />

Phase (deg)<br />

10 2<br />

10 1<br />

10 0<br />

10 −1<br />

10 0<br />

10 −2<br />

0<br />

−200<br />

−400<br />

10 0<br />

10 1<br />

10 1<br />

Bodeplot u−>dPi,dPo<br />

Frequency (Hz)<br />

Fig. 3.14: Bode plot <strong>of</strong> 8 th -order linearised hydraulically driven mechanical <strong>system</strong> model<br />

including transmission line and valve dynamics. Input valve input voltage, u,<br />

outputs dP i (full), dPo (dashed).<br />

shown. The upper and lower positions are just within the operational area where the cushioning<br />

does not influence the actual <strong>system</strong>. Since the eigenvalues are relatively lightly<br />

damped, very small differences in the eigenvalue shift locally result in larger deviations but<br />

the general behaviour is described fairly well by the reduced order model. Both for the<br />

upper as the lower position the anti-resonance between the hydromechanical mode at approx.<br />

7 Hz and the first transmission line related mode moves up from 60 Hz to maximally<br />

90 Hz.<br />

Influence <strong>of</strong> the valve dynamics<br />

The model can be extended by adding the reduced linear second order model <strong>of</strong> the three<br />

stage valve dynamics derived from (3.19) and (3.17). In Fig. 3.14 the Bode plot <strong>of</strong> this<br />

model is given. The main effect <strong>of</strong> the valve is an additional lag in phase, which becomes<br />

prominent for frequencies > 100 Hz. Together with the high gains introduced with the<br />

transmission lines resonances, the valve phase lag forms an important stability risc in feedback<br />

design, which has to be taken into account.<br />

In <strong>system</strong>s with moderate requirements, the valve bandwidth can be decreased in order<br />

to filter most <strong>of</strong> the transmission line resonances. But also in that case, the phase lag <strong>of</strong><br />

the valve will be dangerous in some frequency area e.g. in combination with resonances<br />

due to structural deformation. In the lower frequency region, the nonlinearities <strong>of</strong> the valve,<br />

e.g. due to (3.16), will form the most important factor.<br />

In the time domain, the step response <strong>of</strong> this model is given in Fig. 3.15 on a 200 ms<br />

range. The steady state response to the input and output pressures is neglectable. This<br />

means it is not very feasible to induce sustained accelerations with hydraulically driven<br />

10 2<br />

10 2<br />

10 3<br />

10 3


3.3 Integrated <strong>system</strong> 97<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

−3<br />

Step response u−>dPi,dPo,dPo−dPi<br />

−4<br />

0 0.02 0.04 0.06 0.08 0.1<br />

time (s)<br />

0.12 0.14 0.16 0.18 0.2<br />

Fig. 3.15: Step response <strong>of</strong> 8 th -order linearised hydraulically driven mechanical <strong>system</strong><br />

model including transmission line and valve dynamics. Input valve input voltage,<br />

u, outputs dPi (full), dPo (dashed). dPo ; dPi (dash-dot). The high frequency<br />

resonances result from the transmission lines, the slower vibration with<br />

approx. 0:16s period time is the rigid body mode.<br />

mechanical <strong>system</strong>s. A quasi static gain <strong>of</strong> a moving actuator is attained from valve input<br />

voltage to actuator velocity. Of course in practice, the limited stroke is the main limiting<br />

factor for the low frequency accelerations. In <strong>control</strong> design for acceleration (simulation),<br />

however, the virtually zero gain from valve input to force/ acceleration will be an important<br />

factor to be taken into account.<br />

In both pressures the 160 ms time period (7Hz) <strong>of</strong> the hydromechanical mode is most<br />

prominent. The difference in response is mainly due to the transmission line resonances to<br />

which the pressure measured over the valve is much more sensitive. Due to the interference<br />

<strong>of</strong> the two transmission line modes, the high vibrational amplitude comes up and falls within<br />

a 40 ms time range.<br />

3.3.3 Additional mechanical modes<br />

The models presented in the previous paragraphs include the dynamics <strong>of</strong> the hydraulic actuator(s)<br />

into the higher frequency areas <strong>of</strong> approx. 200 Hz. In deriving accurate models <strong>of</strong> a<br />

hydraulically driven mechanical <strong>system</strong>s over this broad frequency area, one can usually not<br />

assume the mechanics to be entirely rigid. In this section, two <strong>of</strong> such cases are examined<br />

for a 1-d.o.f. <strong>system</strong>. First, the possible nonrigidness <strong>of</strong> a foundation is taken into account.<br />

In the second case, the driven <strong>system</strong> consists <strong>of</strong> a main mass connected to a parasitic mass/<br />

spring/ damper.<br />

A schematic drawing <strong>of</strong> hydraulically driven mass, m s, connected to a nonrigid foundation<br />

is given in Fig. 3.16. The foundation with mass, m f is fastened to the ground with a


98 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 3.16: Schematic drawing <strong>of</strong> a hydraulic actuator moving a consisting <strong>of</strong> a single mass<br />

ms in 1-d.o.f. pushing from a non rigidly attached foundation, acting as a mass/<br />

spring/ damper <strong>system</strong> mf �bf �cf .<br />

spring <strong>of</strong> stiffness cf and a damper, bf . The equations <strong>of</strong> <strong>motion</strong> become<br />

msxs = AdPo (3.38)<br />

mf xf = ;AdPo ; bf _xf ; cf xf<br />

Although the actual force fa = ApdPo necessary to accelerate the mass, ms does not<br />

change with the nonrigid foundation, the feedback path relating _q = _x s ; _xf changes. In<br />

case <strong>of</strong> the structure with simple mass given in Fig. 3.2, the feedback path becomes _q =<br />

ApdPo=(mss), assuming no viscous friction. Every pole <strong>of</strong> the feedback path becomes a<br />

transmission zero in the transfer function from input (valve flow) to output pressure, dP o.<br />

With the foundation the feedback path changes to<br />

considering the additional equation <strong>of</strong> <strong>motion</strong>,<br />

_q = (ms + mf )s 2 + bf s + cf<br />

mss(mf s 2 + bf s + cf ) ApdPo� (3.39)<br />

mf xf + bf _xf + cf xf + fa =0: (3.40)<br />

p<br />

cf =mf ,<br />

A complex pair <strong>of</strong> transmission zeros with the undamped eigenfrequency, ! f =<br />

<strong>of</strong> the foundation alone results. The structure <strong>of</strong> the poles <strong>of</strong> this interconnected <strong>system</strong> is<br />

somewhat more complex. The foundation will add two complex poles and the transmission<br />

zeros will always lie in between, since the passivity properties will be preserved. Usually the<br />

mass <strong>of</strong> the foundation will be much higher than the mass to be moved. In this case, although<br />

the respective eigenfrequencies can be close, there will not be a large shift in connecting the<br />

models <strong>of</strong> the hydraulically driven mass, ms and the foundation. In Fig. 3.17 the Bode plots<br />

<strong>of</strong> this model structure with three different parameter values are given. In all cases (1f,2f,3f)<br />

the model <strong>of</strong> the hydraulically driven mass is as specified in the previous paragraphs. The<br />

parameters concerning the foundation are given in Table 3.4. f is the relative damping<br />

given by 2 f cf = !fbf .<br />

In both cases 1f and 2f, in which the mass <strong>of</strong> the foundation is relatively large (the<br />

realistic cases), the nonrigid foundation results in a relatively small distortion <strong>of</strong> the open


3.3 Integrated <strong>system</strong> 99<br />

Amplitude<br />

Phase (deg)<br />

10 2<br />

10 1<br />

10 0<br />

10 −1<br />

10 0<br />

10 −2<br />

100<br />

0<br />

−100<br />

−200<br />

10 0<br />

−300<br />

Case mf =ms !f f<br />

1f 10 2 10 0:05<br />

2f 10 2 4 0:05<br />

3f 1 2 10 0:05<br />

Table 3.4: Parameters taken for the foundation.<br />

10 1<br />

10 1<br />

Bodeplot Phi_im−>dP_i<br />

Frequency (Hz)<br />

Fig. 3.17: Bode plots <strong>of</strong> input valve flow, im, to pressure difference over the valve, dP i,<br />

with 1-d.o.f. hydraulically driven mass connected to a nonrigid foundation. Case<br />

1f (full), 2f (dashed), 3f (dash-dotted).<br />

loop characteristics <strong>of</strong> the original <strong>system</strong>. The additional resonance and anti-resonance are<br />

clearly visible and also the, mostly oil spring stiffness-moveable mass related resonance is<br />

seen to be shifted slightly. In the third case, where the mass <strong>of</strong> the foundation and moveable<br />

mass are equal, the interaction between the modes is shown to be much higher. Although<br />

this example is highly hypothetical in case <strong>of</strong> a foundation, these typical effects can occur<br />

if one would place two platforms on top <strong>of</strong> each other like McInroy et al. [93]. In all<br />

cases the transfer functions from im to dPi clearly remain strict positive real i.e. passive.<br />

For the high and low frequencies, the response is not sensitive to the characteristics <strong>of</strong> the<br />

foundation.<br />

The second example <strong>of</strong> a possible parasitic resonating behaviour is schematically drawn<br />

in Fig. 3.18. The construction to be moved itself, e.g. the <strong>simulator</strong>, is nonrigid. In practice<br />

such a situation can occur if one connects a large projection screen in front <strong>of</strong> and heavy<br />

projectors on top <strong>of</strong> the cockpit <strong>of</strong> a <strong>simulator</strong>. In this 1-d.o.f. case the parasitic behaviour<br />

is modelled as two masses connected through a spring/ damper. Although this schematic<br />

example will not catch the nonlinear behaviour <strong>of</strong> a flexible distributed mass with gross<br />

spatial movement, some typical, linear parts <strong>of</strong> the characteristics can be investigated.<br />

10 2<br />

10 2<br />

10 3<br />

10 3


100 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Fig. 3.18: Schematic drawing <strong>of</strong> a hydraulic actuator moving a mass m s in 1-d.o.f., which<br />

is connected to a parasitic mass/ spring/ damper <strong>system</strong> mp�bp�cp.<br />

Case mp=ms !p p<br />

1p 1 2 10= p 2 0:05= p 2<br />

2p 1 2 10 p 2 0:05= p 2<br />

3p 3 2 10 p 2 0:05= p 2<br />

Table 3.5: Parameters taken for the parasitic mass.<br />

The equations <strong>of</strong> <strong>motion</strong> for this <strong>system</strong> become<br />

msxs = cp(xp ; xs) +bp(_xp ; _xs) +ApdPo (3.41)<br />

mpxf = cp(xs ; xp) +bp(_xs ; _xp)<br />

In this case the velocity measured at the actuator is equal to the inertial velocity <strong>of</strong> the<br />

mass, ms. The transfer function from applied actuator force, f a = ApdPo, to the velocity,<br />

_q = _xs changes to<br />

_q =<br />

mps 2 + bps + cp<br />

mss(mps 2 + rmbps + rmcp) ApdPo� (3.42)<br />

with the mass ratio, rm = (ms + mp)=ms, since the actuator force is not the only force<br />

on ms as also the spring and damper force are taken into account. Again, the zero at zero<br />

frequency is retained and an additional complex pair is introduced. This time predicted at<br />

p cprm=ms. In principle, the structure <strong>of</strong> (3.39) and (3.42) is equal. As a result, one can not<br />

discriminate between the case <strong>of</strong> a nonrigid foundation or a nonrigid structure to be moved,<br />

given the measurement from valve input to a pressure difference (dP i or dPo) or actuator<br />

velocity, _q. To determine what causes the parasitic resonance, additional measurements<br />

have to be taken e.g. from accelerometers attached to the foundation or <strong>simulator</strong>. This<br />

identification problem will be considered in the next chapter.<br />

In Table 3.5 the parameters <strong>of</strong> three cases are given for the flexible structure <strong>of</strong> Fig. 3.18.<br />

All the parameters <strong>of</strong> the hydraulic actuator are also taken equal to the ones presented in<br />

the previous section (Table 3.2). The total moveable mass (m s + mp) is taken equal to


3.4 Hydraulically driven Stewart platform 101<br />

Amplitude<br />

Phase (deg)<br />

10 2<br />

10 1<br />

10 0<br />

10 −1<br />

10 0<br />

10 −2<br />

100<br />

0<br />

−100<br />

−200<br />

10 0<br />

−300<br />

10 1<br />

10 1<br />

Bodeplot Phi_im−>dP_i<br />

Frequency (Hz)<br />

Fig. 3.19: Bode plots <strong>of</strong> input valve flow, i, to pressure difference over the valve, dP i, with<br />

1-d.o.f. hydraulically driven structure consisting <strong>of</strong> two masses interconnected by<br />

spring and damper. Case 1p (full), 2p (dashed), 3p (dash-dotted).<br />

4000 kg in all cases. The first case (1p) results in exactly the same model parameters<br />

as case 3f, considering (3.39) and (3.42). The transfer function from i to dPi for these<br />

parameters, plotted in Fig. 3.19, is almost the same (the viscous friction <strong>of</strong> the actuator<br />

results in very small differences). The other cases shown have somewhat more realistic<br />

behaviour considering a resonant frequency twice as high (2f) and further taking 25% <strong>of</strong><br />

the part <strong>of</strong> the moving object nonrigidly connected (3f). In reducing this part, the zeros<br />

and poles mostly related to the parasitic resonance shift back to lower values towards the<br />

original eigenfrequency <strong>of</strong> the hydraulically driven load since there is less interaction.<br />

The effect <strong>of</strong> interaction will be similar to the parasitic resonating spring/ damper/<br />

masses treated in this section if one measures the flow to pressure/ velocity transfer function<br />

<strong>of</strong> one actuator in case <strong>of</strong> a hydraulically driven Stewart platform where more than one<br />

actuator is connected to the <strong>system</strong>. This structure will be treated in the next section.<br />

The parasitic resonance shows a moderate peak in the frequency domain in the more<br />

or less realistic setting <strong>of</strong> case 3p. In open loop the characteristics <strong>of</strong> the standard hydromechanical<br />

mode will dominate. With feedback this will change and will change more<br />

drastically if performance specifications require a bandwidth close to these parasitic effects.<br />

This problem will be more closely looked at in Chapter 5 where the closed loop <strong>system</strong> is<br />

validated.<br />

3.4 Hydraulically driven Stewart platform<br />

Having considered the hydraulic actuator and its most relevant properties to be taken into<br />

account in modelling in a 1-d.o.f. setting and considering the description <strong>of</strong> the 6-d.o.f. me-<br />

10 2<br />

10 2<br />

10 3<br />

10 3


102 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

dP o<br />

R<br />

b J T fa - A - - - b<br />

l�x<br />

Mt<br />

?<br />

-L 6<br />

C<br />

-B<br />

6<br />

R?<br />

?<br />

6<br />

-A<br />

_q _x<br />

- b - b<br />

Jl�x -<br />

om<br />

Fig. 3.20: Basic structure hydraulically driven <strong>motion</strong> <strong>system</strong>.<br />

chanical <strong>system</strong> model <strong>of</strong> the Stewart platform introduced in Chapter 2, the two can be<br />

combined. As interaction will be the most prominent additional effect to be taken into account,<br />

the structure <strong>of</strong> this phenomenon will be analyzed first, using a simplified model. The<br />

inertial effect <strong>of</strong> the actuators will be treated in the hydraulically driven setting. Finally, the<br />

full model will be considered performing analysis in the frequency domain (through local<br />

linearisation) for several platform poses using numerical design values <strong>of</strong> the SRS for the<br />

model parameters.<br />

3.4.1 Basic model structure hydraulically driven <strong>system</strong>s<br />

As a resume, the basic structure <strong>of</strong> a hydraulically driven parallel <strong>motion</strong> <strong>system</strong> is given<br />

in Fig. 3.20. Through the servo valves and transmission lines, oil flows to the actuator<br />

compartments, om, can be regulated. Strong coupling with the mechanics results through<br />

actuator velocities, _q, requiring a in/decreasing amount <strong>of</strong> oil in the compartments. Together<br />

with a small leakage term, ;L, which provides for some damping, the nett flow will result<br />

in in/decreasing pressure rises, dP<br />

_<br />

o, through the oil stiffness, C. Actuator pressure times<br />

piston area, A, will force the <strong>system</strong> to accelerate. With parallel <strong>system</strong>s the actuator forces,<br />

fa, have to be transformed to platform forces though the transposed jacobian, J T l�x . Some<br />

additional damping force results from the viscous friction along the actuators, B. The acceleration<br />

<strong>of</strong> the actuators can be calculated as the nett sum <strong>of</strong> forces times the inverse mass<br />

matrix, M ;1<br />

t , considered along the platform coordinates. Coulomb, coriolis and centripetal<br />

forces are not shown in this <strong>system</strong> and neither are the gravity terms.<br />

The mass matrix as seen from the actuators consists <strong>of</strong> the <strong>simulator</strong> pose, sx, dependent<br />

jacobian, Jl�x(sx) as defined by (2.42), and an almost constant <strong>simulator</strong> mass matrix, M t,<br />

6<br />

M ;1<br />

;1<br />

act = Jl�xM t J T l�x : (3.43)<br />

This matrix determines the coupling between the actuators. Given similar hydraulic actuators<br />

in equal position, all other matrix operators can be considered as one operator times<br />

the identity matrix. So, the eigenvectors <strong>of</strong> the linearised version <strong>of</strong> this <strong>system</strong> can be calculated<br />

from the mass matrix. This still holds if decentralized feedback (diagonal feedback<br />

structure) is applied. As the mass matrix is symmetric the eigenvector matrix can be made<br />

?<br />

-1


3.4 Hydraulically driven Stewart platform 103<br />

unitary. This results from the singular value decomposition<br />

M ;1 T<br />

act = U U (3.44)<br />

With the basic model for the hydraulically driven Stewart platform, it will be shown that the<br />

matrix, U, can be used in a state transformation matrix, which decouples the 12 th order <strong>system</strong><br />

into 6 second order <strong>system</strong>s each having the same structure as the 1-d.o.f. hydraulically<br />

driven mechanical <strong>system</strong>.<br />

Considering for a moment dP o and _q as state1 , the state space equations for Fig. 3.20<br />

become<br />

dP<br />

_<br />

o<br />

q<br />

=<br />

;CL<br />

M<br />

;CA<br />

;1<br />

actA ;M ;1<br />

actB<br />

The state transition matrix can be decomposed as<br />

dP o<br />

_q<br />

+ C<br />

0<br />

om<br />

(3.45)<br />

;CL<br />

M<br />

;CA<br />

;1<br />

actA ;1<br />

;M actB =<br />

U<br />

0<br />

0<br />

U<br />

;CL<br />

A<br />

;CA<br />

; B<br />

U T<br />

0<br />

0<br />

U T : (3.46)<br />

This can be used to decouple the <strong>system</strong> using the assumption that all matrices involved,<br />

apart from the mass matrix, have a scalar times identity structure, I. From this observation,<br />

a decoupled <strong>system</strong> results if the <strong>system</strong> is described w.r.t. a new set <strong>of</strong> inputs ^ om<br />

U<br />

=<br />

T ^<br />

om and outputs ^ dP o = U T dP o, _ ^q = U T _q.<br />

If lower captions are used for the diagonal elements <strong>of</strong> the respective operators, e.g. b<br />

from B = bI, the six decoupled second order <strong>system</strong>s can be described by the state space<br />

equations<br />

"<br />

_^<br />

dP o�i<br />

^q i<br />

#<br />

= ;cl ;ca<br />

ia ; ib<br />

dP ^<br />

o�i<br />

_^q i<br />

+<br />

c<br />

0<br />

^<br />

om�i� (3.47)<br />

with i =1�::: �6. So the <strong>system</strong> can be described with six second order <strong>system</strong>s in parallel.<br />

Moreover, exciting the <strong>system</strong> with an oil flow, om, directed along one <strong>of</strong> the columns, ui,<br />

<strong>of</strong> U will only result in response <strong>of</strong> the pressure, dP o, and velocity, _q along this very same<br />

direction, ui. Therefore the following definition makes sense.<br />

Definition 3.2 Each column <strong>of</strong> the matrix, U, is defined as the rigid body modal direction<br />

<strong>of</strong> the hydraulically driven Stewart platform. The mode, eigenfrequency/ damping, <strong>of</strong><br />

each sub<strong>system</strong>, (3.47), is defined as a rigid body mode <strong>of</strong> the hydraulically driven Stewart<br />

platform.<br />

The eigenvalues <strong>of</strong> the sub<strong>system</strong>s can be calculated from the respective determinants,<br />

i = ; p ic( l p<br />

c= i +<br />

2<br />

bp<br />

i=c<br />

2<br />

ja)� (3.48)<br />

1 The most appropriate choice for the state equations in modelling e.g. in numerical simulation, explicit model<br />

<strong>based</strong> <strong>control</strong>, are <strong>of</strong> course the platform pose, sx and velocity, _x as state instead <strong>of</strong> q and its derivative and require<br />

variational actuator stiffness. The model which results from that choice will be given with equation (3.66) in<br />

Section 3.4.4


104 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

where the product <strong>of</strong> l and b is assumed to be neglectable. The eigenfrequencies can be<br />

considered the rigid modes and are mainly determined by the hydraulic stiffness, c, and<br />

the singular values <strong>of</strong> the inverse mass matrix, i, taken w.r.t the actuators. These singular<br />

values can be interpreted as the inverse <strong>of</strong> the generalised masses, which the actuators will<br />

have to accelerate moving along the six orthogonal directions described by the columns <strong>of</strong><br />

the unitary decoupling matrix, U.<br />

The characterisation with U can already be found by H<strong>of</strong>fman [51], who uses this decoupling<br />

to analyse a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>, a Stewart platform, in combination<br />

with a decentralised feedback. It has not seen to be applied after that time. However, as<br />

will be shown in the sequel, both in model analysis, identification and decoupling <strong>control</strong> it<br />

can be very appropriate, since the decoupling is seen to hold in practice even in the face <strong>of</strong><br />

moderate additional variations/ dynamics.<br />

The parallel structure makes the calculation <strong>of</strong> the directions in U simpler. It can be<br />

calculated from a ’scaled’ jacobian, Jm without inversion by assuming M ;1<br />

t does not vary<br />

and can be calculated in advance<br />

Jl�x(sx)M ;1<br />

t J T l�x (sx) = Jm(sx)J T m (sx)<br />

= U (sx) 1=2 (sx)V T (sx)V (sx) 1=2 (sx)U T (sx) (3.49)<br />

As shown in the previous sections, the stiffness <strong>of</strong> the actuators, C, also varies with<br />

the position <strong>of</strong> the actuator. With different stiffness values, the decoupling with U will<br />

not be exact anymore. However, in many practical cases, as will be analyzed by taking<br />

numerical design values in the next sections and experimentally found parameters in the next<br />

chapter, a strongly decoupled <strong>system</strong> results calculating U and only from the platform<br />

pose dependent Jacobian and a constant inverted platform mass matrix M t.<br />

In the next section, first, the specific properties <strong>of</strong> the SRS <strong>motion</strong> <strong>system</strong> will be discussed.<br />

Then, possible neglection <strong>of</strong> the variation <strong>of</strong> M t and appropriate choice <strong>of</strong> its value<br />

will be treated.<br />

3.4.2 Dynamical and kinematical properties <strong>of</strong> the SRS-<strong>motion</strong> <strong>system</strong>;<br />

Design aspects<br />

Having introduced the generic formulas to describe the mechanics and hydraulics <strong>of</strong> a (<strong>flight</strong><br />

<strong>simulator</strong>) <strong>motion</strong> <strong>system</strong>, additional insight can be gained by considering a specific design<br />

example <strong>of</strong> such a <strong>system</strong> and analyzing the Simona Research Simulator characteristics.<br />

The analysis presented in this section can be done before actually building a <strong>motion</strong> <strong>system</strong><br />

and can/ should therefore be taken into account in the process <strong>of</strong> design. The parameters for<br />

the kinematics <strong>of</strong> the SRS were already given in Table 2.1.<br />

The inertial properties <strong>of</strong> the SRS-<strong>motion</strong> <strong>system</strong> discussed here will be the values calculated<br />

in the design <strong>of</strong> a fully operational <strong>flight</strong> <strong>simulator</strong>. In that case the centre <strong>of</strong> gravity<br />

(cog) <strong>of</strong> the moving body excluding the actuators is predicted 1 m above the upper centre<br />

gimbal point (cgb). The cog will be taken as the origin <strong>of</strong> the moving reference frame. The<br />

vector from cog to cgb will thus be given by<br />

x m T<br />

cgb = 0 0 1<br />

(3.50)


3.4 Hydraulically driven Stewart platform 105<br />

It is not required to take the c.o.g. as the origin but usually specific design values, such as<br />

the mass matrix are more easily interpreted. Further, an important point in simulation, the<br />

design eye point (DEP) <strong>of</strong> the pilot is predicted very close to this point, only :2 m higher.<br />

x m T<br />

DEP = 0 0 ;0:207<br />

(3.51)<br />

As discussed, <strong>flight</strong> simulation usually requires a neutral position <strong>of</strong> the <strong>motion</strong> <strong>system</strong>. The<br />

high pass wash-out filter characteristics should keep the <strong>motion</strong> <strong>system</strong> not to far away from<br />

this point to prevent the actuators from running out <strong>of</strong> stroke. The most straight forward<br />

choice <strong>of</strong> the neutral position is at half stroke <strong>of</strong> all the actuators. The moving reference<br />

frame <strong>of</strong> the SRS is in that case 3:38 m above the lower centre gimbal point, which is taken<br />

as the origin <strong>of</strong> the inertial grounded frame (or its nominal position in case <strong>of</strong> a flexibly<br />

attached foundation).<br />

x g<br />

T<br />

neutral = 0 0 ;3:38 (3.52)<br />

Since the <strong>motion</strong> <strong>system</strong> will normally be close to this point 2 the (dynamic) characteristics<br />

are <strong>of</strong>ten specified at this point. The jacobian in neutral position, sx n, is given by:<br />

Jl�x(sxn) =<br />

2<br />

6<br />

4<br />

0:1943<br />

;0:4668<br />

0:2725<br />

0:2725<br />

;0:4668<br />

;0:4269<br />

;0:0452<br />

0:3817<br />

;0:3817<br />

0:0452<br />

;0:8832<br />

;0:8832<br />

;0:8832<br />

;0:8832<br />

;0:8832<br />

0:3386<br />

;1:2204<br />

;1:5589<br />

1:5589<br />

1:2204<br />

1:6047<br />

;1:0955<br />

;0:5091<br />

;0:5091<br />

;1:0955<br />

;0:7011<br />

0:7011<br />

;0:7011<br />

0:7011<br />

;0:7011<br />

7<br />

5<br />

0:1943 0:4269 ;0:8832 ;0:3386 1:6047 0:7011<br />

(3.53)<br />

Each column <strong>of</strong> this matrix specifies at what velocity the actuator should run in case only<br />

one <strong>of</strong> the platform states is to be moved with unit velocity ( _ l = Jl�x _x). The columns,<br />

in this specific pose, are orthogonal except for the pitch/ surge and roll/ sway direction.<br />

This strongly influences the hydromechanical modes <strong>of</strong> the <strong>system</strong> evaluated along the platform<br />

directions. The rows <strong>of</strong> the matrix specify the platform forces felt in case <strong>of</strong> a unit<br />

force <strong>of</strong> one <strong>of</strong> the actuators. These rows are not orthogonal showing strong coupling <strong>of</strong><br />

neighbouring actuators and strongly influence the mass matrix evaluated along the actuator<br />

coordinates.<br />

Recalling the discussion on the jacobian in Section 2.2.9, it is possible to calculate what<br />

can be the maximum velocity in a certain platform direction given a maximum actuator<br />

velocity, given the jacobian. If the infinity norm (max) <strong>of</strong> the actuator velocity vector is<br />

given, the 1-norm (added absolute values) <strong>of</strong> the rows <strong>of</strong> the inverse jacobian specify the<br />

maximum attainable platform velocity.<br />

k _xik1 = kJ ;1<br />

i� k1k _ l1 (3.54)<br />

Given a maximal applicable actuator force, a similar formula can be used to calculate the<br />

maximal attainable platform force (take 1-norm <strong>of</strong> each row <strong>of</strong> the jacobian). In both cases<br />

2Directing the gravitational vector properly <strong>of</strong>ten results in a certain <strong>of</strong>f set from neutral as an airplane is usually<br />

operated slightly pitched.<br />

3


106 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Acceleration limits [m/s2,100 deg/s2]<br />

10 2<br />

10 1<br />

10 0<br />

10 −1<br />

10 −1<br />

10 −2<br />

x,y z m_z m_x,m_y<br />

10 0<br />

Frequency [Hz]<br />

Fig. 3.21: Approximate accelerational limits for sinusoidal platform <strong>motion</strong> considered at<br />

the neutral position, along the platfrom directions, surge (x), sway (y), heave (z),<br />

roll (mx), pitch (my) and yaw (mz).<br />

it should be noted that the other platform variables are allowed to vary. A more appropriate<br />

bound is therefore given by the maximal attainable velocity/ force in a specific direction not<br />

moving/ forcing the others.<br />

For the velocity this bound can be found by<br />

And for the force<br />

_xi�max = _ lmax=kJ �ik1<br />

10 1<br />

(3.55)<br />

fxi�max = fa�max=kJ ;1<br />

i� k1 (3.56)<br />

The maximal actuator force and velocity are parameters, which are determined by the specific<br />

design <strong>of</strong> the hydraulic <strong>system</strong>. The maximum valve flow, max, together with the<br />

hydraulic actuator operational area, Ap, determine the maximum velocity<br />

_qmax = max=Ap =SRS 1 m=s (3.57)<br />

The maximal applicable actuator force can be found by multiplication <strong>of</strong> the supply pressure,<br />

Ps, with this area, Ap,<br />

fa�max = PsAp =SRS 40 kN (3.58)<br />

The operational area is clearly a trade-<strong>of</strong>f parameter in design for velocity vs. force given a<br />

specific available amount <strong>of</strong> energy. There are, however, also other constraints in choosing<br />

this area e.g. radial stiffness, and <strong>of</strong> course also the stroke <strong>of</strong> the actuator limits <strong>motion</strong>.<br />

A first estimate <strong>of</strong> this last limit can be found by using (3.55) from which the jacobian to


3.4 Hydraulically driven Stewart platform 107<br />

Pos. displacement x y z x y z<br />

[m� deg] 1.34 1.10 0.74 22.3 20.9 45.2<br />

Neg. displacement x y z x y z<br />

[m� deg] -1.05 -1.10 -0.69 -22.3 -30.0 -45.2<br />

Est. displacement (1 it.) x y z x y z<br />

[m� deg] 1.34 1.46 0.71 23.1 22.5 52.9<br />

Max. velocities _x _y _z !x !y !z<br />

[m=s� deg=s] 2.14 2.34 1.13 36.8 35.7 81.7<br />

Max. force fx fy fz mx my mz<br />

[kN,kNm] 47.7 46.3 212.0 159.1 174.1 168.1<br />

Max. acceleration x y z _!x _!y _!z<br />

[m=s 2 �kdeg=s 2 ] 13.5 12.9 55.4 1.39 1.24 1.01<br />

Freq. <strong>of</strong> pos. to vel. constr. x y z x y z<br />

[Hz] 0.32 0.39 0.29 0.31 0.32 0.35<br />

Freq. <strong>of</strong> vel. to acc. constr. x y z x y z<br />

[Hz] 0.99 0.87 7.71 5.85 5.56 2.01<br />

Table 3.6: SRS-properties considered at neutral position and c.o.g.<br />

euler parameter velocities for the rotational variation can be calculated. Iteratively using this<br />

formula (Newton-Raphson) and lmin�max, the least available stroke, a precise limit can<br />

be found very fast as opposed to a bilinear search. This Newton-Raphson iteration usually<br />

requires three or four steps, ending up far below measurable error in a well designed <strong>system</strong>.<br />

With a given mass matrix, it is also possible to calculate the maximal attainable acceleration.<br />

The specific mass matrix (design values) <strong>of</strong> the SRS is discussed in the next section. All<br />

these values calculated for the SRS in the neutral position are specified in Table 3.6.<br />

Given sinusoidal inputs, the velocity and positional constraints limit the maximal attainable<br />

acceleration for lower frequencies. For the 1-d.o.f. hydraulic actuator, Viersma [153]<br />

specifies exact performance limit diagrams in the velocity domain. In Fig. 3.21 approximate<br />

performance diagrams for the 6-d.o.f. <strong>motion</strong> <strong>system</strong> are given in the platform coordinates<br />

using the limits specified in Table 3.6. The precision <strong>of</strong> the approximate velocity and acceleration<br />

constraints for the sinusoidal <strong>motion</strong> is favourably influenced since the maximum<br />

velocity <strong>of</strong> the sinusoid is attained at neutral position where the velocity constraint is exact<br />

and the maximal acceleration is attained (with limited positional <strong>of</strong>fset at these high frequencies)<br />

at zero velocity with no centripetal and coriolis accelerations. Gravity, however,<br />

results in a more serious <strong>of</strong>fset to be taken into account. Given the force which needs to<br />

be applied to compensate gravity, the least available maximal force required to move in a<br />

specific direction determines the maximal attainable acceleration.<br />

The frequencies where the positional constraint, p max, and velocity constraint, vmax,<br />

meet, fpv, and velocity and accelerational constraint meet, f av, are given by fpv = vmax<br />

2 pmax<br />

and fpv = amax<br />

2 vmax .Afirst estimate <strong>of</strong> fpv is equal for all platform directions and does only<br />

depend on the actuator specifications: ~ fpv = vact�max<br />

2 pact�max =SRS<br />

1<br />

2 0:625<br />

=0:25Hz. From


108 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

−z [m]<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Upper Gimbal<br />

Lower Gimbal 1 Lower Gimbal 2 (6)<br />

−1.5 −1 −0.5 0<br />

−y [m]<br />

0.5 1 1.5<br />

Fig. 3.22: Envelope and largest circle within this envelope <strong>of</strong> a 2-d.o.f. parallel planar manipulator.<br />

Parameters taken as for actuator 1 and 6 <strong>of</strong> the SRS except for merging<br />

<strong>of</strong> the upper gimbals.<br />

Fig. 3.21 it can be observed that the actual constraint will be attained at a frequency which<br />

is somewhat higher. It is interesting to observe that already from :3 Hz and higher it is not<br />

so much the finite stroke <strong>of</strong> the actuators which limits the (<strong>flight</strong>) simulation but the velocity<br />

constraint. For these frequencies it is <strong>of</strong> no use to build a larger <strong>motion</strong> <strong>system</strong>. One should<br />

first choose a larger valve.<br />

Only looking at heave, roll and pitch, the operational area <strong>of</strong> the actuators could have<br />

been chosen somewhat smaller as velocity is the main constraint for these directions. For<br />

surge and sway, however, the force constraint comes in at much lower frequencies. In this<br />

respect the jacobian (kinematical structure) <strong>of</strong> the Stewart platform is also a trade-<strong>of</strong>f design<br />

parameter where e.g. with the SRS neutral position high surge velocity can be attained<br />

(twice as high as the maximal actuator velocity since the actuators are mainly rotating)<br />

compromised with limited applicable force in this direction.<br />

The specific values given in Table 3.6 should be treated with extreme care since they<br />

can vary considerably over the <strong>motion</strong> envelope. Especially the <strong>motion</strong> envelope itself has<br />

a complex non convex structure. Consider a simplified (planar) <strong>system</strong> with two actuators<br />

having limited stroke manipulating a point. In Fig. 3.22 this structure is drawn with the<br />

lower gimbals the same distance (2:5 m) apart as actuator 1 (2,4) and 6 (3,5) in the SRS<br />

and the same constraints on the minimal and maximal length (2:08 m,3:33 m). The upper<br />

gimbals have been merged (are :2 m apart in the SRS). Choosing the neutral point somewhat<br />

lower in this construction, would lead to a considerable growth <strong>of</strong> the maximal attainable<br />

sway moving straight away from this point. But the constraint towards the attainable heave<br />

would become even more asymmetric in doing so. Thus changing the neutral point in this<br />

way will probably not be a appropriate decision.<br />

Appropriately describing the <strong>motion</strong> envelope <strong>of</strong> the general 6-d.o.f. Stewart platform


3.4 Hydraulically driven Stewart platform 109<br />

in task space is still an open problem [27]. The inverse problem <strong>of</strong> checking whether a<br />

specific task space envelope fits in a specific kinematical structure can be interesting for<br />

(<strong>flight</strong>) simulation and can sometimes be solved. Consider e.g. a normed task space for the<br />

2-d.o.f manipulator <strong>of</strong> Fig. 3.22. With the 2-norm, ( p x T x), this space describes a ball and<br />

the maximal radius (norm) <strong>of</strong> the circle can be found in the half stroke neutral position and<br />

is exactly equal to this half stroke (0:625 m).<br />

With the general Stewart platform, exactly the same bound can be found if the platform<br />

is not allowed to rotate since the upper gimbals in that case are also only allowed to move in<br />

this same ball. So the maximal ball <strong>of</strong> translational <strong>motion</strong> <strong>of</strong> the SRS which is applicable<br />

can be found in the all actuator half stroke neutral position and has radius r =0:625 m. For<br />

any ball larger than this, translations can be found which run (some <strong>of</strong>) the actuators out <strong>of</strong><br />

stroke.<br />

In case <strong>of</strong> orientational variations together with translation, a radius can be found which<br />

satisfies the constraints but is not necessarily maximal. Recall the <strong>motion</strong> <strong>of</strong> any upper<br />

gimbal<br />

which can be bounded by<br />

xa�i(sx) =c + T ( 13)ai<br />

kxak2<br />

(3.59)<br />

p 2ksxk (3.60)<br />

In case <strong>of</strong> the SRS kak = 1:6m when using the centre upper gimbal point as origin (and<br />

gets larger/worse in using DEP or cog). So in case <strong>of</strong> the SRS r = :44m and this means a<br />

maximal rotation <strong>of</strong> =5:7 . It would be interesting to know whether it is possible to use<br />

more structure concerned with the interaction between rotation and translation in order to<br />

arrive at possibly less conservative results.<br />

In some cases it is favourable to define a (hyper) ellipsoid instead <strong>of</strong> a ball in which<br />

the task space states are allowed to move. E.g. in case <strong>of</strong> Fig. 3.22 it would be appropriate<br />

to choose a larger allowable variation in y-direction. Also this problem is in some cases<br />

tractable in optimising a norm given quadratic constraints, which can be transformed to<br />

linear matrix inequalities. The problem has to made convex, however. This can be attained<br />

by replacing the constraints w.r.t. the minimal actuator lengths by convex approximations<br />

in the task space e.g. a planes in the tangent space <strong>of</strong> the l min balls. The ellipsoidal form is<br />

attained by multiplying xa by a (positive definite) scaling matrix S.<br />

E.g. maximize =x T x such that<br />

(Sx ; b) T (Sx ; b) 0 8kxk (3.62)<br />

By use <strong>of</strong> the S-procedure [18], the ’uncertainty’ ball <strong>of</strong> x’s can be removed from the equations,<br />

leaving a tractable problem.


110 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Parameter Description Value Unit<br />

ma Mass upper part actuator (piston) 120 [kg]<br />

mb Mass lower part actuator (cylinder) 135 [kg]<br />

ra Distance upper gimbal to cog piston 0:7 [m]<br />

rb Distance lower gimbal to cog cylinder 0:5 [m]<br />

ia Inertia upper part actuator wrt cog 20 [kgm 2 ]<br />

ib Inertia lower part actuator wrt lower gimbal 36 + mbr 2<br />

b<br />

Table 3.7: Inertial parameters SRS-actuators.<br />

[kgm 2 ]<br />

Limitations <strong>of</strong> a hydraulically driven <strong>motion</strong> <strong>system</strong> with specific kinematical and dynamical<br />

properties have been quantified in this section. This knowledge can be used in<br />

design <strong>of</strong> the <strong>system</strong> itself but also in the design <strong>of</strong> experiments in which it is <strong>of</strong>ten important<br />

not to approach such limits to closely.<br />

3.4.3 Actuator inertial effects<br />

In model <strong>based</strong> <strong>control</strong> choice <strong>of</strong> an appropriate model is important. The model has to<br />

describe the <strong>control</strong> relevant aspects <strong>of</strong> the <strong>system</strong>. On the other hand it should not be too<br />

complex since it has to be calculated on a real-time <strong>control</strong> computer. W.r.t. modelling, the<br />

question arises whether the mass properties <strong>of</strong> the actuators have to be taken into account.<br />

And if so, how this should be done in a convenient way.<br />

With the currently used <strong>simulator</strong> <strong>motion</strong> platforms the weight <strong>of</strong> the actuators is relatively<br />

low (ca. 250 kg) w.r.t. the platform (up to 12:5 tons). In these cases the weights do<br />

not contribute significantly to the <strong>motion</strong> <strong>of</strong> the total <strong>system</strong>. As the designed weight <strong>of</strong> the<br />

<strong>simulator</strong> was reduced drastically (to ca. 3:2 tons) with the Simona research <strong>simulator</strong>, one<br />

has to reevaluate this assumption.<br />

Let’s look into the mass contribution <strong>of</strong> the <strong>system</strong>. The one body <strong>simulator</strong> has according<br />

to the design (finite element modelling) roughly the following mass matrix, M c,<br />

in platform coordinates w.r.t the c.o.g. which is calculated 1 meter above the centre upper<br />

gimbal point.<br />

Mc =<br />

2<br />

6<br />

4<br />

3200 0 0 0 0 0<br />

0 3200 0 0 0 0<br />

0 0 3200 0 0 0<br />

0 0 0 7000 0 500<br />

0 0 0 0 7000 0<br />

0 0 0 500 0 8000<br />

3<br />

7<br />

5<br />

(3.63)<br />

The inertia <strong>of</strong> the <strong>simulator</strong> is somewhat higher than the weight but this does not necessarily<br />

mean that the impact is higher since the actuators are connected on a circle with radius<br />

larger than 1m. The main axes <strong>of</strong> inertia do not exactly correspond to the platform body<br />

axes. The y-axis does, as the <strong>system</strong> can be mirrored w.r.t the xz-plane. The projectors on<br />

the upper back <strong>of</strong> the <strong>simulator</strong> are the main reason for the xz cross term.


3.4 Hydraulically driven Stewart platform 111<br />

Case/Actuator 1 2 3 4 5 6<br />

Neutral position lmid lmid lmid lmid lmid lmid<br />

Extreme pitch/surge lmin lmin lmax lmax lmin lmin<br />

Extreme twist (yaw) lmax lmin lmax lmin lmax lmin<br />

Extreme roll/sway lmid lmin lmin lmax lmax lmid<br />

Table 3.8: Actuator lengths for platform poses considered.<br />

Case/Platform pose x y z 1 2 3<br />

Neutral position 0 0 -3.38 0 0 0<br />

Extreme pitch/surge 2.00 0 -2.43 0 -0.32 0<br />

Extreme twist (yaw) 0 0 -3.02 0 0 -0.52<br />

Extreme roll/sway 0.14 0.41 -3.25 0.20 -0.01 0<br />

Table 3.9: Platform poses considered.<br />

Recalling the additive mass matrix term, Ma�n, in (2.123), given by<br />

Ma�n =<br />

6X<br />

i=1<br />

J T ai�x (Mma�i + Mia�i�ib�i )Jai�x� (3.64)<br />

which consists <strong>of</strong> the contribution <strong>of</strong> all six actuators. In neutral position it becomes the following,<br />

using the relations determining (2.122), and the inertial parameters for the actuator<br />

given in Table 3.7.<br />

Ma�n =<br />

2<br />

6<br />

4<br />

500 0 0 0 ;525 0<br />

0 500 0 525 0 0<br />

0 0 675 0 0 0<br />

0 525 0 1400 0 0<br />

;525 0 0 0 1400 0<br />

0 0 0 0 0 1325<br />

3<br />

7<br />

5<br />

(3.65)<br />

The actuators contribute considerably to the total mass matrix (about 15 %). The terms in<br />

the upper right and lower left block contribute to a lower cog <strong>of</strong> approx. .1 m. Remarkable<br />

is the fact that the weight experienced in moving in z-direction will be higher than in x<br />

or y-direction, due to the actuators. The actuator mass matrix will vary as a function <strong>of</strong><br />

platform position since the actuators move w.r.t. the platform. These variations, however,<br />

are rather small. By taking a constant mass matrix in platform coordinates into the model<br />

<strong>based</strong> <strong>control</strong>ler, a relatively simple but accurate choice is made.<br />

If a Stewart Platform with even less relative weight is used, the full model should be<br />

taken into account. Apart from extra calculational effort this does not result in additional<br />

problems if the modelling steps described, are taken.<br />

With respect to the nonlinear quadratic velocity terms in the model describing the coriolis<br />

and centripetal terms, the constraints on the velocities are important. The most restrictive


112 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

10 2<br />

10 0<br />

−500<br />

−1000<br />

10 2<br />

10 0<br />

0<br />

100<br />

0<br />

−100<br />

10 0<br />

10 2<br />

10 0<br />

10 2<br />

Fig. 3.23: Bode plots from input valve voltage (u) to pressure difference (dP i) <strong>of</strong> linearised<br />

multivariable hydraulically driven mechanical <strong>system</strong> models. Masses: Simona<br />

design values. Rows 1,2 gives response <strong>of</strong> input 1 to outputs 1 (full), 2-6 (dotted).<br />

Rows 3,4 give responses <strong>of</strong> the diagonal elements along the decoupled rigid body<br />

modal directions (1,1:full), (2,2-6,6:dotted). Column 1: mid position, Column<br />

2: extreme pitch/ surge, Column 3: extreme twist, Column 4: extreme roll/<br />

sway. X-axis: frequency (Hz).<br />

constraint, which holds for both pitch and roll <strong>of</strong> 35 =s =0:61rad=s in the neutral position,<br />

was taken as the rotational velocity bound in calculating the influence <strong>of</strong> these nonlinear<br />

terms in [119] and it was found that these only contribute up to ca. 500 N for the SRS.<br />

3.4.4 Analyzing the SRS hydraulically driven <strong>motion</strong> <strong>system</strong> model<br />

Having introduced the specific parameters <strong>of</strong> the SRS, the SRS-model can be analyzed considering<br />

the interaction between the hydraulics and mechanics in this 6-d.o.f. <strong>system</strong>. It will<br />

be evaluated whether the hydromechanical rigid body modes describe the main characteristics<br />

<strong>of</strong> the <strong>system</strong>. Along the rigid body modal directions, it is expected that the <strong>system</strong><br />

locally behaves as six independent 1-d.o.f. hydraulic actuators moving a mass. This has to<br />

be checked also as is the predictability <strong>of</strong> these directions. The influence <strong>of</strong> the main mechanical<br />

nonlinearity, the platform pose dependent variation <strong>of</strong> the jacobian, J lx, has to be<br />

quantified.<br />

The state equations, constructed from Fig. 3.20, describing the nonlinear dynamical<br />

10 0<br />

10 2<br />

10 0<br />

10 2


3.4 Hydraulically driven Stewart platform 113<br />

frb 1 2 3 4 5 6<br />

1: (y� x) (x� - y) ( z) ( y� -x) ( x�y) (z)<br />

[Hz] 4.5 4.5 7.6 13.3 13.5 14.8<br />

2: (x� z� - y) (y� x� - z) ( z� - x) ( y� -x� z) ( x�y) (z)<br />

[Hz] 4.1 5.8 9.1 15.5 16.1 17.4<br />

3: ( z� -z ) (y� x� y) (x� - y� x) ( y�x) ( x� z� -y) (z� z� y)<br />

[Hz] 3.6 4.5 4.5 16.1 16.3 17.2<br />

4: (x� -y� - y) (y� x� x� - y) ( z) ( x�z�-y) ( y�x) (z� - x�x)<br />

[Hz] 4.9 5.4 7.7 12.0 14.3 18.0<br />

Table 3.10: <strong>Model</strong> eigenfrequencies, frb, in Hz <strong>of</strong> the rigid body modes and roughly the<br />

rigid body modal directions. 1: Neutral position, 2: Extreme pitch/ surge, 3:<br />

Extreme twist (yaw), 4: Extreme roll/ sway. i is the rotation around the i th<br />

axis.<br />

model from the output oil flows, om, are given by<br />

x<br />

_<br />

dP o<br />

;bI aI<br />

;aI ;lI<br />

=<br />

;1<br />

M<br />

t J T lx (sx) 0<br />

Jlx(sx) 0<br />

0 I<br />

:::+ M ;1<br />

t<br />

0 Ch(sx)<br />

_x<br />

dP o<br />

+ 0<br />

I<br />

G(sx) +C(sx� _x)<br />

0<br />

::: (3.66)<br />

om + :::<br />

In this case it is assumed that there are similar actuators (operational area, a, viscous friction,<br />

b and leakage, l) easily replaced by the respective matrices, A� B� L otherwise. By taking<br />

the platform pose, sx� as state variable, the model is valid for singular poses as well. Only<br />

if actuators would reach zero length, the jacobian is not defined.<br />

The two-sided coupling with all ( om,dPo) pairs to the transmission lines and valves can<br />

readily be made. For a reduced but fairly good model, each transmission line has four states<br />

and each valve has two. Further, the six platform pose coordinates, sx, have to be calculated<br />

through integration <strong>of</strong> the velocities. The model thus consists <strong>of</strong> 12 + 6(2 + 4) + 6 = 54<br />

states.<br />

This model with the SRS-parameters has been analyzed for several poses. Together<br />

with the neutral position, three extreme poses were taken into account. Actuator lengths<br />

are given in Table 3.8. The extreme pitch/ surge and twist only consist <strong>of</strong> extreme actuator<br />

positions. As the oil stiffness varies parabolically with the position, the extreme roll/ sway<br />

pose is important since it contains actuators with maximal (extreme position) and minimal<br />

(mid position) stiffness.<br />

The platform poses, which are given in Table 3.9, do result from these actuator lengths.<br />

The euler parameters can be translated to angles by taking 12 per :1 up to :6 with 1


114 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Case/mode 1 2 3 4 5 6<br />

Neutral position 8.8 8.8 3.0 1.0 1.0 0.8<br />

Extreme pitch/surge 14.8 7.4 3.0 1.0 0.9 0.7<br />

Extreme twist (yaw) 19.0 12.2 12.2 0.9 0.9 0.8<br />

Extreme roll/sway 9.0 7.8 3.6 1.8 1.0 0.6<br />

Table 3.11: Generalised masses (tons). The large differences in (generalised) mass easily<br />

result in interaction. E.g. consider a force along the 8.8 ton direction in neutral<br />

position, which is only 1 misaligned into the 0.8 ton direction. This will result<br />

in 20% interaction. In more extreme positions e.g. pitch/ surge this will even<br />

amount to more than 40%!<br />

accuracy. In the first two rows <strong>of</strong> Fig. 3.23 the Bode plots <strong>of</strong> the linearised models in these<br />

positions are given from the first valve input voltage to the various pressures, dP i, taken<br />

over the valves. The hydromechanical modes are visible and are highly interacting in the<br />

frequency area between 1 Hz and 30 Hz. The phase remains within 90 for the (1� 1)<br />

transfer function up till the point where the valve itself starts to be relevant (> 70 Hz).<br />

Above 30 Hz the measurable interaction strongly fades as the nondiagonal terms lack zeros<br />

in the transfer functions, which can be interpreted as springs in between input/ cause and<br />

output/ point <strong>of</strong> measurement.<br />

In the last two rows the transfer functions have been considered from the other inputs and<br />

outputs (only diagonal terms) resulting from the approximate decoupling matrix U defined<br />

in (3.44). This matrix was calculated from the approximate mass matrix (not taking into<br />

account the variation due to the specific inertial properties <strong>of</strong> the actuators) and does not<br />

take into account the varying actuator stiffness. Still the modes in the <strong>system</strong> can clearly<br />

be distinguished. In the hydromechanical frequency interaction area, only one mode per<br />

transfer function results in a single shift in phase <strong>of</strong> 180 . All the nondiagonal transfer<br />

functions in these coordinates are small over the full frequency area considered.<br />

Considering stability from a <strong>control</strong> point <strong>of</strong> view, the connection <strong>of</strong> a hydraulic circuit<br />

still enables a simple passivity preserving dissipating feedback from pressure/ force supplied<br />

to the mechanical <strong>system</strong> to the oil flow through the valve instead <strong>of</strong> velocity feedback to<br />

force in a mechanical <strong>system</strong>.<br />

The eigenfrequencies related to the hydromechanical rigid body modes <strong>of</strong> the <strong>system</strong> are<br />

given in Table 3.10. There is considerable variation for the different rigid body modes both<br />

for a specific pose as for the different poses considered. As each extreme position considered<br />

was only given a specific variation <strong>of</strong> the platform pose w.r.t the neutral position, the rigid<br />

body modal directions in platform coordinates, given by J ;1<br />

lx U, can still be interpreted.<br />

E.g. in extreme pitch/ surge, symmetric involving x� z� y and asymmetric modes involving<br />

y� x� z. If all platform coordinates are varied this structure will be lost.<br />

Looking at the Bode plots in Fig. 3.23, it can be observed that building up a unit pressure<br />

in a specific direction will take very different input values (more than a factor 10 variation).<br />

Very slight misalignment (wrong input) will easily result in large coupling. This is roughly<br />

linear with the generalised masses, which affect the eigenfrequencies through a square root.


3.4 Hydraulically driven Stewart platform 115<br />

Case/actuator 1 2 3 4 5 6<br />

Neutral position 18 18 18 18 18 18<br />

Retraction gain -1.08 -1.08 -1.08 -1.08 -1.08 -1.08<br />

Extraction gain +0.90 +0.90 +0.90 +0.90 +0.90 +0.90<br />

Extreme pitch/surge 26 67 -42 -42 67 26<br />

Retraction gain -1.12 -0.57 -1.19 -1.19 -0.57 -1.12<br />

Extraction gain +0.86 +1.29 +0.76 +0.76 +1.29 +0.86<br />

Extreme twist (yaw) -24 48 -24 -24 48 -24<br />

Retraction gain -0.87 -1.21 -0.87 -1.21 -0.87 -1.21<br />

Extraction gain +1.11 +0.72 +1.11 +0.72 +1.11 +0.72<br />

Extreme roll/sway 21 32 25 6 11 18<br />

Retraction gain -1.10 -1.15 -1.12 -1.03 -1.05 -1.09<br />

Extraction gain +0.89 +0.83 +0.86 +0.97 +0.95 +0.90<br />

Table 3.12: Static load as percentage <strong>of</strong> maximal load F max = PsA and the resulting normalized<br />

valve gain discontinuity.<br />

The generalised masses, 1= i from (3.44), are given in Table 3.11.<br />

Already in the neutral position, the SRS is expected to have very different characteristics<br />

along the rigid body modal directions. In the sway/ roll and surge/ neg. pitch directions the<br />

<strong>system</strong> will experience approx. 9 tons where the sway/ neg. roll and surge/ pitch directions<br />

only actuator forces to accelerate 1 ton <strong>of</strong> load have to be applied. In the extreme positions<br />

the largest mass ratio (mass matrix conditioning) along the different modes grows from nine<br />

to more than twentyfour in extreme twist. Although the jacobian, J lx, has a main influence<br />

on this variation, its conditioning is not a very well predictor for this factor. It grows from<br />

4:5 to 5:2 going from the neutral position to extreme twist.<br />

With the SRS design parameters, gravity requires approx. 18 % <strong>of</strong> supply pressure in<br />

the neutral position. The platform contributes 15 % and the moving parts <strong>of</strong> the actuators<br />

3 %. The contribution <strong>of</strong> the lower parts is almost neglectable. This percentages can rise<br />

extensively in other platform poses. In the extreme surge/ pitch position considered, actuator<br />

2 and 5 will have to carry two third <strong>of</strong> the theoretically maximal weight. In practice, two<br />

third is relatively high to take care <strong>of</strong> the static forces only. This means that the SRS design<br />

weight and height <strong>of</strong> the c.o.g. are in fact maximal for this <strong>motion</strong> <strong>system</strong> to operate safely.<br />

Another important aspect <strong>of</strong> the static loads on the actuators is the nonlinear characteristic<br />

<strong>of</strong> the valve as given in (3.16). A preload results in a discontinuous valve voltage to flow<br />

gain as im = u p (1 ; sgn(u)dPi) for normalized variables. In Table 3.12 the required<br />

preload for the four platform poses considered is given with the resulting valve gains for<br />

actuator extension and retraction. Negative percentages for the load mean that retraction<br />

preload has to be applied to statically balance the platform. The all actuator minimal length<br />

position is not the global minimum for the potential gravitational energy. As a result, the<br />

<strong>system</strong> will not always go to this pose if for some reason the preload disappears e.g. pump<br />

supply pressure drop to tank pressure in shut down.<br />

All together, the hydraulically driven Stewart platform model can be transformed to six


116 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

Direction Mfii [tons� tons m 2 ] !fi [Hz] fi<br />

xf 45 2 7 0:05<br />

yf 45 2 7:5 0:05<br />

!f 80 2 6 0:10<br />

Table 3.13: Parameters taken for the foundation.<br />

independent 1-d.o.f. hydraulically driven dynamic and static masses for each pose using<br />

the variable unitary decoupling matrix, U (sx), describing the rigid body modal directions.<br />

Analysis <strong>of</strong> the SRS <strong>motion</strong> <strong>system</strong> design showed considerable differences in the rigid<br />

body eigenfrequencies among these 1 d.o.f.-<strong>system</strong>s for the SRS.<br />

3.4.5 Connection <strong>of</strong> a flexible foundation to the SRS <strong>system</strong> model<br />

The hydraulically driven Stewart platform model <strong>of</strong> (3.66) can be extended with a model<br />

including parasitic dynamics such as a flexible foundation. This model is used to predict<br />

in what way the dynamics <strong>of</strong> the <strong>system</strong> is influenced by such distortion. A model <strong>of</strong> a<br />

flexible foundation was introduced in Section 2.4.7 described by (2.126). Combination <strong>of</strong><br />

these models leads to the following state equation<br />

22<br />

6<br />

6<br />

4<br />

6<br />

4<br />

2<br />

6<br />

4<br />

_xf<br />

xf<br />

x<br />

_<br />

dP o<br />

3<br />

7<br />

5 =<br />

2<br />

6<br />

4<br />

I 0 0 0<br />

0 M ;1<br />

f 0 0<br />

0 0 M ;1<br />

t 0<br />

0 0 0 Ch<br />

0 I 0 0<br />

Cf ;Bf ; J T f BJf ;J T f BJlx J T f A<br />

0 ;J T lx BJf ;J T lx BJlx J T lx A<br />

0 ;AJf ;AJlx ;L<br />

32<br />

7<br />

6<br />

5<br />

6<br />

4<br />

3<br />

7<br />

5<br />

xf<br />

_xf<br />

_x<br />

dP o<br />

::: (3.67)<br />

3<br />

7<br />

5 +<br />

2<br />

6<br />

4<br />

0<br />

0<br />

0<br />

I<br />

3<br />

7<br />

5 om<br />

The foundation is allowed to move in the horizontal plane and introduces three additional<br />

modes, foundation surge, xf ,sway,yf , and yaw, !f , described by six states (xf ,<br />

_xf ). As an example, the parameters given in Table 3.13, are filled into the equations. These<br />

describe roughly the characteristics <strong>of</strong> the foundation on which the <strong>system</strong> was tested in the<br />

central workshop. The mass matrix <strong>of</strong> the foundation has numbers on the diagonal, which<br />

are ten times higher than those <strong>of</strong> the platform. The stiffness matrix and damping matrix<br />

are chosen such that the undamped eigenfrequencies and damping factors specified for the<br />

foundation alone result. The high mass <strong>of</strong> the foundation in relation to the platform lets<br />

these frequencies only slightly altered in connecting the <strong>motion</strong> <strong>system</strong>.<br />

In Fig. 3.24 <strong>of</strong> Appendix A, the <strong>system</strong> is evaluated in the frequency domain from 1 Hz<br />

to 50 Hz. The upper left bode amplitude plots shows the model actuator pressure difference<br />

responses to the oil flow into the first actuator (not all 36 responses as the other 30 are<br />

3<br />

7<br />

5


3.4 Hydraulically driven Stewart platform 117<br />

10 1<br />

10 0<br />

10 −1<br />

10 1<br />

10 0<br />

10 −1<br />

10 0<br />

10 0<br />

10 1<br />

10 1<br />

Frequency (Hz)<br />

10 1<br />

10 0<br />

10 −1<br />

10 1<br />

10 0<br />

10 −1<br />

10 0<br />

10 0<br />

10 1<br />

10 1<br />

Frequency (Hz)<br />

Fig. 3.24: Bode amplitude plots <strong>of</strong> linearised multivariable hydraulically driven mechanical<br />

<strong>system</strong> models without transmission lines but with flexible foundation. Upper<br />

left: neutral position from first input actuator flow om1 to all output pressures<br />

(dP o). Upper right: neutral position along the (original) rigid body modal direc-<br />

tions from all ^ om to ^<br />

dP o, nondiagonal terms dashdotted. Lower left: neutral<br />

position along the (original) rigid body modal directions from all ^ om to the<br />

accelerations (m=s 2 )(xf � yf� _!f ) <strong>of</strong> the foundation, nondiagonal terms dashed.<br />

Lower right: extreme pitch/ surge along the (original) rigid body modal directions<br />

from all ^ om to ^ dP o, nondiagonal terms dashdotted.<br />

similar) in the neutral position. The complex dynamics <strong>of</strong> the hydromechanical modes<br />

prevents one from recognising the parasitic dynamics. Using the decoupling transformation<br />

U calculated from the platform mass matrix the bode amplitude plots in the upper right part<br />

<strong>of</strong> the figure results. In the diagonal responses the fact that there are parasitic modes is<br />

directly visible. Three additional modes show up separately along the eigenvectors <strong>of</strong> the<br />

lowest three original platform modes. Further, the nondiagonal terms are not zero anymore<br />

though still quite small.<br />

As discussed in Section 3.3.3, both a flexible foundation as a mass/ spring/ damper<strong>system</strong><br />

connected to the platform result in similar characteristics evaluated at (combinations<br />

<strong>of</strong>) actuator coordinates. To evaluate whether the foundation is causing parasitic dynamics,<br />

one can measure accelerations at the foundation itself. The model predicts the acceleration<br />

frequency response at the neutral position given in the lower left part <strong>of</strong> Fig. 3.24 <strong>of</strong> Appendix<br />

A. This was calculated by multiplying the velocity frequency response <strong>of</strong> the model<br />

by j!. Remarkable is the fact that the peaks in the response, which are most prominent, are


118 3 Hydraulically driven <strong>motion</strong> <strong>system</strong>s<br />

not in all cases mainly the original modes <strong>of</strong> the foundation. Input along the higher platform<br />

eigenfrequencies are expected to result in relatively high (approx. 15 Hz) frequency vibrations<br />

and <strong>of</strong> the three main transfer functions the yaw platform direction to yaw foundation<br />

modal response peaks are maximal at approx. the original platform yaw mode.<br />

The fact that the parasitic modes influence only three platform modes is not generally<br />

true. In the lower right part <strong>of</strong> Fig. 3.24, the <strong>system</strong> is evaluated again along the original<br />

platform eigendirections. But now the platform has been put into the extreme surge/ pitch<br />

mode. The fact that the platforms c.o.g. is now approx. 2 m in front <strong>of</strong> the c.o.g. <strong>of</strong> the<br />

foundation makes the sway mode <strong>of</strong> the platform interact with both the foundation yaw<br />

and sway mode and through the foundation yaw mode with the platform yaw mode. The<br />

response along this direction shows four resonances. Also the decoupling is almost completely<br />

lost between 4 Hz and 15 Hz.<br />

So, parasitic dynamics such as a flexible foundation can have an important influence<br />

on the <strong>motion</strong> <strong>system</strong> characteristics. This is not always directly visible in the measurable<br />

responses (inputs and outputs <strong>of</strong> the actuators).<br />

3.5 Chapter Resume<br />

The <strong>motion</strong> <strong>system</strong>s rigid bodies interacting with the hydraulic actuator stiffnesses, the<br />

hydromechanical rigid body modes, describe the main characteristics <strong>of</strong> the hydraulically<br />

driven <strong>motion</strong> <strong>system</strong> dynamics. The eigendirections <strong>of</strong> these modes can be quite accurately<br />

predicted by taking the eigenvectors <strong>of</strong> the approximate actuator mass matrix consisting <strong>of</strong><br />

the variable jacobian, Jlx(sx), and the platform mass matrix. Along the rigid body modal<br />

directions the <strong>system</strong> locally behaves as six independent 1-d.o.f. hydraulic actuators moving<br />

a mass in every platform pose.<br />

Under mild conditions, these <strong>system</strong>s are passive evaluating pairs <strong>of</strong> the oil flow and<br />

pressure difference at the valve or at the actuator <strong>of</strong> the hydraulic servos or (collocated)<br />

combinations <strong>of</strong> these pairs, no matter what parasitic flexible mechanical <strong>system</strong>s are to be<br />

included. Any passive feedback over these pairs will result in stability <strong>of</strong> the full nonlinear<br />

<strong>system</strong>.<br />

The valve dynamics causes loss <strong>of</strong> passivity mostly relevant in combination with the<br />

high frequency (e.g. transmission line) modes. The transmission lines, though mainly resulting<br />

in high frequency dynamics, cause a considerable shift <strong>of</strong> the hydromechanical rigid<br />

body modes and introduce large differences at relative low frequencies (10 times smaller<br />

than the transmission line modes) between the measurable pressure at the valve and the<br />

equivalent force (pressure difference times operational area) the actuator induces into the<br />

<strong>motion</strong> <strong>system</strong>.<br />

Parasatic dynamics such as a flexible foundation or an additional mass/ spring/ damper<br />

<strong>system</strong> connected to the moving platform is easily included in the models introduced. This<br />

dynamics can have considerable influence on the <strong>system</strong>s characteristics and is most easily<br />

recognized evaluating along the hydromechanical rigid body modes. Only measuring at the<br />

actuators, the cause <strong>of</strong> the parasitic dynamics is not identifiable.


Chapter 4<br />

Parameter identification and<br />

model validation<br />

The models presented in the two previous sections were derived from the laws <strong>of</strong> physics.<br />

Where numerical examples were given, the parameter values were taken from design values<br />

or rough approximations <strong>of</strong> what could be expected in practice. These models can be<br />

valuable for the structural analysis <strong>of</strong> (parallel, hydraulically driven) <strong>motion</strong> <strong>system</strong>s. The<br />

strengths <strong>of</strong> these models should, however, be proven by experimental validation. Proper<br />

choice <strong>of</strong> model parameter values is required in using the models to predict the characteristics<br />

<strong>of</strong> the actual <strong>system</strong> and enable direct comparison <strong>of</strong> model and experimental response.<br />

Usually, the parameter choice is also to be <strong>based</strong> on experiments. Experiments from<br />

which the parameters can be identified. Small differences between model and practice,<br />

i.e. a well validated model structure with proper parameter values is also important in the<br />

use <strong>of</strong> a model <strong>based</strong> <strong>motion</strong> <strong>control</strong>ler.<br />

Identification can also be performed with more general model structures which are especially<br />

suited for parameter identification. In nonlinear <strong>system</strong>s it is very hard to find<br />

convenient model structures. The physically motivated structure <strong>of</strong> the nonlinear models<br />

presented in this thesis, however, enables one to investigate what influence changes in the<br />

design <strong>of</strong> the <strong>system</strong> will have on its dynamics and helps in pointing at the relevant parts<br />

<strong>of</strong> the statics and dynamics in a complex <strong>system</strong> as the <strong>flight</strong> simulation <strong>motion</strong> <strong>system</strong>s at<br />

hand. This also helps in the design <strong>of</strong> proper experiments.<br />

As already discussed, the <strong>motion</strong> <strong>system</strong> (model) consists <strong>of</strong> kinematics and dynamics<br />

related to the mechanics i.e. the Stewart platform, and the hydraulic actuators which drive<br />

the <strong>system</strong>. Further some additional parasitic dynamics can be expected from a nonrigid<br />

foundation and a flexible platform/<strong>simulator</strong>. The identification <strong>of</strong> the dynamics specifically<br />

related to the hydraulic actuators has already been discussed in detail by Van Schothorst<br />

[124] and will not form an explicit part <strong>of</strong> this thesis though the hydraulic model structure<br />

presented does. Of course the dynamics resulting from the interaction between the<br />

hydraulics and mechanics will be, since the models presented in the previous chapter appointed<br />

this effect as the most relevant.<br />

This chapter will have to provide a good starting point for model <strong>based</strong> <strong>control</strong> on the<br />

actual <strong>system</strong>. The gap between the theoretically derived model structures in the previous<br />

119


120 4 Parameter identification and model validation<br />

chapter and the actual <strong>motion</strong> <strong>system</strong> to be <strong>control</strong>led will be bridged in three steps.<br />

Parameter estimation. The calibration method for the Stewart platform kinematical<br />

model will be presented, implemented and evaluated in Section 4.1. Next, in Section<br />

4.2, the experiments to infer the dynamic model parameters for the mechanics<br />

are set up, implemented and evaluated. Dynamic experiments to infer the position<br />

and orientation <strong>of</strong> measurement equipment consisting <strong>of</strong> accelerometers and rate gyros<br />

will be treated in Section 4.3.2.<br />

<strong>Model</strong> validation. The model characteristics with the parameters estimated will be<br />

validated in the frequency domain in Section 4.3. Especially the conclusion in the<br />

previous chapter that the hydromechanic rigid body modes are most relevant and that<br />

the <strong>system</strong> can be considered locally as six independent 1-d.o.f.-hydromechanic <strong>system</strong>s<br />

along the rigid body modal directions, will have to be evaluated experimentally.<br />

If this conclusion is justified in practise, the models, which describe this dynamics,<br />

are expected to provide a good starting point for model <strong>based</strong> <strong>control</strong>.<br />

Quantification <strong>of</strong> parasitic dynamics. The cause <strong>of</strong> the main parasitic dynamics<br />

observed in Section 4.3 will be clarified in the final part <strong>of</strong> this chapter, Section 4.3.2,<br />

by additional measurements showing vibration <strong>of</strong> the foundation and non rigid shuttle.<br />

Together with the hydraulically generated parasitic dynamics stemming from the<br />

valve and transmission lines, these parasitic dynamic effects are to be considered the<br />

limitations to the <strong>control</strong> scheme <strong>based</strong> on the basic model.<br />

All experiments discussed in this chapter are open loop <strong>based</strong> i.e. no feedback is involved<br />

apart from a very low bandwidth <strong>control</strong>ler, which is to stabilise the <strong>system</strong> without<br />

relevant influence on the information inferred from measurements for calibration/identification<br />

(higher frequencies or static relations).<br />

4.1 Calibration<br />

In this section the parameters <strong>of</strong> the kinematical model <strong>of</strong> the <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> will<br />

be identified (calibrated) and the kinematical model itself will be validated. In robot calibration<br />

[15], parameter identification <strong>of</strong> both the kinematics and dynamics <strong>of</strong> the mechanical<br />

<strong>system</strong> is referred to if calibration is considered. In this thesis, calibration will assume to be<br />

concerned with the identification <strong>of</strong> the kinematical model parameters or <strong>of</strong> specific measurement<br />

equipment. Parameter identification will generally refer to the model parameters.<br />

It is most convenient to perform calibration first as this step can be made independent <strong>of</strong> the<br />

<strong>system</strong> dynamics.<br />

Calibration is a very important topic in robotics. Especially if positional accuracy is<br />

relevant since this can <strong>of</strong>ten be improved by an order <strong>of</strong> a magnitude [15]. Further, one<br />

can expect that most experimental effort will have to be put into calibration as related to<br />

the time spend in testing the <strong>motion</strong> <strong>control</strong> strategies in practice [52]. In <strong>flight</strong> <strong>simulator</strong><br />

<strong>motion</strong> <strong>system</strong>s positional accuracy is clearly not the main issue but indirectly it affects the<br />

accelerational accuracy and is therefore considered in this thesis.


4.1 Calibration 121<br />

The basics <strong>of</strong> calibration can be found in textbooks such as [101] and books containing<br />

selections <strong>of</strong> articles published such as [15]. The framework adapted in this thesis is <strong>based</strong><br />

on an article <strong>of</strong> Hollerbach and Wampler [52]. In this article, a large set <strong>of</strong> robot calibration<br />

methods is unified considering joints and selecting closed kinematic loops. Further, a number<br />

<strong>of</strong> numerical issues is discussed, which provide a theoretical basis on the identifiability<br />

problems <strong>of</strong> kinematical parameters in practice. This method easily integrates the parallel<br />

driven structures but does not explicitly do so by presenting equations which are explicit<br />

in the end effector coordinates. The approach <strong>of</strong> Zhuang in [161] generalizes the idea <strong>of</strong><br />

making use <strong>of</strong> residuals in redundant equations for parallel structures in identifying the unknown<br />

parameters i.e. calibrating such <strong>system</strong>s and these ideas will also be incorporated in<br />

the calibration procedure.<br />

4.1.1 Calibrating the Stewart platform<br />

In this research no specific new theory will presented considering the calibration <strong>of</strong> a Stewart<br />

platform. The contribution will be a fully worked out applicable method consisting <strong>of</strong> a<br />

combination <strong>of</strong> known only slightly modified steps, which, not <strong>of</strong>ten seen in literature, is<br />

brought into practice and is put into a general framework. A large part <strong>of</strong> the procedure<br />

followed has been worked out as a part <strong>of</strong> two master projects [16] [151] in calibrating the<br />

SRS <strong>motion</strong> <strong>system</strong> with the dummy platform <strong>of</strong> Fig. 1.5. Here the results for the practical<br />

part <strong>of</strong> the calibration will involve the calibration <strong>of</strong> the <strong>motion</strong> <strong>system</strong> with the SRS-shuttle<br />

<strong>of</strong> Fig. 1.4.<br />

The calibration procedure proposed consists <strong>of</strong> only two basic steps.<br />

A redundant measurement procedure for platform pose reconstruction.<br />

Identification <strong>of</strong> the kinematical parameters <strong>of</strong> each platform leg separately using the<br />

redundancy <strong>of</strong> the actuator length measurement.<br />

Only a small part <strong>of</strong> the literature on calibration is devoted to the parallel manipulators.<br />

These <strong>system</strong>s differ from most serial <strong>system</strong>s considering the number <strong>of</strong> kinematical<br />

parameters involved, which is high. For the Stewart platform, a kinematical model,<br />

schematically given by Fig. 2.7, was presented by (2.38) from which<br />

klmi + loik 2 =(T ai +c ; bi) T (T ai +c ; bi) i =1�::: �6 (4.1)<br />

follows, which involves 42 unknown kinematical parameters. Seven for each equation consisting<br />

<strong>of</strong> an unknown <strong>of</strong>fset length, loi, which together with the measured length, lmi represents<br />

the total length <strong>of</strong> the i’th leg. The unknown upper and lower gimbal vectors, a i<br />

and bi, each consist <strong>of</strong> three parameters. These parameters will be identified in this section<br />

after redundantly measuring the platform pose consisting <strong>of</strong> the rotation matrix, T and the<br />

translation, c.<br />

This kinematical model assumes that all the gimbals are perfect ball and u-joints as<br />

schematically given by Fig. 2.8 and the prismatic joint axis <strong>of</strong> the sliding actuator exactly<br />

intersects the two gimbal points. In [154, 155] a more general kinematical structure is<br />

considered, allowing each axis <strong>of</strong> rotation in one gimbal to cross the other at a specific


122 4 Parameter identification and model validation<br />

distance not perfectly orthogonal. This model involves 132 unknown kinematic parameters,<br />

which would heavily complicate the calibration problem. Validation <strong>of</strong> the simple model<br />

after calibration can point out whether this is necessary.<br />

Reducing the number <strong>of</strong> unknown parameters to be identified in one step usually helps<br />

the conditioning <strong>of</strong> the problem to be solved and therefore improves the accuracy <strong>of</strong> the<br />

results. In [162] by Zhuang and Roth, the separate calibration <strong>of</strong> each leg <strong>of</strong> the Stewart<br />

platform is made possible by assuming a redundant measurement <strong>of</strong> the platform pose. This<br />

will leave one redundant actuator length variation measurement per actuator from which<br />

the seven unknown kinematical parameters in a i, bi and loi will have to be identified. In<br />

[162] also the calibration <strong>of</strong> the <strong>of</strong>fset length, l oi, is decoupled from the others, but as this<br />

strongly reduces the number <strong>of</strong> measurements, which can be taken into account, this part <strong>of</strong><br />

the procedure has not been adopted.<br />

The solution method to identify the kinematical parameters is iterative with only local<br />

convergence, like the forward kinematical problem discussed in Section 2.3. In [59, 60] a<br />

method is presented to calculate every solution to the calibration <strong>of</strong> a i and bi nonredundantly<br />

taking only six measurements into account and using a seventh measurement in deciding<br />

which solution is the right one. As it involves rootfinding <strong>of</strong> a twentieths order polynomial<br />

equation, this method is usually numerically not suitable for full (noisy) calibration in<br />

practice but it can be used for validation <strong>of</strong> the calibration procedure used here.<br />

The mobility <strong>of</strong> a <strong>system</strong>, roughly its degrees <strong>of</strong> freedom, can be specified with the<br />

mobility index, M as done by Hollerbach and Wampler [52]. It can be calculated given the<br />

number <strong>of</strong> links, L, including the base link and the number <strong>of</strong> constraints, D i, for each joint.<br />

Putting the method <strong>of</strong> Zhuang and Roth [162] for the Stewart platform actuator <strong>of</strong> Fig. 2.8,<br />

in the framework given in [52], the mobility index, M, <strong>of</strong> a separate leg <strong>of</strong> the platform is<br />

M =6(L ; 1) ; J i=1 Di = 3(4 ; 1) ; (4+5+6)=6 (4.2)<br />

The four links consist <strong>of</strong> base, lower and upper actuator part and the platform. There is<br />

a u-, prismatic and a ball joint. The calibration index, C, a redundancy measure given by<br />

Hollerbach and Wampler [52], can also be calculated as the total number <strong>of</strong> sensors, S, is<br />

seven, assuming an independent pose measurement, with<br />

C = S ; M =1: (4.3)<br />

This means there is only one redundant measurement involved from which all seven kinematical<br />

parameters have to be constructed. As in <strong>system</strong> identification <strong>of</strong> a dynamic miso<br />

<strong>system</strong> with several parameters, this is only possible if more than one, here at least seven,<br />

measurements are taken and if the inputs, the platform poses chosen, can be made sufficiently<br />

rich to make the parameters identifiable.<br />

As opposed to the discussion <strong>of</strong> serial manipulator examples given in [52] for <strong>system</strong>s<br />

with C =1, the number <strong>of</strong> measurements in the joints in a parallel <strong>system</strong> is one instead<br />

<strong>of</strong> the number <strong>of</strong> measurements for the end-effector. An advantage <strong>of</strong> C = 1 is that all<br />

equations involved and the residuals resulting, are <strong>of</strong> the same type. This is favourable in<br />

conditioning.<br />

In applying a calibration method in practice, the desired precision <strong>of</strong> the result should<br />

be known since as a rule <strong>of</strong> thumb the measurement equipment should be an order <strong>of</strong> a


4.1 Calibration 123<br />

−z<br />

−z<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

3<br />

2<br />

1<br />

0<br />

View in perspective<br />

0<br />

x<br />

Side view<br />

2 −2<br />

0<br />

−y<br />

−1<br />

−2 −1 0<br />

x<br />

1 2<br />

2<br />

−z<br />

−y<br />

3<br />

2<br />

1<br />

0<br />

Front view<br />

−1<br />

−2 −1 0<br />

−y<br />

1 2<br />

2<br />

1<br />

0<br />

−1<br />

Top view<br />

−2<br />

−2 −1 0<br />

x<br />

1 2<br />

Fig. 4.1: Two calibration measurement frames for the case <strong>of</strong> the Stewart platform. The<br />

fixed frame consists <strong>of</strong> three points. With the SRS they were chosen somewhere<br />

between the lower gimbals. The moving frame consists <strong>of</strong> three points chosen<br />

somewhere on the upper gimbal blocks. Knowledge on the exact positioning <strong>of</strong><br />

the measurement frame points is not necessary for this method.<br />

magnitude more precise [15]. For the measurement <strong>of</strong> the relative position <strong>of</strong> the prismatic<br />

joint a linear position transducer <strong>of</strong> the Temposonic type is used. The accuracy <strong>of</strong> these<br />

sensors is 0:1 mm (resolution with [55] 0:009 mm). A calibrated <strong>motion</strong> <strong>system</strong> with<br />

an positional accuracy <strong>of</strong> 1 mm will be strived for. The independent measurement <strong>of</strong> the<br />

platform pose discussed next will have to be done with equipment <strong>of</strong> :0:1 mm accuracy<br />

also.<br />

4.1.2 Redundant measurement <strong>of</strong> the platform pose<br />

The redundant measurement <strong>of</strong> the platform pose is <strong>based</strong> on the method presented by Geng<br />

and Haynes [43]. One defines two measurement frames e.g. as in Fig. 4.1. One fixed on<br />

the ground and one on the moving platform/<strong>simulator</strong> using three points for each frame.<br />

The <strong>motion</strong> <strong>of</strong> a point with respect to the fixed frame can be determined by measuring the<br />

distance <strong>of</strong> this point to the three points fixed to the ground if the distance between the<br />

fixed points is known and if the moving point remains in a known area w.r.t. plane spanned<br />

by the three points. With three additional length measurements (two to the second moving<br />

point and one to the third) the orientation can also be determined if the distance between the<br />

moving points is known.<br />

In case <strong>of</strong> the calibration <strong>of</strong> the shuttle <strong>of</strong> the SRS the distance between the moving<br />

points fixed to the shuttle are not precisely known and can not easily be measured directly<br />

since the shuttle shape does not allow so. This problem is solved by taking three distance<br />

measurements for each moving point (nine in total) from which the relative position <strong>of</strong> the<br />

moving points can be calculated. The relative distance <strong>of</strong> these points is redundant if more


124 4 Parameter identification and model validation<br />

Fig. 4.2: Determination <strong>of</strong> the position <strong>of</strong> a moving point in space using three length measurements.<br />

In case <strong>of</strong> the SRS, the three fixed points, Gi, can be thought in between<br />

the lower gimbal blocks. Each moving point, P i, was taken as part <strong>of</strong> the<br />

upper gimbal blocks.<br />

than one pose measurement is taken. This redundancy can and will be used to decide on<br />

a confidence level for the measurements in the weighted least squares calibration <strong>of</strong> the<br />

kinematical parameters. As only one distance to a specific point can be measured at the<br />

time this redundancy does not take significant additional effort to be measured since the<br />

platform has to be moved to a specific pose three times at least (using three pairs <strong>of</strong> three<br />

measurements instead <strong>of</strong> three, two and one measurement respectively).<br />

In Fig. 4.2 a top view <strong>of</strong> a three point distance measurement with three lengths, l i,<br />

i = 1�::: �3 is given to determine the position <strong>of</strong> a moving point P i with respect to the<br />

frame defined by the three points Gi, i =1�::: �3. Point, G1, is taken as the origin and the<br />

line through G1 and G2 as the y-axis. Using the cosine formula:<br />

l 2<br />

2 = l2<br />

1 + b2<br />

2 ; 2l1b2cos( )=l 2<br />

1 + b2<br />

2 ; 2b2y1� (4.4)<br />

y1 can be calculated. By letting the third fixed point, G 3, define the fixed measurement<br />

xy-plane, similar application <strong>of</strong> the cosine rule will lead to a parameter, s. This parameter,<br />

s, is the distance and direction <strong>of</strong> the origin to the point S(s x�sy� 0) in Fig. 4.2. So also the<br />

following holds<br />

s<br />

p 2 a3 + b2<br />

3<br />

= sx<br />

=<br />

a3<br />

sy<br />

b3<br />

(4.5)<br />

Further, the line in the xy-plane through S, orthogonal to the line through G 1 and G3, runs<br />

through (x1�y1� 0), i.e. direction @y=@x = ;a3=b3 can also be given by<br />

y1 ; sy<br />

= ;<br />

x1 ; sx<br />

a3<br />

b3<br />

Combining these equations to get rid <strong>of</strong> sx and sy leads to<br />

q<br />

2<br />

b3y1 + a3x1 = s a<br />

3<br />

: (4.6)<br />

+ b2<br />

3<br />

(4.7)


4.1 Calibration 125<br />

So in applying the cosine rule, s can be replaced by x 1 and y1, from which x1 can be solved<br />

l 2<br />

3 = l2<br />

1 + a2<br />

3 + b2<br />

3 ; 2l1<br />

=<br />

q<br />

2<br />

cos( ) a3 + b2<br />

3<br />

l 2<br />

q<br />

2<br />

1 + a2<br />

3 + b2<br />

3 ; 2s a3 + b2 3<br />

Finally j z1 j can be determined from<br />

= l 2<br />

1 + a2<br />

3 + b2<br />

3 ; 2(b3y1 + a3x1) (4.8)<br />

l 2<br />

1 = x21<br />

+ y2 1 + z2 1<br />

(4.9)<br />

The sign <strong>of</strong> z1, above or below the plane <strong>of</strong> measurement, should be known in advance. In<br />

this way the three vectors from the fixed frame to the moving points, p i, can be calculated.<br />

The frame defined through Gi given in Fig. 4.2 is convenient in determining the vectors to<br />

the moving reference points pi. In identifying the kinematical parameters it will be more<br />

suitable to define a slightly different fixed and moving frame from G i and Pi respectively,<br />

which are almost equal to the ground frame and moving frame defined for the SRS (Fig. 2.7).<br />

The new fixed reference measurement frame origin, G m, is given by a translation<br />

oGm =(g1 +g2 +g3)=3 (4.10)<br />

with respect to the frame given in Fig. 4.2. The orientation remains unchanged. The moving<br />

reference measurement frame origin, Pm, is specified w.r.t Gm accordingly with an origin<br />

(translation cpg)<br />

oPm =(p1 +p2 +p3)=3 ; oGm =cpg<br />

(4.11)<br />

and the orientation, with the line through P 1 and P2 chosen parallel to the y-axis, consists<br />

<strong>of</strong> three unit vectors along the new axes. With pij = pj ; pi, npy = p12= p p T 12 p12,<br />

npz = (npy p13)= p p T 13 p13 and npx = npy npz. These together form the rotation<br />

matrix from Pm to Gm<br />

TGm�Pm =[npx npy npz ]: (4.12)<br />

And this is the pose measurement we were looking for. In short it is given by (c pg, Tpg),<br />

which basically can be interpreted along the same lines as the pose defined for the Stewart<br />

platform.<br />

All together the used platform pose measurement requires nine absolute length measurements<br />

and the knowledge <strong>of</strong> the absolute distance between the fixed reference points.<br />

The quality <strong>of</strong> the calibration procedure heavily depends on these measurements. The positioning<br />

<strong>of</strong> the reference points themself is not that important. A rough idea will help in<br />

selecting a proper initial estimate in the calibration and the conditioning <strong>of</strong> the optimization<br />

problem to be solved is improved if all the distances to be measured (e.g. in Fig. 4.2) are<br />

about the same. Further the platform reference points have to reside at a predefined side <strong>of</strong><br />

the plane defined by the fixed reference points.


126 4 Parameter identification and model validation<br />

Fig. 4.3: In practise, the measurement device could not be used to take direct measurement<br />

<strong>of</strong> the length between two points at an angle ( 6= 0). A funnel (a cone shaped<br />

utensil) had to be placed in between. The picture shows that the measurement<br />

cable will not run directly from Gi to Pi but will follow the rounded edges <strong>of</strong> the<br />

funnel from Gi to D. This requires funnel length measurement compensation.<br />

4.1.3 Stewart platform pose measurement in practice<br />

The length measurements to reconstruct an independent, redundant platform pose in case <strong>of</strong><br />

the SRS-<strong>motion</strong> <strong>system</strong> have been performed with relatively low cost string encoders made<br />

by AMS Gmbh. A measurement cable winds onto an accurately machined cable drum. A<br />

coil spring delivers a constant pull in force to maintain cable tension. An incremental rotary<br />

encoder on the drum gives a digital output. The resolution <strong>of</strong> these encoders is 0.03 mm and<br />

the calibrated relative precision 0.1 mm. The standard version <strong>of</strong> this equipment does not<br />

allow the cable to be used at different angles w.r.t. the output point. This would damage the<br />

device.<br />

To allow calibration for an absolute length measurement, also for measurements taken<br />

at an angle, a funnel was designed. As shown in Fig. 4.3, this slightly affects the length<br />

measurement between Gi, which is chosen at the bottom <strong>of</strong> the funnel and P j. In practice<br />

this difference can take values up till 4 mm (R=15 mm) and should be corrected.<br />

Considering Fig. 4.3, lpg, the length between Pj and Gi is equal to<br />

p<br />

lpg = (X0 ) 2 + Z2 (4.13)<br />

the measured length will be equal to<br />

d1 + R<br />

p<br />

= (X0 ) 2 + Z2 ; 2RX0 + R (4.14)<br />

By using the measured lengths to reconstruct a first estimate <strong>of</strong> p i the required correction


4.1 Calibration 127<br />

can be approximated by taking<br />

0 = tan X 0<br />

Z<br />

(4.15)<br />

or more precise (in this case not necessary since R=lpg was small, only used for evaluation)<br />

reconstructing from<br />

Z tan( )=X 0 + R<br />

1 ; cos( )<br />

cos( )<br />

(4.16)<br />

It is assumed that the direction (Z) <strong>of</strong> the funnels (in the frame G m) is known.<br />

The choice for the approximate positions <strong>of</strong> the reference measurement points is given<br />

in Fig. 4.1. Considering Fig. 2.10, the fixed reference points (the encoders) have been<br />

placed in between the lower gimbal pairs (4,5), (2,3), (6,1) respectively. This allows the<br />

funnels to point along the z-axis <strong>of</strong> the measurement (and ground) frame. The upper, moving<br />

reference points have been taken in between (slightly lower: 80 mm) the same pairs <strong>of</strong><br />

upper gimbal points. In most platform poses, the lengths l pg will therefor not be the same<br />

(slightly less favourable conditioning) but in practice prevention from actuators running into<br />

the measurement cables is more important. As a result, the fixed ground frame and platform<br />

frame almost coincide with the measurement reference frames. This helps in interpretation<br />

<strong>of</strong> numerical results but does not improve the calibration itself.<br />

To allow measurement in between a favourable maximum and minimum length the limited<br />

(2 m) range string encoders were enlarged with steel cables (as used for fishing!) which<br />

were calibrated with the string encoders themselves. Two sets <strong>of</strong> these additional <strong>of</strong>fset cables<br />

were used (for Gi to Pj with i = j approx. 1.5 m and 2.0 m in case <strong>of</strong> i 6= j). The <strong>of</strong>fset<br />

<strong>of</strong> the string encoders themselves is also necessary to allow absolute measurement and was<br />

measured at a small, approx. 60 mm, retracted up right position with a sliding calipers to<br />

the funnels. The distance ( 1:8 m) between the string encoders could be measured with a<br />

large (2 m range) sliding calipers.<br />

Careful attention in performing these (<strong>of</strong>fset) measurements largely improved the calibration.<br />

By calibration before and after a measurement session also some drift compensation<br />

(1.4 mm for one string encoder) could be applied. Another important point <strong>of</strong> consideration<br />

was the dry cable friction in the funnels. Before each measurement, the cables had to be<br />

vibrated a little bit to get rid <strong>of</strong> play.<br />

To allow proper choice <strong>of</strong> the sensor range to be used, the set <strong>of</strong> platform poses, approximate<br />

since the actual kinematics is not known in advance, used for calibration can be<br />

simulated. In Fig. 4.4 all the positions <strong>of</strong> upper gimbal 1 are given for all 64 combinations<br />

<strong>of</strong> setting the six actuators in a minimum length position <strong>of</strong> 20 cm above minimal and below<br />

maximal stroke. These positions are sufficiently far away from the cushioning area <strong>of</strong> the<br />

actuators to be able to hold the <strong>system</strong> stable at these positions in taking a measurement.<br />

It is assumed that this set <strong>of</strong> positions will give relatively high variation in order to<br />

enhance the identifiability <strong>of</strong> the kinematical parameters. From the front view it can be observed<br />

that the positions are roughly the extremes <strong>of</strong> the 2-d.o.f. structure given in Fig. 3.22<br />

which are rotated if considered from the side view. The fact that some points are relatively


128 4 Parameter identification and model validation<br />

−z<br />

3<br />

2<br />

1<br />

0<br />

−2<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

View in perspective<br />

0<br />

Side view<br />

2 −2<br />

0<br />

−2 0 2 4<br />

x<br />

0<br />

2<br />

−z<br />

−y<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

Front view<br />

0<br />

−2 −1 0<br />

−y<br />

1 2<br />

2<br />

1<br />

0<br />

−1<br />

Top view<br />

−2<br />

−2 0 2 4<br />

x<br />

Fig. 4.4: Positions <strong>of</strong> upper gimbal 1 for calibration.<br />

close to each other in this picture does not imply that they will contain the same information<br />

since the orientation <strong>of</strong> the platform, which will be different, is not shown.<br />

If compared to the unconstraint actuator, it can be observed that especially the range <strong>of</strong><br />

angles to be reached by the platform actuators is relatively limited. E.g. the angles <strong>of</strong> the line<br />

between the lower gimbal point 1 and the dots in Fig. 4.4. This will affect the conditioning<br />

<strong>of</strong> the calibration as will be discussed in the next sections.<br />

4.1.4 Identification <strong>of</strong> the kinematical parameters<br />

After the determination <strong>of</strong> the platform poses, this information plus the actuator displacement<br />

measurements can be used to identify the kinematical parameters. Taking an initial<br />

estimate <strong>of</strong> these parameters in (4.1), each measurement regarding the j’th actuator, will<br />

have a residual. One can try to change the parameters in order to minimize these residuals.<br />

Using the quadratic form <strong>of</strong> (4.1), the jacobian <strong>of</strong> kinematical parameter variations, k, to<br />

residual variation, r, has a favourable linear structure<br />

Jrk = @r<br />

@k = Xo + 7<br />

i=1 Xiki<br />

with the kinematical parameters, k, <strong>of</strong> one actuator defined by<br />

and the i’th row (measurement) <strong>of</strong> the jacobian<br />

(4.17)<br />

k T =[a T b T loj] (4.18)<br />

Jrk(i� ) = 2[(lm(i) +lo(i)) (c T i Ti + b T Ti +a T ) (c T i +aT T T<br />

i + bT )]: (4.19)


4.1 Calibration 129<br />

It is immediately clear from the last equation that every row can be split in a non kinematical<br />

parameter dependent row<br />

Xo(i� )=2[lm(i) c T i Ti c T i ] (4.20)<br />

from which the matrix, Xo, can be constructed and parts which are linearly dependent on<br />

only one kinematical parameter e.g.<br />

X1(i� )k(1) = 2[0 1 0 0 T T<br />

i ( � 1)]a(1): (4.21)<br />

A weighted least squared solution to the linearised problem <strong>of</strong> minimising the residuals<br />

is given by<br />

kwls =(J T rk WJrk) ;1 JrkW r (4.22)<br />

In robot calibration, this is usually effective since the equations are only mildly nonlinear<br />

[52]. The weighting matrix, W , can be used to discriminate between the measurements.<br />

This can be appropriate since the confidence level for the different platform pose measurements<br />

can differ. If one <strong>of</strong> the string encoders showed e.g. a little hysteresis due to friction<br />

once in a while, the redundancy in the platform pose measurement can be used to detect<br />

this. With<br />

Wii =<br />

1<br />

0:5 + 1000 j lPP ; mlpp j<br />

(4.23)<br />

the confidence is upper bounded by 2 and goes to zero if high difference is detected between<br />

the measured distance, lpp, and the mean distance, mlpp, between two points, Pi.<br />

Iteration on (4.22) is similar to a Newton Raphson iteration. For convergence, it helps if<br />

the jacobian is Lipschitz, far from singular, well conditioned and starts with a well chosen<br />

initial estimate. Due to the linearity in the parameters, the jacobian is clearly Lipschitz.<br />

Further the structure <strong>of</strong> each parameter (Xi) is almost equal. The choice <strong>of</strong> platform poses<br />

can be used for conditioning. E.g. not varying the orientation, T , will result in a singular<br />

jacobian, since all solutions with only a translational vector added to the actual solutions <strong>of</strong><br />

both a and b will have equal residuals.<br />

The singular values <strong>of</strong> the jacobian, Jrk, are also called the observability indices <strong>of</strong> the<br />

calibration problem. In a comparative study cited by Hollerbach and Wampler in [52] it was<br />

concluded that optimal conditioning <strong>of</strong> these indices or maximising the minimal index gave<br />

good quantative results.<br />

The calibration problem <strong>of</strong> the actuator kinematics was idealised by an unconstrained<br />

simulation i.e. no constrains <strong>of</strong> platform poses to be chosen and all actual parameters zero.<br />

Choosing all (729) combinations <strong>of</strong> the tri state (-1,0,1) for translation and (-.5,0,.5) for the<br />

euler parameters results in a jacobian with condition number two at the solution, which is<br />

almost optimal. By Hollerback and Wampler [52] it is referred to a research concluding the<br />

condition number should be better (lower) than 100 to have reasonable results. With the<br />

SRS Stewart Platform to be calibrated with constrains on the actuator lengths the condition<br />

number increases to sixty using the 64-state set given in Fig. 4.4.


130 4 Parameter identification and model validation<br />

−z<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

−0.01<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

view in perspective<br />

0<br />

0.01<br />

side view<br />

0.01<br />

0<br />

−0.01<br />

−0.01<br />

−0.02 −0.01 0<br />

x<br />

0.01 0.02<br />

−z<br />

−y<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

front view<br />

−0.01<br />

−0.02 −0.01 0<br />

−y<br />

0.01 0.02<br />

0.02<br />

0.01<br />

0<br />

−0.01<br />

top view<br />

−0.02<br />

−0.02 −0.01 0<br />

x<br />

0.01 0.02<br />

Fig. 4.5: SRS calibration result with the design structure on a 1:100 scale and the differences<br />

1:1.<br />

e_length [mm]<br />

e_trans [mm]<br />

e_rot [meps]<br />

5<br />

0<br />

−5<br />

0 20 40 60<br />

5<br />

0<br />

−5<br />

0 20 40 60<br />

5<br />

0<br />

−5<br />

0 20 40 60<br />

non calibrated <strong>system</strong><br />

5<br />

0<br />

−5<br />

0 20 40 60<br />

5<br />

0<br />

−5<br />

0 20 40 60<br />

5<br />

0<br />

−5<br />

0 20 40 60<br />

calibrated <strong>system</strong><br />

Fig. 4.6: Positional errors for all 64 platform poses used for calibration made by the noncalibrated<br />

and the calibrated <strong>system</strong>. The first row shows length measurement<br />

prediction errors <strong>of</strong> the actuators length measurement given the redundantly measured<br />

pose. The next rows give the transformation <strong>of</strong> these errors to translational<br />

(2 nd row) and rotational (3 rd row) prediction errors <strong>of</strong> the platform pose.


4.1 Calibration 131<br />

4.1.5 Results in calibrating the SRS <strong>motion</strong> <strong>system</strong><br />

The calibration procedure leads to kinematical parameter (vectors) defined in the reference<br />

measurement frames. By construction <strong>of</strong> the forward kinematics <strong>of</strong> the newly found parameters<br />

for the actuators in half stroke, the transformation, sx gp, between the two measurement<br />

frames can be calculated. All upper gimbal positions can now also be referred to the ground<br />

measurement frame, Gm.<br />

a Gm =cgp + Tgpa Pm (4.24)<br />

As the parameters were not constrained to a plane, the <strong>simulator</strong> ground and moving<br />

frame are not trivially defined anymore. As for the reference measurement frames, the<br />

ground frame origin is chosen as the mean <strong>of</strong> the lower gimbal points. The ground frame xdirection<br />

is chosen as the addition <strong>of</strong> the vectors between the origin and gimbal point 1 and<br />

6 and the xy-plane is formed by making the vector from the origin to the mean <strong>of</strong> gimbals 2<br />

and 3 part <strong>of</strong> it. This defines the transformation to a ’design’ ground frame.<br />

Both the upper as the lower gimbals can now be referred to the design frames. In Fig. 4.5<br />

the gimbal points found through calibration are given in a design frame which is a 100<br />

times compressed making the differences, which are less than 0.1 %, visible. As opposed<br />

to the dummy platform [16] [151], no large deviations are found in the construction <strong>of</strong> the<br />

shuttle. It is now interesting to see what the differences are in predicting the actuator length<br />

measurements i.e. using the design values and the calibrated values. These deviations for<br />

all six actuators can also be transformed to the deviations in platform pose using<br />

sx = J ;1<br />

l�sx (sx) l (4.25)<br />

In Fig. 4.6 it is shown that the positional accuracy <strong>of</strong> the calibrated platform has almost<br />

improved an order <strong>of</strong> a magnitude.<br />

The accuracy <strong>of</strong> the noncalibrated <strong>system</strong> is not better than 5 mm in actuator lengths,<br />

2 mm in translation and 10 mrad in orientation. With the calibrated <strong>system</strong> this improves<br />

as shown in detail in Fig. 4.7 to .75 mm in actuator length predictability, apart from one<br />

outlier also predicted with upper reference point distance measurement. In the more important<br />

platform pose translational and orientational accuracy the errors are reduced to less<br />

than 0.5 mm and 2 mrad respectively.<br />

The initial estimate in the calibration procedure was more than another magnitude worse<br />

(> 100 times) since the pose <strong>of</strong> the measurement reference frame was only roughly (within<br />

100 mm) taken into account. Still the iteration on the kinematical parameters converged<br />

quickly within 4 steps for each actuator leg. Convergence is not guaranteed with such<br />

initial errors, but since the procedure can be done <strong>of</strong>f line, as opposed to the platform pose<br />

reconstruction treated in Section 2.2.11, this is not so important.<br />

One important aspect, the direction <strong>of</strong> the gravitational field has not been identified yet.<br />

This can be done using an additional measurement device (e.g. an inclinometer) as has been<br />

done in [16] [151]. But the pressure (force) measurements in the actuators can also be used<br />

to detect this as shown in the next section.


132 4 Parameter identification and model validation<br />

residual [mm]<br />

error [mm]<br />

Variation upper triangle measurement<br />

3<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

0 20 40 60<br />

0.5<br />

−0.5<br />

Translational pose prediction errors<br />

1<br />

0<br />

−1<br />

0 20 40<br />

pose no.<br />

60<br />

residuals in [mm]<br />

residual [meps]<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

Errors in actuator length prediction<br />

1<br />

−2<br />

0 20 40 60<br />

Orientational pose prediction errors<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

0 20 40<br />

pose no.<br />

60<br />

Fig. 4.7: Positional errors for all 64 platform poses used for calibration made by the calibrated<br />

<strong>system</strong> in detail. Also showing the variation in upper reference point distance<br />

measurement, which is used to calculate the confidence levels.<br />

Resume<br />

The following conclusions can be drawn w.r.t. the calibration procedure described.<br />

It has been shown that the kinematic mechanical parameters <strong>of</strong> a Stewart platform<br />

used as a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> can be calibrated resulting in a <strong>system</strong> with a<br />

positional accuracy better than 1 mm (for DEP). This is an order <strong>of</strong> a magnitude more<br />

precise than before calibration.<br />

The fact that the kinematical model itself is accurate within these error bounds for the<br />

SRS is another important conclusion.<br />

Further, it means that with calibration such <strong>motion</strong> <strong>system</strong>s can be made more accurate<br />

without putting more extreme tolerances on the fabrication and constructional<br />

process.<br />

The two step calibration procedure has been shown to work in practice.<br />

For the first step, measuring relative platform pose variations w.r.t. unknown, but<br />

platform and inertial world fixed, reference frames is sufficient redundant information<br />

necessary for calibration.<br />

The parallelism <strong>of</strong> the platform can be used in the second step to identify the 42<br />

kinematical parameters in six separate actuator groups. This enhances accuracy.<br />

It is important to realise that the result strongly depends on the accuracy <strong>of</strong> some<br />

<strong>of</strong> the measurements taken. These are the absolute length measurement <strong>of</strong>, and the<br />

distances in between, the redundant sensors.


4.2 Stewart platform model parameter identification 133<br />

As the calibration is concerned with sub mm precision on a sup m <strong>motion</strong> <strong>system</strong>, it is<br />

recommended that the procedure is rerun after each reconstruction <strong>of</strong> a platform e.g. in<br />

moving the SRS <strong>motion</strong> <strong>system</strong> to another site.<br />

4.2 Stewart platform model parameter identification<br />

Now the kinematic model has been validated and its parameters have been calibrated, the<br />

parameters concerning the dynamics <strong>of</strong> the <strong>motion</strong> <strong>system</strong> can be identified. The main<br />

unknowns in the Stewart platform model are its mass properties.<br />

First, the experiments to reconstruct the parameters related to the gravity forces, the<br />

location <strong>of</strong> the centre <strong>of</strong> mass and the weight itself will be discussed. Then, the identification<br />

<strong>of</strong> the inertial parameters is treated. These can be determined experimentally without much<br />

concern about the precise dynamics <strong>of</strong> the hydraulics. In this way, a reasonably accurate<br />

approximate model, i.e. (2.112) with a general mass matrix, <strong>of</strong> the mechanics presented in<br />

Chapter 2 taking into account the actuator intertia as in (2.122) can be identified.<br />

In the model <strong>based</strong> <strong>control</strong> structure to be presented in the next chapter, this part <strong>of</strong> the<br />

model will be important since the forces required for the accelerations to be applied for<br />

simulation will for most be generated by feed forward which requires an accurate model.<br />

Identification <strong>of</strong> the hydraulics, considering the separate actuators, has been discussed in<br />

[124]. The model parameters concerning the hydraulics are therefore assumed to be known.<br />

Validation <strong>of</strong> the dynamics resulting from the interaction in the hydraulically driven Stewart<br />

platform will be discussed in the next section.<br />

The set up <strong>of</strong> the hydraulically driven <strong>motion</strong> <strong>system</strong> supplied with the six valve inputs<br />

and measurement equipment to determine both pressure and positional information <strong>of</strong> the<br />

actuators, the legs, provides a complex but sufficiently rich environment by itself to perform<br />

almost all experiments required. The only additional measurement equipment which is<br />

used in this research are accelerometers and rate gyros which do more directly reflect the<br />

simulation performance on the platform and with which one can more accurately appoint<br />

the cause <strong>of</strong> parasitic dynamics resulting from flexible modes in the shuttle or foundation.<br />

4.2.1 Gravitational force determination<br />

The parameters to be identified, the static mass experienced and the position <strong>of</strong> the centre <strong>of</strong><br />

gravity, can be used to incorporate a gravity compensation in a <strong>control</strong>ler. This is important<br />

since gravity forces form a considerable part <strong>of</strong> the total force (up to 70 % <strong>of</strong> full pressure<br />

required as specified in Table 3.12). Further, a reasonable mass matrix results from taking<br />

the centre <strong>of</strong> gravity as the origin <strong>of</strong> the moving body axes in the dynamic model.<br />

Considering the model (2.122) in a static pose, leaves<br />

Gt =<br />

fg<br />

mg =(fg co) = J T l�x (sx)fa (4.26)<br />

with the gravity force, fg = Tgg[0 0 mg] T , transformed to the fixed lower gimbal frame by<br />

the rotation matrix, Tgg. Usually the matrix Tgg is approximately the identity matrix.


134 4 Parameter identification and model validation<br />

z<br />

f 1<br />

co<br />

f 2<br />

fg l1 l2<br />

c1<br />

f 3<br />

l<br />

3<br />

X<br />

Fig. 4.8: Simplified planar example <strong>of</strong> identifying mass and centre <strong>of</strong> gravity. Through the<br />

three actuator forces <strong>of</strong> a parallel 3-d.o.f. <strong>system</strong> the gravity vector (length and<br />

direction) plus line <strong>of</strong> action can be identified with two force equilibria and one<br />

moment equation. By rotating the <strong>system</strong> and identifying another (rotated) line<br />

<strong>of</strong> action, the intersection <strong>of</strong> both lines provides for the position <strong>of</strong> the centre <strong>of</strong><br />

gravity. In a spatial 6-d.o.f.-<strong>system</strong>, also two measurements are sufficient but the<br />

two lines <strong>of</strong> action can cross in practice. By identifying the point, which is closest<br />

to a set <strong>of</strong> lines <strong>of</strong> action, a more robust estimation can be attained.<br />

Further, the moments, mg, are induced by the vector product <strong>of</strong> this force with the vector<br />

<strong>of</strong> the upper centre gimbal point to the centre <strong>of</strong> gravity, c o. The vectors, fg, the gravity force<br />

i.e. platform mass times gravity and the centre <strong>of</strong> gravity, c o, have to be identified.<br />

From Fig. 4.8, it can be observed that one can only identify the part <strong>of</strong> c o, which is<br />

orthogonal to fg considering one pose as the moments result from a vector product i.e. a<br />

line <strong>of</strong> action can be identified on which the centre <strong>of</strong> gravity point resides. By measuring<br />

the forces required in other static poses in which the platform has been rotated (around a<br />

vector which is at least partly orthogonal to the gravitational direction), a set <strong>of</strong> lines can<br />

be determined from which co can be obtained. In theory all these lines will intersect at this<br />

point but in practice they will cross. The point which is closest to all lines in some sense<br />

can be taken as best guess for the centre <strong>of</strong> gravity. Here the least squares optimal solution<br />

<strong>of</strong> all (Euler) distances will be taken.<br />

Consider a specific static pose measurement <strong>of</strong> the pressure differences at the valves,<br />

[f T g mT T<br />

g ]=AdP i J(sx), assuming the input and output pressures are equal statically. A<br />

valid point <strong>of</strong> the line containing the centre <strong>of</strong> gravity, o, can be found by<br />

with ng = fg=(mg) and<br />

c1<br />

z<br />

l<br />

1<br />

f 1<br />

o = ; 1<br />

f T g fg<br />

fg Pngmg� (4.27)<br />

q<br />

f T<br />

g fg = mg. Now, the line <strong>of</strong> action can be described by<br />

o + ng. Every line, ( o + ng)i, found for a pose (ci�Ti), can be transformed back to the<br />

platform coordinates in one specific pose e.g. the neutral pose.<br />

( on + nngn)i = T T<br />

i ( o + ng) ; ci (4.28)<br />

f g<br />

co<br />

l<br />

2<br />

c2<br />

f 2<br />

f 3<br />

l<br />

3<br />

X


4.2 Stewart platform model parameter identification 135<br />

Configuration A B C<br />

m [tons] 2.60 2.10 4.25<br />

co(x) [m] 0.00 -0.12 0.00<br />

co(y) [m] 0.00 0.00 0.00<br />

co(z) [m] -0.23 -0.26 -0.45<br />

Table 4.1: The Simona Research Simulator gravitational parameters<br />

The vector, dlp, with smallest distance <strong>of</strong> some point, xp, to such a line is given by<br />

dlp = Pnng (xp ; on) (4.29)<br />

The vector, co, to the point, which minimizes the sum <strong>of</strong> squared distances to n <strong>of</strong> such<br />

lines, is given by<br />

co = 1<br />

n<br />

n<br />

i=0 Pnng�i<br />

;1 1<br />

n<br />

n<br />

i=0 Pnng�i oni (4.30)<br />

The conditioning <strong>of</strong> this optimization problem greatly depends on the ability to perform a<br />

measurement in which the summation <strong>of</strong> the projection matrices, P nng�i, equally span the<br />

3D-space i.e. whether the platform has been rotated sufficiently.<br />

Due to the limited stroke <strong>of</strong> the actuators <strong>of</strong> the Stewart platform, rotation can not be<br />

performed freely. In fact, performing valid pressure measurements can only be done if<br />

the actuators are out <strong>of</strong> the cushioning area (approx. 15 cm from max or min). Further,<br />

incorporating the mass <strong>of</strong> the actuators in the total mass is less tolerable with large rotations.<br />

With the SRS, the fast gravitational force identification procedure consisted <strong>of</strong> measuring<br />

pressures in five poses: neutral, +roll, -roll, +pitch and -pitch from neutral considering the<br />

limits: (lmin + :15,lmax ; :15).<br />

The more elaborate procedure takes into account all 64 poses used in the kinematic<br />

calibration. Over time, the mass and position <strong>of</strong> the centre <strong>of</strong> gravity <strong>of</strong> a <strong>simulator</strong> under<br />

construction, as the SRS, is likely to change. Therefore this procedure is to be repeated<br />

more <strong>of</strong>ten than full kinematic calibration.<br />

The conditioning <strong>of</strong> the matrix to be inverted in (4.30) is 12 and as a result the x- and<br />

y-position <strong>of</strong> the centre <strong>of</strong> gravity can be found approx. 10 times as accurate as the zposition.<br />

Fortunately, as this point is mainly used for gravitational compensation, the x- and<br />

y-position are also much more important.<br />

The gravitational forces <strong>of</strong> the three SRS <strong>system</strong> configurations evaluated in this research,<br />

A) The dummy platform alone,<br />

B) The shuttle and<br />

C) The dummy platform with additional load


136 4 Parameter identification and model validation<br />

have been identified. The results are given in Table 4.1.<br />

In the poses considered, the observed gravity forces are varying related to a mass <strong>of</strong><br />

approx. 20 kg due to the relative <strong>motion</strong> <strong>of</strong> actuators w.r.t. the platform. In all cases the<br />

c.o.g. is found above the centre gimbal plane. As the dummy platform has symmetricity, it<br />

is natural to find the c.o.g. in the origin <strong>of</strong> the xy-plane. The c.o.g. <strong>of</strong> the shuttle, which is<br />

somewhat to the back, is expected to move to the front if the visual <strong>system</strong> will be added.<br />

Measurements are best taken with a very slowly moving platform (e.g. sinusoidal <strong>motion</strong><br />

in z-direction < :5Hz, amplitude < 2mm over a whole number <strong>of</strong> periods) to eliminate<br />

most <strong>of</strong> the friction forces. Coulombic friction forces are small for the prismatic joint, the<br />

actuator with hydrostatically bearings, [124] but the rotation <strong>of</strong> the gimbal joint also has<br />

friction. These friction forces together are seen to be as large as 250 N.<br />

Both the calibration procedure as the gravitational force determination require a simple<br />

<strong>control</strong>ler to stabilize the <strong>system</strong> and moving the <strong>system</strong> to a specific platform pose and hold<br />

it there. The standard local (actuator per actuator) pressure feedback [153] in combination<br />

with positional PI-feedback with low bandwidth (1/10 <strong>of</strong> lowest eigenfrequency in neutral<br />

position i.e. approx. 0.5 Hz) will usually be sufficient.<br />

To determine the inertial properties <strong>of</strong> the <strong>system</strong>, discussed in the next section, experiments<br />

preferably require vibrations <strong>of</strong> higher frequency, e.g. approx. 4 Hz. This will still<br />

be attainable with a local <strong>control</strong> structure as no harsh requirements will be stated towards<br />

cross talk in the identifying procedure.<br />

4.2.2 Identification <strong>of</strong> the inertial properties<br />

Basic concept<br />

The main unknown parameters left to be determined are those contained in the mass matrix,<br />

the inertial properties <strong>of</strong> the <strong>system</strong>. In short, these parameters will be estimated by<br />

considering force and acceleration (through position) measurements and evaluating these<br />

signals in a domain where the basic equation force is equal to mass times acceleration in<br />

matrix form, f = Mx, holds. The mass matrix, M, can be deduced by performing six<br />

approximately sinusoidal vibrations which span the 6-d.o.f. coordinate space with sufficient<br />

acceleration (position) and force amplitude to be measured accurately and sufficiently small<br />

positional amplitude to have a <strong>system</strong> with only moderate nonlinear effects. It is assumed<br />

that no relevant parasitic flexible modes are hit at this harmonic.<br />

Basically from calculation <strong>of</strong> the platform forces fp = J T l�x (sx)AdP i and position<br />

(sx = f ;1 (l)), the amplitudes <strong>of</strong> the ground harmonic signal parts can be determined. The<br />

amplitudes <strong>of</strong> the accelerations, xa, resulting from the ground harmonic with frequency ! s<br />

are easily obtained from the positional as j xa j=j ! 2<br />

s xa j. Now, taking all experiments<br />

together in a matrix equation,<br />

Mt = FapX ;1<br />

a � (4.31)<br />

Mt can be identified, with for the i’th experiment M txai fapi and the force basic harmonic<br />

amplitudes, fap, stacked in the columns <strong>of</strong> Fap and stacking the accelerational amplitudes,<br />

xa, into Xa.


4.2 Stewart platform model parameter identification 137<br />

mpp (40kNm,kN)<br />

pitch (3eps,m)<br />

6<br />

4<br />

2<br />

0<br />

−2<br />

−4<br />

−6<br />

0.2<br />

0.1<br />

0<br />

−0.1<br />

x 10 −3<br />

0 0.5 1<br />

−0.2<br />

0 0.5<br />

time (s)<br />

1<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

−0.01<br />

−0.02<br />

x 10 −4<br />

0 0.5 1<br />

−0.03<br />

0 0.5<br />

time (s)<br />

1<br />

Fig. 4.9: Decomposition <strong>of</strong> platform forces and position into ground harmonic and residual<br />

signals in a vibrational (pitch) test <strong>of</strong> the dummy platform without additional load<br />

(configuration A.) for mass matrix determination. Upper left plot gives positional<br />

ground harmonics for the second euler parameter (pitch in three times the euler<br />

parameter, eps) and the ground harmonics for the other d.o.f.’s. Upper right are<br />

the residual positional measurements during the test. Lower left are the platform<br />

forces/moments ground harmonics, normalised to the maximum force <strong>of</strong> one actuator<br />

(times 1m for the moments). The lower right plot provides the measured<br />

residual forces and moments.


138 4 Parameter identification and model validation<br />

Estimation in the presence <strong>of</strong> nonlinear terms<br />

To evaluate under what conditions a reasonable estimate <strong>of</strong> the SRS <strong>motion</strong> <strong>system</strong> mass<br />

matrix, Mt, results, the nonlinear model will be considered. As proposed in the previous<br />

chapter, the mass matrix is assumed constant in the platform frame i.e. the relative <strong>motion</strong> <strong>of</strong><br />

the actuator parts is not taken into account. Referring to (2.122) and including a dissipating<br />

velocity term, Bt(sx) _x,<br />

Mt(sx)x + Ct(sx� _x) _x + Bt(sx) _x + Gt(sx) =J T l�x (sx)fa<br />

it is immediately clear that even in order to identify the mass matrix, M t, locally at some<br />

sx, dynamic experiments (x 6= 0) have to be performed in which nonlinear terms are<br />

unavoidable. Fortunately, as the term, Ct, is quadratic in the velocity it does for most<br />

have second and higher harmonic signal components since 2 sin( ) cos( ) = sin(2 ) and<br />

2cos 2 ( ) = (1 + cos(2 ))). So in convolution with the ground harmonic, C t almost disappears.<br />

Further, the ground harmonic function can be split up into two parts. One in phase<br />

with the position and one shifted over =2 rad, or the part which is in phase with velocity.<br />

By only considering the part <strong>of</strong> the signal (force, position) which is in phase with position,<br />

also the dissipating forces drop out.<br />

In Fig. 4.9 the necessary measurements are provided for one typical experiment out <strong>of</strong><br />

six in estimating the mass matrix. Only taking the ground harmonic into account allows<br />

accurate determination <strong>of</strong> the ground harmonic accelerational amplitude. As the ground<br />

harmonics along all degrees <strong>of</strong> freedom are taken into account, no severe requirements hold<br />

for the <strong>control</strong>ler during the experiment.<br />

Another cause left <strong>of</strong> calculating a biased estimate <strong>of</strong> the mass matrix will be the dry<br />

friction terms. Also these are velocity (sign) related and will also mainly influence the<br />

ground harmonic which is =2 out <strong>of</strong> phase.<br />

In practice, the <strong>system</strong> will not move in one platform coordinate only. Often the parasitic<br />

<strong>motion</strong> will also be out <strong>of</strong> phase. The <strong>system</strong> will therefore move along an ellipsoidal<br />

trajectory instead <strong>of</strong> a line as can be observed from Fig. 4.9, which shows phase shifted<br />

positional ground harmonics. This will be a possible cause <strong>of</strong> still having friction like terms<br />

in phase with the ground harmonic used in the estimation procedure.<br />

Another nonlinear term to be considered is the position dependent jacobian, J l�x(sx).<br />

Linearising in a static pose, sxo, with static gravitational actuator forces, fao results in<br />

Mt x = J T (sxo) fa + 6 @Jl�sx(sxo)<br />

i=0<br />

(sxo) sxifao<br />

(4.32)<br />

@sxi<br />

With sxi sufficiently small, these terms can be neglected. To check what is sufficiently<br />

small, the partial derivative <strong>of</strong> the jacobian, Jl�sx to the states sx is to be evaluated. As<br />

given in (2.42), this jacobian consists <strong>of</strong> two kind <strong>of</strong> vectors. All the unit direction vectors,<br />

lni <strong>of</strong> the actuators and the vector products, T a i lni. Using the time derivative <strong>of</strong> ln given<br />

in (2.52), and decomposition with (2.47) the partial derivative to sx is given by<br />

@ln 1<br />

=<br />

@sx klk PlnJai�x<br />

I 0<br />

0 J!<br />

(4.33)


4.2 Stewart platform model parameter identification 139<br />

with the transform from euler parameter variations to angular velocity given by J ! =<br />

2G( ) (2.23). Now<br />

@(T a ln)<br />

@sx<br />

= T a<br />

@ln @T<br />

+<br />

@sx @sx a ln (4.34)<br />

where @T=@sx follows from _<br />

T = ~!T. Validation is easily done by calculation <strong>of</strong> the<br />

difference between two jacobians for which the pose is slightly altered.<br />

Experiment design<br />

Calculation <strong>of</strong> each @Jl�sx=@sxi-matrix in the neutral position gives matrices with roughly<br />

the same gains as Jl�sx itself. With the gravity forces at the same magnitude as the dynamics<br />

forces, the relative effect <strong>of</strong> the variation (to compensate for gravity) in the linearisation<br />

is nearly the same as the ratio <strong>of</strong> positional over accelerational amplitudes. As the test<br />

frequency becomes higher, this ratio decreases quadratically.<br />

In practice the test frequency <strong>of</strong> the ground harmonic, ! t is chosen somewhat below<br />

the lowest open loop ’rigid’ eigenfrequency. In this manner, it possible to successfully<br />

perform the test with a simple non model <strong>based</strong> <strong>control</strong>ler. In case <strong>of</strong> configuration A.<br />

and B., !t = 2 4 and for the dummy with additional load it was taken somewhat lower<br />

!t =2 =0:3. At these frequencies it is possible to perform the measurements with reasonable<br />

pressure (up till 25 % <strong>of</strong> full load) and velocity (up till 20 % <strong>of</strong> maximum velocity)<br />

amplitudes by requiring harmonic reference signals for each platform coordinate in a separate<br />

test <strong>of</strong> positional amplitude <strong>of</strong> 10mm in the translational and :003 for k 13k in<br />

rotational directions. Also an equivalent ratio <strong>of</strong> measurement errors in position (0.1 mm)<br />

and pressure ( 25 mV noise on 10V scale) in relation to the amplitudes is attained in this<br />

way (both 1%). Typical measurement signals, for the forces, S f , and positions, Sp, for such<br />

a test are given in Fig. 4.9. The actual test length was tl =5s and the measurements were<br />

sampled at 500 Hz.<br />

Determination <strong>of</strong> the ground harmonics<br />

The harmonic coefficients <strong>of</strong> the positional and force (pressure) measurements contain the<br />

information <strong>of</strong> the mass properties. They are calculated by using a matrix, H, with a basis<br />

function in each column over the time span 0�::: �tl sampled at the same frequency<br />

H =[1t ; :5tl sin(!tt) cos(!tt) sin(2!tt) cos(2!tt) :::] (4.35)<br />

Now the coefficient matrices for the platform forces C f and positions Cp are the least square<br />

estimates found by<br />

[Cf Cp] =(H T H) ;1 H T [Sf Sp] (4.36)<br />

For the ground harmonic (third, c 3, and fourth, c4 rows <strong>of</strong> the coefficient matrices), the<br />

phase, p, <strong>of</strong> the platform coordinate, which is tested, is calculated. The columns fap and<br />

xap <strong>of</strong> the matrices Fap and X <strong>of</strong> (4.31) are now found by<br />

[f T ap x T<br />

ap ]=real(e;j p ([c3f + jc4f c3p + jc4p])) (4.37)


140 4 Parameter identification and model validation<br />

Configuration Place in 6x6 A B C<br />

Mass matrix Mt<br />

mx [tons] (1,1) 2.6 2.0 4.3<br />

my [tons] (2,2) 2.6 2.0 4.3<br />

mz [tons] (3,3) 2.7 2.3 4.3<br />

Ixx [tons m 2 ] (4,4) 2.3 2.6 4.1<br />

Iyy [tons m 2 ] (5,5) 2.3 2.8 4.0<br />

Izz [tons m 2 ] (6,6) 3.8 3.5 6.7<br />

Ixz�zx [tons m 2 ] (4,5),(5,4) 0.0 0.3 0.0<br />

M15�51 [tons m] (1,5),(5,1) 0.0 0.0 0.1<br />

Table 4.2: The Simona Research Simulator identified nonzero inertial parameters. These<br />

are mainly appearing on the diagonal <strong>of</strong> the mass matrix, M t, apart from<br />

the Ixz�zx term in case <strong>of</strong> the shuttle (configuration B.). At least 1 % error<br />

(25 ; 50 kg) can be expected since this is the maximum singular value <strong>of</strong> the<br />

nonsymmetric part <strong>of</strong> the estimated mass matrix.<br />

In this way, a maximal part <strong>of</strong> the ground harmonic for each test direction is taken into<br />

account and only the part <strong>of</strong> the force and positional signal which are in phase with this<br />

<strong>motion</strong>.<br />

In Fig. 4.9 it is shown that a significant part <strong>of</strong> the forcing signal does not belong to the<br />

ground harmonic and will not be taken into account in the mass matrix estimation.<br />

Experimental results <strong>of</strong> the mass matrix estimation<br />

The estimated mass matrices for the three configurations are given in Table 4.2. An advantage<br />

<strong>of</strong> taking the identified centre <strong>of</strong> gravity as origin in the moving body frame is that the<br />

mass matrix becomes almost diagonal. The dummy platform has the advantage <strong>of</strong> having<br />

its main axes <strong>of</strong> inertia along the platform frame axes. For the shuttle, configuration B, the<br />

xz-plane is a plane <strong>of</strong> symmetry (making the y-axis a main axis <strong>of</strong> inertia, running I yz�zy<br />

and Ixy�yx zero) but also has a cross term Ixz�zx <strong>of</strong> approx. 250 kg in case <strong>of</strong> an empty<br />

shuttle.<br />

In principle, the mass matrix should be symmetric. The nonsymmetricity <strong>of</strong> the identified<br />

mass matrix can give some impression <strong>of</strong> the errors made due to measurement inaccuracy,<br />

etc. In case <strong>of</strong> configuration A., the maximum singular value <strong>of</strong> the non symmetric<br />

part is 26 kg(m 2 ) and in B. and C. it amounts to 50 kg i.e. 1 % <strong>of</strong> the estimate itself for the<br />

dummy platform and somewhat higher for the shuttle. This means no more accurate results<br />

were obtained than specified in Table 4.2.<br />

Resume<br />

A very compact test sequence was presented, which allows estimation <strong>of</strong> the relevant<br />

dynamic mechanical parameters <strong>of</strong> the Stewart platform in the presence <strong>of</strong> hydraulic


4.3 Frequency response model validation 141<br />

actuators.<br />

In practice five static poses (at least two in theory) measuring pressure and position,<br />

allowed identification <strong>of</strong> the total mass experienced and the position <strong>of</strong> the centre <strong>of</strong><br />

gravity.<br />

With six persistently exciting periodic <strong>motion</strong> tests, the inertial properties <strong>of</strong> the <strong>motion</strong><br />

<strong>system</strong> could be identified through a (6 by 6) mass matrix.<br />

By only taking the ground harmonic in phase with positional <strong>motion</strong>, velocity related<br />

(friction) and higher order terms (nonlinearities) have minimal influence on the<br />

procedure.<br />

Together with the earlier identified hydraulic actuator parameters, these mechanical<br />

parameters allow a reasonably tight fit to practice <strong>of</strong> the complex multivariable nonlinear<br />

hydraulically driven mechanical <strong>system</strong> model.<br />

In the next section, this will be validated in comparing the responses <strong>of</strong> the models derived<br />

with the parameters determined with ’open loop’ <strong>motion</strong> <strong>system</strong> frequency response<br />

measurements.<br />

4.3 Frequency response model validation<br />

At this point, the model structures have been formulated and the model parameters have<br />

been estimated. With this information, the calibrated models can be evaluated by comparison<br />

<strong>of</strong> model and actual responses. In this section the model open loop <strong>system</strong> characteristics<br />

will be validated in the frequency domain. Especially with mechanical <strong>system</strong>s, also<br />

hydraulically driven, many important properties are revealed considering responses to sinusoidal<br />

inputs.<br />

Characterising nonlinear <strong>system</strong>s by describing functions<br />

With frequency response measurements not only the input-output relation <strong>of</strong> a linear <strong>system</strong><br />

can uniquely be described but also many nonlinear <strong>system</strong> structures can be characterized<br />

by performing harmonic excitation. E.g. a sinusoidal input describing function identification<br />

procedure approximately characterises a nonlinear <strong>system</strong> by considering amplitude<br />

and phase <strong>of</strong> the response as a function <strong>of</strong> frequency and input sinusoidal amplitude, as<br />

extensively discussed in Van Schothorst [124]. This only requires a filter hypothesis, which<br />

says that the non-linear <strong>system</strong> has low-pass characteristics.<br />

Further, the linear dynamics <strong>of</strong> a non-linear <strong>system</strong> (for some operating point) can be<br />

defined as the describing function <strong>of</strong> the <strong>system</strong> as the amplitude approaches zero i.e. is<br />

sufficiently small. The describing function can be found measuring a frequency response<br />

such that the amplitude <strong>of</strong> the response is flat. This requires the design <strong>of</strong> an input amplitude<br />

filter found e.g. by iteratively measuring the frequency response.<br />

Appropriate experiments characterising hydraulically driven mechanical <strong>system</strong>s<br />

Systems experiencing structural vibrations such as mechanical <strong>system</strong>s are <strong>of</strong>ten charac-


142 4 Parameter identification and model validation<br />

terised experimentally using modal analysis techniques. A large part <strong>of</strong> the modal analysis<br />

also heavily relies on measuring frequency responses [39, 90]. With modal analysis, the velocity<br />

response <strong>of</strong> a mechanical <strong>system</strong> to force excitation is called mobility. With a single<br />

degree <strong>of</strong> freedom (one mode, vibration) e.g. a mass-spring-damper <strong>system</strong>, the mobility<br />

frequency response has a zero at zero and at infinite frequency and a complex pole pair<br />

assuming reasonably low damping with an undamped eigenfrequency at the square root <strong>of</strong><br />

the mass vs. spring stiffness ratio.<br />

Similar responses can be expected measuring the response <strong>of</strong> our <strong>motion</strong> <strong>system</strong> from<br />

input valve voltage to the pressure difference over the valve. As discussed in the previous<br />

chapter, the hydraulically driven mechanical <strong>system</strong> has roughly a mobility like characteristic<br />

if the response <strong>of</strong> the pressure difference over the actuator compartments is considered<br />

from input actuator valve flow. Measuring from input valve voltage sets a low pass filter<br />

in series with this <strong>system</strong> thereby fulfilling the filter hypothesis. Viscous damping not only<br />

damps the resonance but also shifts the zero at zero frequency to the left in the complex<br />

plane. As shown in Section 3.3.1 considering the models, the response <strong>of</strong> the multi degree<br />

<strong>of</strong> freedom multi input valve/output pressure difference <strong>of</strong> the <strong>motion</strong> <strong>system</strong> at hand is expected<br />

to have a mobility like characteristic with multiple modes. Also with parasitic effects<br />

from transmission lines or mechanical flexibility, this does not change, apart from inducing<br />

additional anti-resonance and resonance pairs.<br />

4.3.1 Frequency response dummy platform<br />

The models <strong>of</strong> the previous chapter appointed the rigid body modes as the main characteristics<br />

<strong>of</strong> the hydraulically driven mechanical <strong>system</strong>. In this section, this will be evaluated by<br />

the actual frequency responses <strong>of</strong> the dummy platform (configuration A. with no additional<br />

load). In this configuration relatively little influence <strong>of</strong> parasitic effects, e.g. stemming from<br />

flexible deformation, is expected to enter the frequency area <strong>of</strong> interest i.e. the frequency<br />

area up till 50 Hz where the rigid body modes reside.<br />

With the inertial parameters found in the previous section and the parameters for the hydraulic<br />

actuators given in Table 3.2, the theoretic basic model structure <strong>of</strong> the hydraulically<br />

driven <strong>motion</strong> <strong>system</strong> given in Section 3.4.1 is specified. Calculation <strong>of</strong> the eigenfrequencies<br />

<strong>of</strong> the rigid body modes in the neutral position for the dummy platform, configuration A.,<br />

results in 6:8 Hz for a surge/pitch and sway/roll mode, 13:6 Hz for a yaw mode, 17:7 Hz<br />

for heave and 23:4 Hz for the different sign surge/pitch and sway/roll modes.<br />

With a HP frequency analyzer ([91]) the relation at the neutral pose from each valve<br />

steering voltage to each actuator pressure difference was measured between 5 Hz and<br />

50 Hz by a swept sine at 1:5% <strong>of</strong> full input voltage. With the analyzer 201 logarithmically<br />

spaced datapoints per input output relation were obtained. As this analyzer only resolves<br />

one input output relation at the time, 36 separate experiments were performed. Most <strong>of</strong> the<br />

frequency domain measurements are given in 6x6 form in Appendix A.<br />

The <strong>system</strong>s pose was stabilized by a moderate position feedback <strong>of</strong> gain k p = 0:2<br />

resulting in a bandwidth <strong>of</strong> 0:05 Hz. At 5 Hz the influence <strong>of</strong> this feedback can be<br />

1% and grows at the first resonance to maximally 2%. At the frequencies larger than<br />

10 Hz it will be as little as 0:1% and decrease quadratically. This influence is estimated


4.3 Frequency response model validation 143<br />

Amplitude<br />

Phase (deg)<br />

10 1<br />

10 0<br />

0<br />

−200<br />

−400<br />

−600<br />

10 1<br />

10 1<br />

Frequency (Hz)<br />

10 1<br />

10 0<br />

0<br />

−200<br />

−400<br />

−600<br />

10 1<br />

10 1<br />

Frequency (Hz)<br />

Fig. 4.10: Bode plots <strong>of</strong> the highly interacting responses from the first valve input voltage<br />

to the six pressure input differences (dPi), measured at the dummy platform<br />

(configuration A.) at the left. On the right side the model response (18 th order)<br />

is given incorporating the rigid body modes and three additional modes for the<br />

foundation. Six by six separated Bode plots can be found in Fig. A.1 and Fig. A.2<br />

<strong>of</strong> Appendix A<br />

by taking the measured ’closed loop’ gain, k m, and recalling the open loop response gain,<br />

kg, to be measured is found through kg = km(1 ; kckm) -1 with kc, the position loop gain<br />

including the transfer function from (normalized) pressure to position. For small (complex)<br />

values, kckm is the relative gain variation.<br />

j kckm j = kpAdpnormkm(m! 2 ) ;1<br />

= 0:2 25 10 -4 200 10 5 7=(8:8 10 3 (2 5) 2 )<br />

= 0:008<br />

(4.38)<br />

The measured Bode plot <strong>of</strong> the measurement from the first valve input to the six pressure<br />

differences is given in the left column <strong>of</strong> Fig. 4.10. The predicted rigid body modes are<br />

present and additionally some parasitic resonances are present between 10 ; 13 Hz and at


144 4 Parameter identification and model validation<br />

Direction Mfii [tons� tons m 2 ] !fi [Hz] fi<br />

xf 45 2 10:5 0:05<br />

yf 45 2 10:0 0:035<br />

f 140 2 12:5 0:05<br />

Table 4.3: Parameters taken for the foundation.<br />

approx. 30 Hz.<br />

Additional measurements at the floor, which will be discussed in the next section,<br />

pointed out that the vibrations at 10 ; 13 Hz could be assigned to the foundation. By modelling<br />

the foundation as a 3-d.o.f mass/spring/damper <strong>system</strong> as modelled in Section 3.4.5<br />

with the parameters <strong>of</strong> Table 4.3, an 18 th -order model results.<br />

In general the damping structure <strong>of</strong> mechanical <strong>system</strong>s is hard to predict theoretically.<br />

The damping <strong>of</strong> the first three rigid body modes is well described by the experimentally<br />

found values for leakage and viscous friction <strong>of</strong> the hydraulic actuators. For the three highest<br />

frequency modes some additional damping had to be introduced.<br />

This can be explained by the fact that the gimbal rotation is probably also dissipating<br />

energy and not explicitly taken into account. Further, it is observed in modal analysis that<br />

damping does not always act as the linear viscous effect, which is part <strong>of</strong> the model used<br />

here. Alternatively, hysteretic damping structures are proposed in [39, 90], in which the<br />

frequency dependence <strong>of</strong> the dissipation is changed. A drawback <strong>of</strong> that structure is that it<br />

can not be incorporated in a state space (time-domain) model. Therefore hysteretic damping<br />

has not been included in the models proposed in this thesis for <strong>motion</strong> <strong>system</strong>s.<br />

The Bode plot from the first valve input to the six pressure differences <strong>of</strong> the (6 6)<br />

18 th -order model is given in the right column <strong>of</strong> Fig. 4.10. The measured transfer functions<br />

are fully shown in Fig. A.1 and Fig. A.2 <strong>of</strong> Appendix A. The model gives a reasonably<br />

close description <strong>of</strong> the characteristics observed from the measurements. Both the phase as<br />

amplitude frequency responses are very much alike. Note that only the phase <strong>of</strong> the first<br />

input to the first output remains within 90 necessary for passivity.<br />

The largest discrepancy can be observed at 30 Hz where the measurements point out<br />

an additional resonance, which is not incorporated in the model. Limited stiffness <strong>of</strong> the<br />

actuators in the radial direction are a possible cause <strong>of</strong> this vibration but this could not be<br />

confirmed by some additional measurements <strong>of</strong> an accelerometer waxed at the actuators.<br />

Further, the additional phase lag coming in at frequencies 40 Hz can be assigned to<br />

the valve characteristic. The model including this effect will be evaluated with the shuttle<br />

(configuration B.). The frequency responses <strong>of</strong> that <strong>system</strong> were measured up till 500 Hz.<br />

Decoupled response by coordinate transformation<br />

In the frequency responses <strong>of</strong> Fig. 4.10 is it not immediately clear how many modes are to be<br />

used to the describe the 6 6 <strong>system</strong>. Now we can use the basic model structure we obtained<br />

for hydraulically driven mechanical <strong>system</strong>s in Section 3.4.1. An important <strong>system</strong> property<br />

<strong>of</strong> the model is the coordinate transformation along the rigid body modal directions, which


4.3 Frequency response model validation 145<br />

Amplitude<br />

Phase (deg)<br />

10 1<br />

10 0<br />

100<br />

50<br />

0<br />

−50<br />

−100<br />

10 1<br />

10 1<br />

Frequency (Hz)<br />

10 1<br />

10 0<br />

100<br />

50<br />

0<br />

−50<br />

−100<br />

10 1<br />

10 1<br />

Frequency (Hz)<br />

Fig. 4.11: Bode plots along the six estimated decoupling rigid body modal directions (without<br />

parasitic vibrations) from redirecting the input valve voltages and valve pressure<br />

differences measured (left) and modelled (right). In the upper left, the<br />

largest singular value over the frequency <strong>of</strong> the error <strong>system</strong> regarding all the<br />

nondiagonal elements is given in the dashed plot. Six by six separated Bode<br />

plots can be found in Fig. A.3 and Fig. A.4 <strong>of</strong> Appendix A<br />

decouples the <strong>system</strong> into six independent <strong>system</strong>s. Each <strong>of</strong> which incorporates one <strong>of</strong> the<br />

rigid body modes.<br />

With additional parasitic modes this decoupling is not exact anymore but in the previous<br />

chapter it was found that at least in the neutral position, this transformation still puts the<br />

main modes on the diagonal. This transformation with an unitary matrix, U, can be derived<br />

from the mass matrix evaluated from the actuators as defined in (3.44). A quick impression<br />

for these modes can attained by evaluating the singular values for each separate frequency<br />

measured.<br />

In Fig. 4.11, the frequency responses <strong>of</strong> the diagonal elements <strong>of</strong> the 6 6 transformed<br />

transfer function are plotted. (Full transfer functions are shown in Fig. A.3 and Fig. A.4<br />

<strong>of</strong> Appendix A. In the left column, the measured response is given and on the right, the<br />

model response is drawn. Still, the model and the measured response have very much the<br />

same characteristics. Having the rigid body modes decoupled on the diagonal <strong>of</strong> the transfer<br />

function enables a much easier interpretation. E.g. the phases <strong>of</strong> the <strong>system</strong> are now clearly<br />

90 from which the conclusion can be drawn that the linearised <strong>system</strong> is passive at this<br />

point <strong>of</strong> operation as was predicted from the non-linear <strong>system</strong>. As predicted, the dynamics


146 4 Parameter identification and model validation<br />

<strong>of</strong> the foundation and other additional mechanical resonances do not influence this property.<br />

Of course, the valve characteristics will make the <strong>system</strong> lose its property <strong>of</strong> being strict<br />

positive real and cause the additional phase shift at the higher frequencies <strong>of</strong> the measured<br />

response.<br />

In this decoupled form, the two frequency responses, the measured response and the<br />

one derived from the model, can be compared in more detail. Further, it is easier to adjust<br />

the model in this form in order to let the model more tightly fit reality. For example the<br />

damping structure is most conveniently changed in these coordinates since each rigid mode<br />

can be varied independently. Also parameters like massmodes (eigenvalues massmatrix)<br />

can be adjusted. In the model, the mass viewed along the z-direction can be changed to<br />

a 13% higher value to have higher correspondence <strong>of</strong> the modelled and measured z-mode.<br />

Probably the actuator masses are somewhat higher than expected.<br />

Important is the fact that the non diagonal part <strong>of</strong> the responses is almost neglectable<br />

as the largest singular value <strong>of</strong> this part <strong>of</strong> the transfer is much smaller than the rigid body<br />

modes, which appear separated on the diagonal transfer functions. In the measured responses<br />

it is clear that there is some distortion at the resonance frequencies coming into the<br />

low gain parts <strong>of</strong> the other modes.<br />

Resuming, the model which describes<br />

the rigid body mechanical <strong>system</strong><br />

together with the basic hydraulic <strong>system</strong>,<br />

and in this environment some additional dynamics from the non rigid foundation,<br />

gives a reasonable description <strong>of</strong> the <strong>system</strong> over the frequency range, which will be important<br />

for <strong>control</strong> (5 ; 50 Hz). Both the rigid modal eigenfrequencies and directions can<br />

be predicted from the model structure together with the kinematic and inertial model parameters<br />

identified. The damping structure had to be adapted somewhat using the measured<br />

frequency responses. Something which is not unusual since the damping structure <strong>of</strong> a<br />

mechanical <strong>system</strong> is hard to predict.<br />

Frequency domain identification<br />

A more tightly fitted local description <strong>of</strong> the <strong>system</strong> at the neutral pose can be obtained by<br />

frequency domain identification. A flexible tool was used to identify a multivariable model<br />

from frequency domain data with the method <strong>of</strong> De Callafon et al [28] as already reported<br />

in [75]. No reasonable results could be obtained, however, from directly fitting the original<br />

6 by 6 by 201 datapoints. Results improved drastically by separately identifying the rows <strong>of</strong><br />

the data transformed by the coordinate transformation, U. Rows instead <strong>of</strong> columns were<br />

taken as the singular directions could be predicted more closely at the directly measured<br />

output than at the input through the valve.<br />

As these models will usually contain redundant modes from non exact decoupling, especially<br />

at the dominant lowest rigid modes at 6:8 Hz, the full multivariable model can be<br />

reduced. By using the block balancing approach <strong>of</strong> Wortelboer [157], the reduction problem<br />

can be transformed to finding a good multivariable representation <strong>of</strong> each observed mode.


4.3 Frequency response model validation 147<br />

Amplitude<br />

10 0<br />

10 −1<br />

10 1<br />

Input filter shuttle frequency response measurements<br />

Frequency (Hz)<br />

Fig. 4.12: Bode amplitude plot <strong>of</strong> the 6 th -order experiment notch filter.<br />

The resulting 20 th -order model describes both the six rigid modes as four additional<br />

parasitic modes, including the modes at 30 Hz which were not incorporated in the more<br />

global model with physical interpretation. The largest singular value <strong>of</strong> the difference between<br />

model and data at each frequency at which the data was measured is equivalent with<br />

the dashed line in the upper left <strong>of</strong> Fig. 4.11. Thus the error remains well below the relevant<br />

dynamics.<br />

With the inverse coordinate transformation, U T , this model can be transformed to the<br />

actuator coordinates from which the original measurements were taken. The error is approximately<br />

10%.<br />

More research into identification <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> has to point out<br />

in what way models or model parameters can be identified which provide a tight description<br />

<strong>of</strong> the <strong>system</strong> on a global scale, e.g. over all platform positions.<br />

4.3.2 Additional dynamics into the higher frequency area<br />

With the shuttle on top <strong>of</strong> the <strong>motion</strong> <strong>system</strong>, Configuration B., it was decided to measure<br />

the frequency responses over the wide frequency area between 5 Hz and 500 Hz again<br />

with a sine (up) sweep <strong>of</strong> 201 logarithmically spaced frequencies. In this way the possible<br />

flexible modes <strong>of</strong> the shuttle and the influence <strong>of</strong> the transmission lines on the pressure<br />

dynamics, could be observed and evaluated with the <strong>system</strong> characteristics predicted by the<br />

models obtained in the previous chapter.<br />

In performing the experiments, more care w.r.t. the dummy platform measurements had<br />

to be taken not to hit resonances to badly. Especially because the pilot seats, experiencing<br />

some play in the mounting, appeared to have their resonant behaviour close to the main<br />

rigid modes <strong>of</strong> the shuttle. Moreover, the high frequency area, containing the transmission<br />

line dynamics, did not allow, already for the audible noise level alone, input voltage levels<br />

necessary to measure the low frequecy rigid mode area with sufficient signal/noise ratio.<br />

A 6 th -order experiment filter was designed to obtain a sufficiently flat response level,<br />

which is also favourable in spreading the relative effect <strong>of</strong> possible nonlinearities. The measured<br />

bode amplitude plot <strong>of</strong> this filter is drawn in Fig. 4.12. Having the poles <strong>of</strong> the filter<br />

10 2


148 4 Parameter identification and model validation<br />

!p1 !p2 !p3 !n1 !n2<br />

2 5 2 35 2 100 2 7:8 2 42<br />

p1 p2 p3 n1 n2<br />

0:7 0:7 0:7 0:25 0:25<br />

Table 4.4: The parameters <strong>of</strong> the input frequency response measurement filter.<br />

appear just before the lightly damped (notching) zeros on 7:8Hz and 42Hz, gradually reduces<br />

the amplitude level and above 100Hz filters more rigorously. The filter was designed<br />

in the continuous time domain and converted into a digital version with a sample frequency<br />

<strong>of</strong> 5kHz by a bilinear transformation (in Matlab) for implementation on the <strong>motion</strong> <strong>control</strong><br />

computer. By measuring the response <strong>of</strong> the filter implemented on the computer, the <strong>system</strong><br />

frequency responses can be adequately compensated for. The input level <strong>of</strong> the experiments<br />

could be set on 2 % but the filter already starts to reduce this level at 6Hz.<br />

The continuous version <strong>of</strong> the experiment filter, U f (s) is given by<br />

Uf (s) = (s2 =! 2 n1 +2 n1s=!n1 +1)<br />

(s 2 =! 2<br />

p1 +2 p1s=!p1 +1)<br />

(s2 =! 2<br />

n2 +2 n2s=!n2 +1)<br />

(s2 =! 2<br />

p2 +2 p2s=!p2 + 1)(s2 =! 2<br />

p3 +2 p3s=!p3 +1)<br />

(4.39)<br />

with the parameters specified in Table 4.4.<br />

The direct measurements <strong>of</strong> the <strong>system</strong> with the shuttle from the actuator inputs to the<br />

input pressure differences show an even higher number <strong>of</strong> resonances than those with the<br />

dummy platform. However, the structure <strong>of</strong> the <strong>system</strong> characteristics is revealed by decomposition<br />

along the eigen directions <strong>of</strong> the rigid modes predicted through the (rigid body)<br />

model <strong>of</strong> the shuttle using the mass properties identified by the experiments discussed in the<br />

previous section. This requires pre and post multiplication <strong>of</strong> the measured 6 by 6 transfer<br />

function by a constant unitary matrix Ush�n and its transpose.<br />

Also with the shuttle and over the larger frequency area, this transformation puts the<br />

dominant responses on the diagonal, which show only one rigid mode per direction, while<br />

all 36 directly measured responses contain every mode in principal. This result and the<br />

model predicted frequency responses are given in Fig. 4.13 and in Fig. A.7 and Fig. A.8<br />

<strong>of</strong> appendix A all plots are given seperately, while Fig. A.5 and Fig. A.6 <strong>of</strong> Appendix A<br />

provide the Bode plots for the untransformed frequency data.<br />

Due to the nondiagonal inertia matrix, there are more platform directions involved in<br />

each mode considering the neutral pose than in case <strong>of</strong> the dummy platform. Still, in this<br />

pose, one can discriminate between symmetric (x� z� -related) and non-symmetric (y� � -<br />

related) modes. Premultiplying the Ush�n-matrix, whose columns describe actuator displacements,<br />

by the inverse jacobian results in a column wise description <strong>of</strong> the modal direction<br />

in the platform coordinates. Rescaling the columns to a (2-)norm <strong>of</strong> one, by postmulti-


4.3 Frequency response model validation 149<br />

Amplitude<br />

Phase (deg)<br />

10 1<br />

10 0<br />

10 −1<br />

100<br />

0<br />

−100<br />

−200<br />

−300<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

Frequency (Hz)<br />

10 1<br />

10 0<br />

10 −1<br />

100<br />

0<br />

−100<br />

−200<br />

−300<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

Frequency (Hz)<br />

Fig. 4.13: Bode plots <strong>of</strong> the shuttle response redirected along the six estimated decoupling<br />

rigid body modal directions considering the valve input voltages and the valve<br />

pressure difference responses measured (left) and modelled (right). Six by six<br />

separated Bode plots can be found in Fig. A.7 and Fig. A.8 <strong>of</strong> Appendix A<br />

plying with Ncsc gives:<br />

J ;1<br />

lx Ush�nNcsc =<br />

2<br />

6<br />

4<br />

0:987 0:000 0:000 0:083 ;0:078 0:000<br />

0:000 0:988 ;0:032 0:000 0:000 ;0:064<br />

;0:026 0:000 0:000 0:828 0:676 0:000<br />

0:000 0:153 0:089 0:000 0:000 0:988<br />

;0:159 0:000 0:000 0:554 ;0:733 0:000<br />

0:000 ;0:022 0:996 0:000 0:000 ;0:139<br />

3<br />

7<br />

5<br />

(4.40)<br />

The eigenfrequencies, which belong to these modal directions, are 7:3, 7:3, 14:7, 20:3, 22:0,<br />

24:2 Hz. The (60 th -order) model predicts these frequencies (and its directions) appropriately<br />

as are the transmission lines anti-resonances at approximately 70Hz and the characteristic<br />

due to the (anti)-resonances at 200Hz. This does not include the exact behaviour at<br />

the resonant transmission line frequencies since only the actuator cylinder pressure difference<br />

and not the absolute pressures were taken into account as was discussed in the previous<br />

chapter.


150 4 Parameter identification and model validation<br />

The transfer function from valves flow to input pressure differences is passive despite<br />

the transmission lines, flexible shuttle modes and moving foundation. In Fig. 4.13 this<br />

clearly shows by the phase moving between =2 rad. The valve voltage to valve flow<br />

characteristic with two poles modelled at 150Hz results in the additional phase shift, which<br />

(together with the transmission line resonances and flexible shuttle modes) can result in<br />

stability problems when using pressure feedback.<br />

Another property, which clearly shows, is that the hydraulic actuators are velocity generators<br />

at low frequencies below the rigid body resonance. The Bode plots show that this<br />

requires different pressure levels since the mass experienced along the modal directions<br />

varies. Above the resonant frequencies, the actuators become pressure difference derivative<br />

generators and in this respect the fact that all actuators are similar causes indifference for<br />

direction considering the amplitudes. A dual effect will show in considering the velocity<br />

(accelerating) behaviour over the frequency using constant valve input amplitude.<br />

The characteristics <strong>of</strong> the foundation having resonant frequencies at 10:2, 10:7 and<br />

12:7Hz do show the same characteristics in the model and measurements. The model parameters<br />

and the evaluation <strong>of</strong> the foundation causing these resonances were obtained by<br />

performing measurements at the foundation itself, which will be discussed in the next part<br />

<strong>of</strong> this section.<br />

The main discrepancy between the model and the actual <strong>system</strong> is the omission <strong>of</strong> the<br />

flexible modes which show at 33Hz, again, as with the dummy platform probably due to<br />

the radial stiffness <strong>of</strong> the actuators. Further, new resonances show up most prominently at<br />

43� 57 and 65Hz and are still visible around 100Hz. These resonances were most probably<br />

caused by the flexibility <strong>of</strong> the shuttle. To confirm this, accelerometers measurements were<br />

performed at several points <strong>of</strong> the shuttle.<br />

Although the characteristics <strong>of</strong> the pressure measurements in Fig. 4.13 are for most<br />

well contained in the model, the model does not predict the measurements very accurately<br />

considering the numerical values. This probably requires a more thorough identification<br />

procedure and/or more research into the nonlinear behaviour <strong>of</strong> the <strong>system</strong>. The Bode response<br />

is <strong>based</strong> on a linearised version <strong>of</strong> the model. E.g. the nonlinear valve characteristics<br />

could be a cause <strong>of</strong> some <strong>of</strong> the discrepancies.<br />

The vibrating foundation<br />

As already shown in several Bode plots, the vibrating foundation is part <strong>of</strong> the model according<br />

to the equations presented in Section 2.4.7. In principle, the parasitic resonance,<br />

which shows up in e.g. Fig. 4.13, can be caused by any flexibly attached mass at either<br />

side <strong>of</strong> the platform. In experiments and demonstrations it was readily verifiable that also<br />

the floor was moving if the <strong>motion</strong> <strong>system</strong> was made to vibrate at frequencies approaching<br />

10Hz.<br />

To confirm that it was actually the concrete foundation, which is approx. 20 times as<br />

heavy as the shuttle <strong>of</strong> Configuration B., which causes the specific behaviour, and to help<br />

identifying the model parameters and structure, experiments were set up measuring the<br />

acceleration at the floor itself. Creating the experimental environment to perform the measurements<br />

for closer analysis <strong>of</strong> parasitic mechanical behaviour <strong>of</strong> the <strong>system</strong> (so also those


4.3 Frequency response model validation 151<br />

10 0<br />

10 −1<br />

10 −2<br />

10 0<br />

10 −1<br />

10 −2<br />

10 1<br />

10 1<br />

Fig. 4.14: Bode plot amplitude responses <strong>of</strong> measured (dashed) and modelled accelerations<br />

along the xf (upper left), yf (upper right), x2f at (0,-2) (lower right) and constructed<br />

angular acceleration around the z-axis, ! (lower left) due to input valve<br />

vibration along the first three ’rigid modes’.<br />

100<br />

0<br />

−100<br />

−200<br />

−300<br />

100<br />

0<br />

−100<br />

−200<br />

−300<br />

10 1<br />

10 1<br />

Fig. 4.15: Bode plot phase responses <strong>of</strong> measured (dashed) and modelled accelerations<br />

along the x (upper left), y (upper right), x 2 at (0,2) (lower right) and constructed<br />

angular acceleration around the z-axis, ! (lower left) due to input valve vibration<br />

along the first three ’rigid modes’.<br />

10 0<br />

10 −1<br />

10 −2<br />

10 0<br />

10 −1<br />

10 −2<br />

100<br />

0<br />

−100<br />

−200<br />

−300<br />

100<br />

0<br />

−100<br />

−200<br />

−300<br />

10 1<br />

10 1<br />

10 1<br />

10 1


152 4 Parameter identification and model validation<br />

in the shuttle itself) have been part <strong>of</strong> a masters’ project described in [41].<br />

Use was made <strong>of</strong> the (at that time) new compact accelerometers <strong>of</strong> Analog Devices, the<br />

ADXL05 1 , which evolved from airbag <strong>system</strong> initiators and now measure accelerations at a<br />

range <strong>of</strong> 5g at a resolution <strong>of</strong> 5mg from DC into reasonable high frequencies (10 kHz).<br />

Three <strong>of</strong> these sensors were mounted orthogonally in a small package, such that acceleration<br />

measurement in all three directions, xf , yf and zf was possible.<br />

This package was mounted on the floor in between the <strong>motion</strong> <strong>system</strong> on three places<br />

performing 6(x6) experiments in total assuming that the floor itself was rigid but flexibly<br />

attached to the inertial frame. Right in between the lower gimbal points as the frame given<br />

in Fig. 2.13, all three accelerations were measured, again while performing a sine sweep<br />

at the valve input voltage for each actuator at the time (between 5Hz and 500Hz usually<br />

stopping the experiment if no <strong>motion</strong> could be observed anymore above 50Hz). Again, the<br />

<strong>motion</strong> <strong>system</strong> provides a nice 6-d.o.f. force inductor to perform the required experiment 2 .<br />

Recording both accelerations and the forces induced is required in identifying the mass<br />

and spring stiffness, which can hardly be identified if only the resonant behaviour in the<br />

e.g. pressure dynamics is taken into account.<br />

At the coordinates (0� -2)[m] and (2� 0)[m] taken in the frame <strong>of</strong> Fig. 2.13, measurements<br />

were performed in x and z and z direction respectively. All measurements in z-direction<br />

did not show any acceleration (apart from gravitation), allowing to take only three directions<br />

(xf ,yf, f ) into account.<br />

In Fig. 4.14 and Fig. 4.15, the bode amplitude and phase plots <strong>of</strong> the model and measurements<br />

in the xf and yf direction are given as well as for the x2f and reconstructed f<br />

direction due to the inputs along the rigid modal directions most concerned with the <strong>motion</strong><br />

to be observed. Clearly, in Fig. 4.14 and Fig. 4.15 both the rigid shuttle modes can be observed<br />

along with the mode due to the fact that the foundation is not rigidly attached to the<br />

floor. In the neutral state, there is no interaction between the rigid shuttle modes due to the<br />

floor but in other platform poses it can.<br />

Although the accelerometers were relatively noisy, the characteristics observed in the<br />

models can be recognized in the measurements. The model fails to identify a secondary<br />

mode in x-direction (at approx 12 Hz) which is hardly affecting the yaw mode (as it does<br />

with the model). The most relevant behaviour is, however, well described by the model. No<br />

additional effort has been done to more accurately describe the dynamics <strong>of</strong> the foundation<br />

as the <strong>system</strong> had to be moved to another building in the end.<br />

The parameters used in the model <strong>of</strong> the foundation are given in Table 4.5. The total <strong>system</strong><br />

eigenvalues, which are due to this part <strong>of</strong> the mechanics, are shifted somewhat w.r.t. the<br />

values given here for the foundation alone. The direction <strong>of</strong> this shift naturally depends on<br />

the eigenfrequencies <strong>of</strong> the rigid body modes <strong>of</strong> the platform (up in case <strong>of</strong> x and y, down<br />

in case <strong>of</strong> ).<br />

The experiments point at the importance <strong>of</strong> taking into account the dynamic behaviour<br />

<strong>of</strong> the construction <strong>of</strong> the building, foundation, best already in design, in order to prevent<br />

1This sensor is not available anymore and has been replaced by the ADXL105 having higher resolution and<br />

temperature compensation<br />

2As already put forward, the hydraulic actuators are no force generators by themselves. Probably, the experiments<br />

could be enhanced by using some sort <strong>of</strong> force <strong>control</strong> loop.


4.3 Frequency response model validation 153<br />

Phase (deg)<br />

Amplitude (m/s2/10V)<br />

Direction Mfii [tons� tons m 2 ] !fi [Hz] fi<br />

xf 45 2 10:5 0:035<br />

10 2<br />

10 1<br />

10 0<br />

0<br />

−200<br />

−400<br />

−600<br />

yf 45 2 10:0 0:03<br />

f 182 2 12:5 0:07<br />

Table 4.5: Parameters taken for the foundation.<br />

10 1<br />

10 1<br />

Frequency (Hz)<br />

Fig. 4.16: A typical frequency response with a Bode plot <strong>of</strong> measured and reconstructed<br />

(from the other measurements) acceleration approx. along the y-axis in the back<br />

left floor position due to a sinusoidal sweep <strong>of</strong> the input voltage at the 4 th actuator.<br />

this part <strong>of</strong> the <strong>system</strong> influence the characteristics <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> in a<br />

relevant frequency area as it did with the experiments performed at the Central Workshop<br />

<strong>of</strong> mechanical engineering.<br />

Accelerating a flexible shuttle<br />

At this point only measurements were presented, which indirectly provide insight in the behaviour<br />

<strong>of</strong> acceleration in the shuttle. By mounting a (minimal) number (<strong>of</strong> six) accelerometers<br />

onto the shuttle, the rigid body <strong>motion</strong>, felt by a person sitting in the <strong>simulator</strong>, can be<br />

observed. Additional sensors can provide insight into the parasitic flexibility resulting from<br />

the shuttle itself.<br />

10 2<br />

10 2


154 4 Parameter identification and model validation<br />

10 2<br />

10 1<br />

10 0<br />

10 2<br />

10 1<br />

10 0<br />

10 2<br />

10 1<br />

10 0<br />

10 1<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

10 2<br />

Frequency (Hz)<br />

10 2<br />

10 1<br />

10 0<br />

10 2<br />

10 1<br />

10 0<br />

10 2<br />

10 1<br />

10 0<br />

10 1<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

10 2<br />

Frequency (Hz)<br />

Fig. 4.17: Bode amplitude plots <strong>of</strong> the platform acceleration response due to the valve inputs<br />

along the rigid body modal input directions. Left column x� y� z, and on the<br />

right _!x� _!y� _!z. For each plot one <strong>of</strong> the modal direction is not dashed. It is<br />

clear that strong interaction and hard to analyze responses result from evaluation<br />

along the platform directions. Much stronger decoupling will result evaluating<br />

along the rigid body modal directions at the output also.<br />

Reconstruction <strong>of</strong> platform acceleration<br />

To reconstruct the platform acceleration, first the gain, direction and positional parameters<br />

<strong>of</strong> the accelerometers used, have to be identified. By using a method in which the procedure<br />

in identifying the platform mass matrix and centre <strong>of</strong> gravity are combined, the sensors can<br />

be characterised. In general, an accelerometer with gain k, mounted at position, r, w.r.t. a<br />

reference frame and directed along the unit vector, n, will output a signal, s, determined by<br />

s = kn T (c ; g + _! r +! (! r)) (4.41)<br />

where c and _! provide the reference frame translational and rotational acceleration.<br />

By performing sinusoidal <strong>motion</strong>, the accelerational amplitudes, ( c� _!)i, in six directions,<br />

which span the <strong>motion</strong> space, can robustly be derived from the position measurements.<br />

By correlating with the basic sinusoidal frequency, ! t, noise and the quadratic


4.3 Frequency response model validation 155<br />

velocity term will be very small as in (4.36) and will be neglected. (4.41) also holds for<br />

the ground harmonic amplitudes. With some manipulation using (4.41), the accelerometer<br />

ground harmonic output amplitudes <strong>of</strong> all the experiments, s, can be written as<br />

s = Am� (4.42)<br />

the multiplication <strong>of</strong> a known matrix, A, <strong>of</strong> platform accelerational ground harmonic amplitudes<br />

<strong>of</strong> which each i’th row is given by<br />

h<br />

Ai� = aT c�i aT i<br />

_!�i � (4.43)<br />

and the measurement device vector, m, with the yet unknown accelerometer parameters,<br />

direction, n, gain, k, and line <strong>of</strong> action, r, given by<br />

h<br />

m = knT (kn ( g<br />

! 2 t ; r))T i T<br />

(4.44)<br />

Gravity gives an <strong>of</strong>fset to the line <strong>of</strong> action <strong>of</strong> the sensor which is inversely proportional with<br />

the squared frequency, ! 2 t , at which the test has been performed. This <strong>of</strong>fset results from<br />

the fact that gravity is constant in the inertial frame and the direction <strong>of</strong> the accelerometers,<br />

n, moves (and thus rotates) with the platform. If e.g. pitch rotation occurs, gravity will<br />

show at the accelerometer measuring in x-direction proportionally to the sine <strong>of</strong> the angle<br />

rotated. For small angles the sine operation can be dropped and the angle decreases with<br />

the squared frequency with constant rotational acceleration. By inversion <strong>of</strong> A, the gain, k,<br />

and direction, n, <strong>of</strong> the accelerometers can be determined.<br />

m = A ;1 s (4.45)<br />

The position, r, can not, just as in the determination <strong>of</strong> the centre <strong>of</strong> gravity. Using only<br />

the linear terms (not the quadratic velocity term), the sensor can be anywhere on a line<br />

parametrized by the free variable , described by<br />

with<br />

r( )= o + n (4.46)<br />

1<br />

o = ;<br />

mT 11m11 (m11 Pnm21) (4.47)<br />

where [m T 11 mT 21 ]T = m T .<br />

With at least two accelerometers, not measuring in parallel directions, in a measurement<br />

device, the position <strong>of</strong> this device can be defined by the point closest to both lines identified<br />

as r( ). As in the identification <strong>of</strong> the centre <strong>of</strong> gravity (4.30), incorporation <strong>of</strong> more<br />

measurements (accelerometers) can easily be done. In case <strong>of</strong> the shuttle measurements, a<br />

device with three accelerometers was attached in several positions, p o, as given in Fig. 4.19.<br />

For each position, the accelerometer lines <strong>of</strong> measurement were identified and from these<br />

lines po.


156 4 Parameter identification and model validation<br />

Amplitude (160Bar/10V)<br />

Phase (deg)<br />

10 1<br />

10 0<br />

0<br />

−100<br />

−200<br />

−300<br />

−400<br />

10 1<br />

10 1<br />

Frequency (Hz)<br />

Fig. 4.18: Bode plot <strong>of</strong> the diagonal elements <strong>of</strong> the input valve voltage to output pressure<br />

transfer function along the rigid body modal directions reconstructed from the<br />

accelerometer measurements. As in Fig. 4.13 the <strong>system</strong> is strongly decoupled<br />

and each plot shows only one rigid body mode. The transfer function to output<br />

(actuator) pressure difference lacks the transmission zeros at approx. 70Hz and<br />

shows phase shift much earlier in comparison with input pressure difference as<br />

output plotted in Fig. 4.13.<br />

Three positions were chosen at the shuttle floor, the part <strong>of</strong> the shuttle which was expected<br />

to be the most rigid (highest flexible eigenfrequencies). Just behind the pilot’s chairs<br />

in the middle, in the back at port and starboard. At the front side, the ’chin’ <strong>of</strong> the shuttle,<br />

just below the place where the window is to be placed, in the middle and at the starboard<br />

side, the redundant measurements were taken. The identified positions, using 4Hz sinusoidal<br />

<strong>motion</strong>, are given in Table 4.6. These positions give some idea <strong>of</strong> the placing <strong>of</strong> the<br />

accelerometers. In further calculation also the relative positioning <strong>of</strong> the measurement lines,<br />

r( ), and their direction should be taken into account.<br />

Assuming a rigid body shuttle floor, the rigid body frequency response can be identified<br />

by measuring at least six accelerometer responses at the floor. To include some redundancy<br />

for evaluation, seven sensors were used in measuring the response to the sine sweep <strong>of</strong> all<br />

the actuators independently. x, y and z for the mid and y and z for the (port and star) back<br />

10 2<br />

10 2


4.3 Frequency response model validation 157<br />

Position 1:mid 2:port back 3:star back 4:front mid 5:front star<br />

x [m] -0.44 -1.33 -1.32 1.21 1.11<br />

y [m] -0.18 -0.40 0.43 0.07 0.89<br />

z [m] -0.11 -0.11 -0.12 -0.39 -0.39<br />

Table 4.6: Positions identified in the shuttle for the accelerometer device relative to the<br />

centre <strong>of</strong> gravity in platform coordinates.<br />

4<br />

5<br />

x<br />

y<br />

Fig. 4.19: Approximate positions <strong>of</strong> the accelerometer device in the shuttle xy plane as also<br />

given by Table 4.6.<br />

positions. Port y is used for evaluation. The other six are used to construct the platform<br />

response by<br />

1<br />

3<br />

2<br />

x = M ;1<br />

p s (4.48)<br />

where each row <strong>of</strong> Mp is given by m T <strong>of</strong> (4.45). With the platform response, x, it is possible<br />

to reconstruct all the (linear) parts <strong>of</strong> the acceleration resulting from the rigid body <strong>motion</strong><br />

at all positions <strong>of</strong> the shuttle.<br />

Reconstructing rigid body acceleration, pressure and velocity dynamics using accelerometers.<br />

Sine sweeps from 5Hz up to 500Hz with 201 logarithmically spaced frequency points were<br />

made with 150mV amplitude send through the experiment filter <strong>of</strong> Fig. 4.12. All measurements<br />

presented have been compensated for this filter.<br />

Reconstruction <strong>of</strong> rigid body redundant accelerometer measurement.<br />

In Fig. 4.16, a typical accelerometer frequency response to a sinusoidal valve voltage input<br />

is given. It is the response <strong>of</strong> the accelerometer measuring in y-direction at the port back<br />

position. In dashed form the reconstructed response is plotted.<br />

Clearly visible are the rigid body resonance at 8Hz (sway/roll), 15Hz (yaw) and 24Hz<br />

(roll/sway). At 10Hz the resonance due to the foundation can be observed. Modes, which<br />

are not identified yet, are visible at 35� 57 and 80Hz. Irrespective <strong>of</strong> the high number <strong>of</strong><br />

resonances, the two responses correspond quite well. Apart from the valve phase drop, the


158 4 Parameter identification and model validation<br />

Amplitude ("m"/s/10V)<br />

10 1<br />

10 0<br />

10 −1<br />

10 1<br />

Frequency (Hz)<br />

Fig. 4.20: Bode amplitude plot along the rigid body modal directions <strong>of</strong> valve input to the<br />

velocities along these directions reconstructed by the accelerometer measurements.<br />

Again these plots are decoupled showing only one rigid mode per plot<br />

and further the convergence to the typical hydraulic actuator static velocity gain<br />

going to frequency zero shows.<br />

transfer function phase start to drop below =2 around 57Hz and further at 80Hz possibly<br />

due to flexibility between the input (actuator) and the sensor.<br />

Reconstructed dynamics <strong>of</strong> input valve voltage to platform accelerations.<br />

Step by step the responses can be manipulated such that a structured response results which<br />

can be analyzed more easily. In a first step, Fig. 4.17 provides the responses <strong>of</strong> the platform<br />

pose accelerations to the inputs along the rigid body directions. Almost all platform<br />

directions (apart from the yaw-direction) are influenced by more than one rigid body mode.<br />

In the neutral pose, where the measurements where taken, it is possible to discriminate between<br />

symmetric and non-symmetric modes (x� z� ) and (y� � ).<br />

Output pressure dynamics.<br />

More fully decoupled responses (up till the flexible resonances) can be achieved by considering<br />

the reconstructed output pressure differences or velocities along the rigid body<br />

directions. The output pressures are reconstructed using a constant premultiplication given<br />

by<br />

dP<br />

^<br />

o = U T A ;1 J ;T<br />

l�x Mx (4.49)<br />

containing the unitary modal direction matrix, U, the actuator area matrix, A, the jacobian,<br />

Jl�x, and the identified mass matrix, M.<br />

10 2


4.3 Frequency response model validation 159<br />

actuator 1 2 3 4 5 6<br />

yaw 1 -1 1 -1 1 -1<br />

chin up 1 0 0 0 0 1<br />

torsional x-axis 1 -1 -1 1 1 -1<br />

Table 4.7: Input directions flexible mode measurements.<br />

In Fig. 4.18 it is shown that all the rigid modes have been decoupled in this way. It<br />

looks like the six accelerometer measurements allow a reconstruction <strong>of</strong> the (linear part<br />

<strong>of</strong> the) output pressure difference characteristics. The rigid modes appear on the diagonal<br />

elements <strong>of</strong> the 6x6-transfer function. For the lower frequencies, there is a considerable<br />

difference in gain between the different directions due to the variation in mass (and thus the<br />

pressure necessary to move with the same speed). Above the rigid mode eigenfrequencies<br />

the gains are expected to be equal apart from the effects due to the parasitic modes.<br />

The three assumably flexible modes at 43, 57 and 65Hz mainly appear along the direction<br />

<strong>of</strong> the most stiff two rigid modes. Along the pitch/surge mode direction the 43<br />

and 65Hz resonances are experienced. The 57Hz vibration comes in along the roll/sway<br />

direction.<br />

The pressure transfer function constructed from the accelerometers assuming a rigid<br />

construction should be passive apart from the influence <strong>of</strong> the transmission lines (and <strong>of</strong><br />

course the valve characteristics). Passive behaviour is certainly lost at approx. 60Hz where<br />

the phase along the direction <strong>of</strong> the pitch/surge mode (6) drops another rad.<br />

Velocity dynamics.<br />

Another alternative to reconstruct a decoupled transfer function is to use the velocities along<br />

the rigid body modal directions. These frequency responses are plotted in Fig. 4.20. These<br />

plots show the fact that for low frequencies, the hydraulic actuator is a velocity generator.<br />

This enables easy evaluation <strong>of</strong> the differences in the rigid mode resonance gains. It also<br />

shows that for the most stiff modes, the acceleration at the frequencies above the rigid body<br />

modes is the highest. Probably the (high frequency) flexible modes appear mainly along<br />

these directions due to the fact that they are best measured along these directions.<br />

Flexible phenomena.<br />

A closer inspection <strong>of</strong> these flexible modes was done by evaluation <strong>of</strong> the additional accelerometer<br />

measurements taken at the front <strong>of</strong> the shuttle. From visual inspection <strong>of</strong> the<br />

shuttle, it can be expected that the weakest part <strong>of</strong> the shuttle appears at this point where<br />

a large hole has been provided for the front window. Possibly torsional or chin up like<br />

deformation could be induced. By comparison <strong>of</strong> the actual accelerations measured at the<br />

front and those reconstructed from the (back) floor assuming rigid <strong>motion</strong>, this has been<br />

investigated.<br />

In Fig. 4.21 three examples <strong>of</strong> actually measured and reconstructed frequency responses<br />

from inputs along some specific directions are given, some <strong>of</strong> which are expected to induce<br />

energy for the flexible modes. These directions are given in Table 4.7.


160 4 Parameter identification and model validation<br />

10 1<br />

10 0<br />

10 −1<br />

10 1<br />

10 0<br />

10 −1<br />

10 1<br />

10 0<br />

10 −1<br />

10 1<br />

Frequency (Hz)<br />

10 2<br />

0<br />

−200<br />

−400<br />

0<br />

−200<br />

−400<br />

−600<br />

−800<br />

0<br />

−200<br />

−400<br />

−600<br />

−800<br />

10 1<br />

Frequency (Hz)<br />

Fig. 4.21: Bode plots <strong>of</strong> measured (dotted) and reconstructed rigid part <strong>of</strong> the acceleration at<br />

front middle in approx. y (upper row), x (middle row) and y (lower row) direction<br />

due to input vibration along the yaw, ’chin up’ and ’torsional x-ax’ directions<br />

respectively. Left column provides amplitude and right column provides phase<br />

responses. Differences in measured and reconstructed plots can point at flexible<br />

modes.<br />

10 2


4.4 Chapter Resume 161<br />

In the yaw direction it can be observed that the accelerometer output at front can be predicted<br />

fairly well by the accelerometers at the back <strong>of</strong> shuttle on the floor. However, none<br />

<strong>of</strong> the flexible modes to be investigated (43,57,65 Hz) is very pronounced in this direction.<br />

For the chin up input, the reconstructed and measured output for the accelerometer mounted<br />

for the x-direction result in totally different frequency responses. In the reconstructed frequency<br />

response, it is not expected that the modes at 43 and 65Hz will be observed while<br />

they appeared to be very pronounced in the measurements at this point. Finally, with the<br />

torsional x-axis input it is shown that around 57Hz the phase <strong>of</strong> the bode plots starts to<br />

differ 180 . The measurements have a complex zero pair like characteristic at 50Hz while<br />

the reconstructed response has not.<br />

In many ways, the reconstructed and measured responses differ. Too much to closely<br />

predict the characteristics <strong>of</strong> the flexible modes very accurately at this point. In some directions,<br />

it seemed the accelerometers were not sufficiently excited by the experiments performed.<br />

If exact knowledge <strong>of</strong> the flexibilities is required, more thorough modelling together<br />

with further experiments is necessary. In this research the parasitic resonances were<br />

mainly investigated in order to give a proper view on the extent to which the models not<br />

including these effects are valid.<br />

Resume<br />

Through frequency response model validation it was shown that the model structure<br />

proposed fairly well describes the characteristics <strong>of</strong> the actual <strong>system</strong>.<br />

Especially the rigid body modes and directions resulting from the interaction <strong>of</strong> hydraulics<br />

and mechanics, which were predicted rightly to dominate the response, are<br />

well caught by the model.<br />

Low frequency ( 10Hz) parasitic behaviour could be pinpointed to the dynamics <strong>of</strong><br />

the floor, also well described by the model including the floor.<br />

Accelerometers could be used to reconstruct rigid body platform accelerations, velocity<br />

and pressure dynamics and point at possible flexibilities, after careful identification<br />

<strong>of</strong> accelerometer line <strong>of</strong> action and gain.<br />

The shuttle flexible modes could be assigned to part <strong>of</strong> the high frequent parasitic<br />

resonances but could not yet be included in the model structure. This will also require<br />

more rigorous measurement to allow accurate mode shape identification.<br />

4.4 Chapter Resume<br />

In this chapter it has been shown that the parameters <strong>of</strong> the models derived on physical laws<br />

could be identified by performing a very limited number <strong>of</strong> experiments. The responses<br />

<strong>of</strong> the kinematical model <strong>of</strong> the <strong>motion</strong> <strong>system</strong> did have a very close match with the measurements<br />

taken. The characteristics <strong>of</strong> the dynamics modelled for the hydraulically driven


162 4 Parameter identification and model validation<br />

mechanical <strong>system</strong> does correspond fairly well to the actually observed dynamics up till<br />

reasonable high frequencies ( 30Hz) if the dynamics <strong>of</strong> the foundation is taken into account.<br />

The properties due to the valve and transmission lines in the hydraulic <strong>system</strong> are<br />

still important for the over all <strong>motion</strong> <strong>system</strong>. But at the frequencies where these effect have<br />

their main influence ( 30Hz) additional parasitic resonances due to the mechanics have<br />

to be taken into account, especially in case <strong>of</strong> the light weight shuttle.


Chapter 5<br />

<strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong><br />

<strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

5.1 Introduction<br />

In previous chapters it was shown through theoretical modelling and experimental verification<br />

that in <strong>system</strong>s like the Simona <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>, the dynamics due to the<br />

hydraulic actuators and the parallel driven mechanical <strong>system</strong> with six degrees <strong>of</strong> freedom<br />

form an integrated structure with typical characteristics, which can be modelled with reasonable<br />

complex equations. This is in practice the most elaborate part <strong>of</strong> design in using a<br />

model <strong>based</strong> <strong>control</strong> strategy as laid down in this chapter.<br />

Referring to the discussion in Chapter 1, the eventual task <strong>of</strong> high performance <strong>simulator</strong><br />

<strong>motion</strong> <strong>control</strong> is improved <strong>motion</strong> realism. With <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s one wants<br />

to provide the pilot in the <strong>simulator</strong> with appropriate generalized specific forces i.e. both<br />

rotational and translational accelerations plus gravity [92]. Particularly in case <strong>of</strong> tight pilot<br />

<strong>control</strong>, e.g. landing, phase lag <strong>of</strong> realized versus desired accelerations should be minimal.<br />

In this way smaller differences between <strong>simulator</strong> and in-the-air <strong>flight</strong> conditions will exist.<br />

In <strong>flight</strong> simulation, the pilot relies on the perception <strong>of</strong> self-<strong>motion</strong> through several<br />

stimuli, and uses this perception to exercise <strong>control</strong> over the aircraft. Stimuli cueing <strong>system</strong>s<br />

include the visual <strong>system</strong>, the <strong>motion</strong> <strong>system</strong>, the audio <strong>system</strong>, <strong>control</strong> loading and the<br />

aircraft instruments stimulation. In the <strong>control</strong> task, the pilot uses his visual perception to<br />

make a good estimation <strong>of</strong> the aircraft’s attitude and velocity. The approximate frequency<br />

response <strong>of</strong> visual perception can be modelled as a first order low-pass filter with a break<br />

frequency <strong>of</strong> 0.1 Hz [92], which is rather slow.<br />

Motion in the higher frequency area is, however, mostly perceived by the pilot’s vestibular<br />

and tactile sensors, which are sensitive to specific forces and angular accelerations.<br />

These signals are processed rapidly by the central nervous <strong>system</strong> and, therefore, give the pilot<br />

lead information. With this lead information, the pilot can react more quickly to changes<br />

in the vehicle <strong>motion</strong> state. This information is most important at the higher frequencies<br />

(above the bandwidth <strong>of</strong> visual perception). These high-frequency <strong>motion</strong>s are, in simulation,<br />

<strong>of</strong>ten called onset-cues.<br />

163


164 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Because <strong>of</strong> the high-frequency nature <strong>of</strong> these onsets, it is important that the <strong>motion</strong><br />

<strong>system</strong> has sufficient bandwidth. Moreover, it is important that the time-delay in the <strong>motion</strong><br />

simulation is kept as small as possible, since an onset is also time-critical. If the onset is<br />

simulated noticeably late, <strong>simulator</strong> training quality will be decreased considerably. Time<br />

delays, furthermore, lower the pilot-vehicle crossover frequency and may require the pilot<br />

to adapt by applying lead compensation. The reduction <strong>of</strong> the adaptation required is<br />

precisely one <strong>of</strong> the goals <strong>of</strong> future <strong>flight</strong> training [7]. Therefore, a primary objective <strong>of</strong><br />

a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> is to provide the pilot in the <strong>simulator</strong> with appropriate<br />

representations <strong>of</strong>, what one could call, the ”generalized forces” i.e. both translational and<br />

rotational accelerations and gravity. This has to be attained by driving the <strong>simulator</strong> with<br />

six parallel hydraulic servo actuators. With the models <strong>of</strong> the SRS, an important basis for<br />

high performance <strong>motion</strong> has been attained further helped by construction design for <strong>control</strong><br />

including low-mass, low centre-<strong>of</strong>-gravity, and a high-stiffness structure. Its potential<br />

can only be fully exploited by advanced <strong>control</strong>ler design techniques such as used in the<br />

area <strong>of</strong> robotics.<br />

Another necessary requirement is the application <strong>of</strong> appropriate hardware components,<br />

like fast multi-processor Digital Signal Processor (DSP) boards, and s<strong>of</strong>tware which performs<br />

automatic DSP-code generation from higher level programs e.g. Matlab/Simulink.<br />

Thereby, it is possible to use highly structured complex <strong>control</strong> strategies which take model<br />

knowledge into account. This includes fast iteration on <strong>control</strong>ler design methods consisting<br />

<strong>of</strong> setting the specifications, analyzing the <strong>system</strong>, synthesising the <strong>control</strong>ler and<br />

implementing on the actual experimental set up.<br />

5.2 Another look at the <strong>control</strong> problem<br />

Since the <strong>motion</strong> <strong>control</strong>ler to be designed has to be implemented on a complex <strong>system</strong>, the<br />

previous chapters dealt with extensive modelling, which took place with the <strong>control</strong> objective<br />

in mind, even before construction [74], [124]. In the earlier stages measurements could<br />

be done on an experimental set up in which each hydraulic actuator was tested separately.<br />

Further, tests were performed with a dummy platform and the empty shuttle replacing the<br />

eventual <strong>simulator</strong> on top <strong>of</strong> the <strong>motion</strong> <strong>system</strong> [75].<br />

<strong>Model</strong> <strong>based</strong> <strong>control</strong> strategies such as the computed torque like methods are directly<br />

applicable for mechanical <strong>system</strong>s modelled in the standard structure <strong>of</strong> explicit differential<br />

equations whose states can be measured and whose <strong>control</strong> inputs are the torques which are<br />

aligned to the model coordinates. So, in applying a computed torque like technique, at least<br />

the following issues have to be regarded.<br />

- <strong>Model</strong> equations: A model-<strong>based</strong> method requires modelling. <strong>Model</strong>ling parallel <strong>system</strong>s<br />

typically can result in combined constraining algebraic and differential equations.<br />

For instance, the <strong>motion</strong> <strong>of</strong> an actuator mass in the Stewart Platform, will be<br />

fully dependent on the <strong>motion</strong> <strong>of</strong> the platform itself. Preferably the model equations<br />

will have to be parametrized in such a way that an explicit differential structure results.


5.2 Another look at the <strong>control</strong> problem 165<br />

- <strong>Model</strong> coordinates: Parametrization <strong>of</strong> the model also includes choice <strong>of</strong> the model coordinates.<br />

In the Stewart Platform one can choose end-effector/ platform coordinates,<br />

input coordinate/actuator length or any other coordinate <strong>system</strong> to describe the 6 degrees<br />

<strong>of</strong> freedom. Parallel <strong>system</strong>s are usually and most conveniently modelled in<br />

end-effector coordinates. In general these are not the coordinates which can be directly<br />

measured or steered.<br />

- Extent <strong>of</strong> modelling: Every <strong>system</strong> can be modelled in several ways and in various degrees<br />

<strong>of</strong> accuracy and complexity. In case <strong>of</strong> the mechanics <strong>of</strong> the Stewart Platform<br />

the simplified model <strong>of</strong> (2.112) could be taken. But next to the <strong>simulator</strong> body also<br />

the actuators themselves have mass properties which can not always be neglected.<br />

- Measurements: A full state measurement is assumed. In mechanical <strong>system</strong>s this means<br />

that a number <strong>of</strong> independent positions and velocities equal to the number <strong>of</strong> degrees<strong>of</strong>-freedom<br />

<strong>of</strong> the <strong>system</strong> has to be measured. On top <strong>of</strong> this, not only position and<br />

velocity measurement are assumed to be taken, but it is also assumed that these quantities<br />

directly represent the model coordinates.<br />

With the Simona <strong>motion</strong> <strong>system</strong> only actuator lengths (next to actuator pressure) are<br />

measured. So not only the velocity but also the appropriate model position or platform<br />

pose have to be constructed somehow.<br />

- Control inputs: The same goes for the <strong>control</strong> inputs. These are assumed to be the generalised<br />

forces along the model coordinates. These would be the platform forces in<br />

case <strong>of</strong> the Stewart Platform. Since the actuator forces have several (position dependent)<br />

components in platform coordinates, these can not be considered to be the<br />

required <strong>control</strong> inputs using a standard feedback linearisation scheme without any<br />

further manipulation.<br />

Also the forces will have to be generated by an actuator which in practice cannot<br />

apply required forces instantaneously. Actuator dynamics has to be looked into. In<br />

case <strong>of</strong> the Simona <strong>motion</strong> <strong>system</strong>, the actuators are hydraulic servo <strong>system</strong>s. This<br />

kind <strong>of</strong> actuators can not be considered force generators as such.<br />

Direct application <strong>of</strong> the feedback linearising <strong>control</strong> on the <strong>motion</strong> <strong>system</strong> requires<br />

consideration <strong>of</strong> the various problems observed. Other aspects also need to be looked into.<br />

In the sequel several issues will be discussed, and where necessary the <strong>control</strong> strategy will<br />

be modified.<br />

By analysis <strong>of</strong> both theoretic models derived from basic physical laws as experimental<br />

models <strong>based</strong> on measurements taken, an inventory <strong>of</strong> the relevant <strong>control</strong> problems can be<br />

put together.<br />

Control objectives<br />

Flight simulation or fooling a pilots <strong>motion</strong> awareness basically forms a <strong>control</strong> problem<br />

with mixed objectives. The <strong>system</strong> should provide for the accelerations being<br />

simulated without running out <strong>of</strong> stroke. This problem is mainly left to a host which<br />

has to come up with feasible trajectories but the <strong>motion</strong> <strong>control</strong>ler still has to both<br />

track reference accelerations as to stabilize platform pose.


166 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Hydraulics.<br />

Control <strong>of</strong> the long-stroke hydraulic actuators used in <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s<br />

is not easy since the phase lag introduced by the servo valve together with the nonnegligible<br />

high frequent transmission line resonances form a stability problem [125].<br />

Hydraulics/Mechanics<br />

Further, the bilateral coupling <strong>of</strong> the hydraulic mechanical <strong>system</strong> with strong energy<br />

exchange via pressure/flow and force/velocity introduces mechanical pose dependent<br />

resonances with interaction over the actuators. As was shown, the dynamics resulting<br />

from this interaction forms the most relevant part <strong>of</strong> the <strong>system</strong>.<br />

Mechanics<br />

With the actuators mounted in parallel to the <strong>simulator</strong>, the construction forms a Stewart<br />

platform. Due to the resulting kinematic loops, care should be taken to model this<br />

<strong>system</strong> with only explicit differential equations [74]. This required modelling in appropriate<br />

coordinates i.e. the platform pose. As only the actuators lengths are being<br />

measured, a transformation to platform pose was required to be able to apply model<br />

<strong>based</strong> <strong>control</strong>. This transformation is, dual to the actuator trajectory generation <strong>of</strong> serial<br />

robots, explicitly known from platform pose to actuator length but not injective.<br />

Trajectory generation<br />

The reference acceleration and pose will be calculated at a low sample rate host computer<br />

which incorporates a complex airplane model. References have to be introduced<br />

smoothly to the <strong>motion</strong> <strong>control</strong>ler. But smoothing the signal should not result in responses<br />

with too much phase lag since the timing <strong>of</strong> on-set <strong>motion</strong> is an important<br />

part <strong>of</strong> simulation quality.<br />

In the next section, a <strong>control</strong> strategy will be described, which takes into account the<br />

afore mentioned problems and resulted in a <strong>control</strong>ler which could be implemented on a<br />

real-time <strong>control</strong> computer connected to the <strong>motion</strong> <strong>system</strong>.<br />

5.3 Control Strategy<br />

The <strong>control</strong> strategy will be targeted at generating the appropriate accelerations for the simulation<br />

and at the same time stabilising the platform pose. Taking into account the reference<br />

generating <strong>system</strong>, e.g. the airplane, enables sufficiently smooth accelerating trajectories.<br />

Now, acceleration <strong>of</strong> the <strong>simulator</strong> mechanical <strong>system</strong> results from the forces applied.<br />

Through a model <strong>of</strong> the mechanical <strong>system</strong> the necessary forces can be calculated. Minimal<br />

invasive but stabilising position feedback is attained through use <strong>of</strong> a decoupled coordinate<br />

<strong>system</strong>. The hydraulic actuators can generate the required forces but in the previous<br />

chapters we observed that they face strong interaction from the mechanics. By supply <strong>of</strong><br />

appropriate oil flow the hydro-mechanical coupling can be minimized and local pressure<br />

feedback enables the desired force generation.<br />

By design <strong>of</strong> a <strong>control</strong> structure which has four levels, each <strong>of</strong> which have their own<br />

specifications in close relation with the other levels, one can circumvent having to solve


5.3 Control Strategy 167<br />

Fig. 5.1: Multi level <strong>control</strong> <strong>of</strong> the Simona Motion System.<br />

1. Inner loop feedback: Each hydraulic actuator is feedback transformed into a<br />

force generator by the inner loop feedback using the input pressure, dp i, and estimated<br />

velocity, ^ _q i, <strong>of</strong> an actuator. Input to these sub<strong>system</strong>s are the desired input<br />

pressures, dpi�d.<br />

2. Feedback linearisation: The desired pressures are calculated using a model (2a)<br />

<strong>of</strong> the multivariable mechanics, which calculates gravity, centripetal and coriolis<br />

force compensation and uses the mass matrix to calculate the required forces for<br />

acceleration. This block uses the measured platform state (pose, sx, and estimated<br />

velocity, ^ _x) to do this feedback linearisation step. This state is calculated on line<br />

from the measured actuator position and velocity in an iterative manner (2b).<br />

3. Outer loop feedback: This is outer loop state feedback <strong>of</strong> the platform coordinate<br />

and velocity errors, the difference between the desired xs dr and _xdr and<br />

reconstructed coordinates. Corrective accelerations, xc, result from this block and<br />

are fed to to feedback linearisation block (2.).<br />

4. Reference model-<strong>based</strong> feedforward: Smooth and possibly predictive acceleration,<br />

velocity and position, xdr� _xdr� xs, are reconstructed from the desired accelerations<br />

supplied by the host, xd. They are output to the feedback linearisation<br />

block to be used for feed forward.<br />

a too complex set <strong>of</strong> problems at once. Further, it is shown that this structure leads to an<br />

implementable <strong>control</strong>ler.<br />

General idea<br />

Looking at Fig. 5.1, in which the <strong>control</strong> structure is schematically depicted and described,<br />

we consider the following <strong>control</strong> levels with reference to the applied <strong>control</strong> theory [76].<br />

Level 1. Local hydraulic pressure <strong>control</strong> loops [49], [124], [127].<br />

Level 2. Multivariable feedback linearisation [71], [133].<br />

Level 3. Outer loop position stabilisation [113].


168 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

-Llm -B<br />

R R<br />

- V - ? d- - d?<br />

- M - -<br />

6<br />

1<br />

C<br />

- Ap J -<br />

u dp<br />

f<br />

_x<br />

T -<br />

-Ap<br />

Fig. 5.2: Basic structure hydraulically driven <strong>motion</strong> <strong>system</strong>.<br />

Level 4. Reference model <strong>based</strong> feed-forward [72], [111], [144].<br />

In short, the actuators are turned into pressure generators by local <strong>control</strong>lers. These<br />

<strong>control</strong>lers receive their reference pressure from a feedback linearisation loop in which pressures<br />

can be calculated necessary to track desired accelerations. Desired accelerations are<br />

partly corrections which are required to stabilise the pose <strong>of</strong> the <strong>simulator</strong> and for most<br />

the cues generated to provide the pilot with reasonable <strong>motion</strong> awareness. As these cues<br />

have to be smoothed but not delayed, a reference model <strong>based</strong> <strong>control</strong>ler has to calculate<br />

appropriate cues for the feedback linearisation <strong>control</strong>ler.<br />

The next sections will describe the different <strong>control</strong> levels more closely.<br />

5.4 Inner loop pressure <strong>control</strong><br />

To turn the hydraulic actuators into nice force generators two <strong>of</strong> the afore mentioned <strong>control</strong><br />

problems have to be solved at this level. Feedback <strong>of</strong> the pressure can result in stability problems<br />

since the relatively long transmission lines cause badly damped resonances together<br />

with phase lag <strong>of</strong> the valve. Further, the coupling between the mechanics and hydraulics<br />

results in the pose and load dependent rigid modes <strong>of</strong> the <strong>system</strong>.<br />

As shown in Fig. 4.13, for the shuttle the rigid modes can be observed in the frequency<br />

area between 7Hzand 25 Hz. At200 Hz the transmission lines cause peaking and at 75 Hz<br />

also a notch results. The valve has a bandwidth <strong>of</strong> 150 Hz which can clearly be observed by<br />

looking at the phase. Flexibility in the mechanics caused some additional parasitic modes<br />

between 40 Hz and 80 Hz.<br />

The coupling from which the rigid modes result, can be dealt with using the <strong>control</strong><br />

method introduced by Sepehri [127] and successfully applied by Heintze [49]. A hydraulically<br />

driven <strong>motion</strong> <strong>system</strong> basically has the structure given in Fig. 5.2. Through the valves,<br />

V , the oil flows, , can be steered by the inputs, u. The required oil flows are mainly determined<br />

by the speed at which the volumes in the actuators have to be filled. These are<br />

equal to the velocities, _q, <strong>of</strong> the actuators times the area <strong>of</strong> the piston, Ap. Together with<br />

the oil loss due to the leakage Llm, the net oil flow difference cause the pressures dp to rise<br />

through the hydraulic oil stiffness, C.<br />

fa<br />

_q<br />

J<br />

_<br />

dp = C(V u ; Llmdp ; Ap _q) (5.1)


5.4 Inner loop pressure <strong>control</strong> 169<br />

dp d<br />

?<br />

-1<br />

V ;1<br />

d<br />

d?-<br />

- -<br />

6<br />

- Kdp<br />

Inner loop<br />

<strong>control</strong>ler<br />

A ;1<br />

p<br />

6<br />

-<br />

u<br />

-Llm<br />

V - ? d-<br />

C -<br />

6<br />

Fig. 5.3: Basic structure inner loop <strong>control</strong>lers. On the right, the hydraulic part <strong>of</strong> the <strong>system</strong><br />

structure <strong>of</strong> Fig. 5.2 is given. On the left, the inner loop <strong>control</strong>ler, the cascade<br />

dp structure, compensates for the influence <strong>of</strong> the mechanics and <strong>control</strong>s the<br />

pressure, dp, which results in a theoretically arbitrarily fast first order response dependent<br />

on the <strong>control</strong>ler feedback gains in K dp. In case <strong>of</strong> the SRS, _q can not be<br />

measured and has to be reconstructed.<br />

The acceleration <strong>of</strong> the actuators is determined by the inverse mass matrix, M ;1 , which<br />

causes the interaction, times the forces supplied by the actuators minus the viscous friction<br />

e.g. b due to each hydraulic bearing, B = J T l�x bIJl�x, with Jl�x as the pose dependent<br />

jacobian.<br />

R<br />

dp<br />

-<br />

Ap<br />

-Ap<br />

Mx = J T l�x Apdp ; B _x (5.2)<br />

The actuator and platform velocities are related through _q = Jl�x _x. Gravity forces and the<br />

less relevant coriolis and centripetal forces are assumed to be dealt with at the higher levels.<br />

In Fig. 5.3, the basic structure <strong>of</strong> an inner loop pressure <strong>control</strong>ler, which decouples the<br />

mechanics from the hydraulics, is given together with the hydraulic part <strong>of</strong> the <strong>system</strong> <strong>of</strong><br />

Fig. 5.2. By compensation <strong>of</strong> the oil flow due to actuator velocity, the hydraulics can be<br />

decoupled from the mechanics.<br />

A smooth 50 Hz bandwidth pressure generator was obtained by filtering the pressure<br />

feedback signal properly for a one degree <strong>of</strong> freedom set up in which the hydraulic actuator<br />

were all tested separately. As the inner loop <strong>control</strong>ler should not interact with the mechanics<br />

its characteristics could be designed and tested in this setting at first instance [124]. In<br />

this way the hydraulic servo actuators, which usually are considered velocity engines, are<br />

approximately turned into force generators. However, as already put forward, the measured<br />

pressure difference (at the valve) does not reflect the applied force over the full relevant<br />

frequency area due to the transmission lines. Already at 30 Hz, this difference amounts to<br />

25%.<br />

The structure <strong>of</strong> the implemented inner loop <strong>control</strong>ler as given in Fig. 5.4 is somewhat<br />

different from the basic structure given in Fig. 5.3. The structure implemented on the single<br />

hydraulic actuator as was used by Van Schothorst [124] is taken. As the velocity <strong>of</strong> the<br />

hydraulic actuator is not measured directly it has to be reconstructed and three filters, C 1,<br />

C2 and C3 have been added to deal with the transmission lines and the valve dynamics.<br />

-<br />

fa<br />

_q


170 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Roughly the filters remove signal content from the frequency area above 100Hz where<br />

the valve and transmission line dynamics are most prominent and only mildly change the<br />

characteristics up to 30Hz where the rigid body dynamics resides.<br />

The linear dynamic filters, C1�C2�C3, used for the platform are given in the frequency<br />

domain by<br />

and for the shuttle<br />

C1(s) =<br />

C2(s) =<br />

C3(s) =<br />

1<br />

1<br />

(2 95) 2 s 2 +<br />

1<br />

1<br />

2 120<br />

C2s(s) =<br />

s +1<br />

1<br />

(2 200) 2 s 2 +<br />

1<br />

(2 400) 2 s 2 +<br />

1<br />

2 0:3 s +1<br />

2 95<br />

2 0:2<br />

2 200<br />

1 s +1<br />

2 120<br />

s +1<br />

2 0:2<br />

s +1<br />

2 400<br />

1<br />

(2 65) 2 s 2 +<br />

1<br />

(2 65) 2 s 2 +<br />

1<br />

1<br />

(2 150) 2 s 2 +<br />

2 0:015<br />

2 65<br />

2 0:13<br />

2 65<br />

2 0:6<br />

s +1<br />

2 150<br />

s +1<br />

s +1<br />

(5.3)<br />

(5.4)<br />

(5.5)<br />

(5.6)<br />

a notch had to be added due to the parasitic resonance <strong>of</strong> the mechanical <strong>system</strong> at this<br />

frequency, which should not be hit. Implementation was done through the standard bilinear<br />

transformation to digital filters running at 5kHz.<br />

The flexibility <strong>of</strong> the shuttle, which enters well above 30 Hz has still to be taken into<br />

account. An illustrative picture is given in Fig. 5.5. Nyquist plots <strong>of</strong> the characteristic<br />

loci are given reconstructed from the 6x6 frequency domain measurements taken for the<br />

shuttle by taking the eigenvalues <strong>of</strong> the measured transfer function, T , (T (j!)), for each<br />

frequency, !. Recall that since the multivariable <strong>system</strong> can approximately be decoupled<br />

by a unitary matrix at each operating point, Nyquist plots can be constructed, which can be<br />

analyzed as in the SISO case as given in (3.47).<br />

In the upper left part <strong>of</strong> Fig. 5.5, no feedback filter is applied and a pressure feedback<br />

gain <strong>of</strong> kdp = 0:35 is chosen. All rigid body modes and flexible resonances remain<br />

’circling’ in the right half plane as our passivity analysis <strong>of</strong> Section 3.3.1 predicted. The<br />

additional phase lag <strong>of</strong> the valve rotates the transmission line dynamics into the left half<br />

plane encircling the stability point, ;1.<br />

The gain <strong>of</strong> kdp =0:35, is well below the achievable gain <strong>of</strong> kdp =0:75 reported on<br />

in the SISO case [124]. For the SISO case this gain could safely be used in combination<br />

with the dynamic filters. The parasitic dynamics due to the flexibility <strong>of</strong> the shuttle, however,<br />

complicates the separation <strong>of</strong> low (rigid body modes, foundation) and high frequent<br />

phenomena (valve and transmission lines). The upper right plot <strong>of</strong> Fig. 5.5 still shows a<br />

resonance (at 65Hz) encircling the stability point. An additional notch solves the stability<br />

problem as shown in the lower left and right plots <strong>of</strong> Fig. 5.5 although the resonances at 43<br />

and 57Hz become more prominent.<br />

In theory, measured velocity compensation should remove all influences, also the flexibilities,<br />

from the pressure dynamics. As velocity could only be estimated with limited<br />

accuracy, a considerable amount <strong>of</strong> this compensation was done with the required velocity<br />

instead. This leaves the pressure feedback dynamics for a large part unchanged.


5.4 Inner loop pressure <strong>control</strong> 171<br />

dpi�d<br />

C1<br />

- -+ ? - -+ -<br />

+<br />

6<br />

C3<br />

e Kdp<br />

e<br />

C2<br />

-<br />

Fig. 5.4: Implemented structure inner loop <strong>control</strong>lers with velocity observer and additional<br />

filters, Ci. The output, ^ i, <strong>of</strong> this <strong>control</strong>ler part is still an intermediate <strong>control</strong><br />

variable, which is input to the valve flow compensation module <strong>of</strong> (5.7).<br />

IMAG<br />

IMAG<br />

5<br />

0<br />

−5<br />

−10<br />

−10 0 10 20<br />

REAL<br />

5<br />

0<br />

−5<br />

−10<br />

−5 0 5 10<br />

REAL<br />

IMAG<br />

5<br />

0<br />

−5<br />

K _q<br />

6<br />

_qobs<br />

−10<br />

−5 0 5 10<br />

REAL<br />

−1.5<br />

−1 −0.5 0 0.5<br />

REAL<br />

Fig. 5.5: Largest characteristic locus for pressure feedback with just feedback gain (k dp =<br />

:35) (upper left), feedback filter applied (upper right), feedback filter with notch<br />

for the flexible platform (lower left and right).<br />

IMAG<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

dpi<br />

- ^ i<br />

_qref<br />

dp i<br />

q


172 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

10 0<br />

10 −1<br />

0<br />

−100<br />

−200<br />

−300<br />

10 1<br />

10 1<br />

Fig. 5.6: Bode plot <strong>of</strong> pressure feedback filter (C 2KdpC1) with flexible platform.<br />

Given the normalized stiffness, C =2Cm 220, similar as in Table 3.2, this results in<br />

a bandwidth <strong>of</strong> 2Cmkdp<br />

=12:2Hz in case only the first order pressure <strong>system</strong> is considered.<br />

2<br />

A Bode plot <strong>of</strong> the resulting feedback <strong>control</strong>ler is shown in Fig. 5.6.<br />

As discussed in Section 3.2.1, the valve has a nonlinear characteristic given by (3.16),<br />

especially in the face <strong>of</strong> load variations. Van Schothorst [124] proposes a <strong>control</strong> structure<br />

u =<br />

10 2<br />

10 2<br />

f ;1<br />

2 (^ i)<br />

q 1 ; sgn( ^ i)dpi<br />

to compensate for these effects. As this was shown to be successful experimentally, it was<br />

made part <strong>of</strong> the inner loop <strong>control</strong>ler for the SMS.<br />

(5.7)<br />

Velocity estimation<br />

The positive velocity compensation loop requires knowledge <strong>of</strong> the actuator velocity. Position<br />

and (indirectly through pressure difference) applied force are measured. Further, it<br />

is known at what velocity the actuators should run. Again, in the thesis <strong>of</strong> Van Schothorst<br />

[124] a range <strong>of</strong> velocity observers is presented. The most simple version consists <strong>of</strong> differentiating<br />

the position signal (with some low pass filtering). With the application <strong>of</strong> the<br />

increased resolution position measurement [55], this version seemed feasible. However, the<br />

position measurement appeared to have a (complex deterministic) error <strong>of</strong> 100 m in amplitude<br />

(and 4:3mm period length) which could result in severe vibrations running at higher<br />

velocities.<br />

Therefore it was decided to use the other sources <strong>of</strong> information as well, i.e. the measured<br />

pressures and the required actuator velocities, which can be calculated by v des =<br />

Jl�x _xdr. A simple second order filter was used to attain cut <strong>of</strong>f at both low (no bias) and


5.5 Multivariable feedback linearisation 173<br />

high (less noise) frequencies for all three inputs <strong>of</strong> the velocity reconstructor.<br />

1<br />

2 0:7<br />

2 6 svdes + sq<br />

(2 6) ^_qobs =(1; )<br />

2 s^q dpi +<br />

1<br />

(2 6) 2 s2 + vdes (5.8)<br />

2 0:7<br />

+ s +1<br />

2 6<br />

The acceleration <strong>of</strong> the actuator, q, had to be reconstructed at the higher level using the<br />

information on the mass properties <strong>of</strong> the <strong>system</strong> and the measured pressure differences<br />

at the valves. The velocity compensation gain was set at k _q = 0:9, just safely below the<br />

theoretic value <strong>of</strong> 1 as the velocity valve gain is still slightly nonlinear. Further, in practice<br />

the feedforward part, vdes, had to be set to 50% , i.e. =0:5.<br />

The parametrization <strong>of</strong> the inner loop <strong>control</strong>, which was implemented, has more or less<br />

been attained through loop shaping. A more structural approach has been taken in [150] by<br />

doing inner loop <strong>control</strong> design using robust <strong>control</strong> techniques. This unfortunately did not<br />

lead to <strong>control</strong>lers which could be implemented safely yet. Additional effort in this direction<br />

is required.<br />

5.5 Multivariable feedback linearisation<br />

Considering the actuators as smooth force generators, a task <strong>of</strong> the upper level is to come<br />

up with proper reference forces. These can be calculated using the model <strong>of</strong> the mechanical<br />

part <strong>of</strong> the <strong>system</strong>. A feedback linearisation structure results.<br />

5.5.1 Feedback linearising robotic manipulators<br />

General feedback linearisation as described in [105] can for rigid body robotic manipulator<br />

models be given in a computed torque <strong>control</strong> structure [133] or an inverse dynamics <strong>control</strong><br />

structure [31]. In [77] it is shown that many <strong>of</strong> such differently named <strong>control</strong> strategies in<br />

fact lead to the same <strong>control</strong> structure.<br />

For a rigid body serial robotic manipulator, the <strong>control</strong>ler is constructed as follows. See<br />

Fig. 5.7. Given the general equations <strong>of</strong> a mechanical model <strong>of</strong> such a <strong>system</strong><br />

M (q)q + C(q� _q) +G(q) = (5.9)<br />

in which similar terms appear as in (2.112). The torques/forces, are assumed to drive the<br />

<strong>system</strong>. In the computed torque <strong>control</strong>ler all terms in (5.9) are compensated for according<br />

to<br />

= M (q)(q d + c) +C(q� _q) +G(q) (5.10)<br />

which leaves decoupled double integrators.<br />

This linear <strong>system</strong> is achieved by compensation with terms, which depend on the measured<br />

state and is therefore called feedback linearisation.<br />

The desired accelerations, q d, can be steered directly, and the double integrators can be<br />

stabilised independently by the outer loop, e.g. a PD position feedback <strong>control</strong>ler for each<br />

input-output channel


174 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Fig. 5.7: Feedback linearising <strong>control</strong> <strong>of</strong> a robotic manipulator<br />

c = K(s)(qd ; q): (5.11)<br />

As the feedback linearising <strong>control</strong> strategy provides a suitable feedforward path for<br />

desired accelerations and turns the non-linear multivariable structure into an easy to <strong>control</strong><br />

decoupled double integrator <strong>system</strong>, the method should be able to deal with most <strong>of</strong> the<br />

specifications set for the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> [83].<br />

However, in the theory presented quite some assumptions are implicitly made. These<br />

have to be checked in applying the method. In the next section it will be put forward that<br />

with parallel robotic <strong>system</strong>s the assumed model structure does not fit in general. The<br />

<strong>control</strong> method has to be modified, which requires additional analysis.<br />

5.5.2 Feedback linearisation <strong>of</strong> a Stewart platform<br />

Also with parallel <strong>motion</strong> <strong>system</strong>s, the force generators should be used to <strong>control</strong> the nonlinear<br />

and multivariable mechanics. With model <strong>based</strong> calculation <strong>of</strong> the required forces to<br />

accelerate along the desired path, xd, given a measured pose and velocity, the <strong>system</strong> is to<br />

be provided with both feed forward and decoupled feedback linearised correction paths to<br />

be used by the higher level <strong>control</strong>lers.<br />

The proposed <strong>control</strong> structure is given in Fig. 5.8. This structure differs from the standard<br />

computed torque <strong>control</strong>ler <strong>of</strong> a mechanical <strong>system</strong>. In modelling for <strong>control</strong> the parallel<br />

Stewart platform configuration, one has to take care <strong>of</strong> generating an explicit set <strong>of</strong><br />

differential equations [74]. Since this is only possible taking the platform pose as the generalised<br />

coordinates, the <strong>control</strong>ler has to incorporate an algorithm which calculates these


5.5 Multivariable feedback linearisation 175<br />

Fig. 5.8: Modified feedback linearising <strong>control</strong> structure<br />

coordinates from the measured actuator lengths, l, and translates desired platform forces<br />

into required actuator pressures.<br />

The actuator lengths, l can be calculated from the platform pose sx.<br />

l = f(sx) (5.12)<br />

As discussed in Chapter 2, in measuring the actuator lengths, the platform pose has to be<br />

reconstructed iteratively e.g. by (2.58).<br />

With the jacobian, Jl�sx, which was defined by<br />

sxk+1 = sxk + J ;1<br />

l�sx (sxk)(lmeasured ; lk) (5.13)<br />

Jl�sx(x) = @l<br />

@sx<br />

(5.14)<br />

In Chapter 2 it was shown that this iteration converges sufficiently fast if some requirements<br />

are fulfilled.<br />

The desired actuator pressures can be constructed by calculation <strong>of</strong> a platform mass<br />

matrix M, coriolis and centripetal forces C _x and gravity G and filling in the desired platform<br />

accelerations, xd for the simulation plus the corrective accelerations, xc, from the outer loop<br />

position <strong>control</strong>. These are functions <strong>of</strong> the reconstructed platform pose and velocity. The<br />

simplified model <strong>of</strong> the Stewart platform taking into account the actuator inertia in addition<br />

to the platform mass matrix given in (2.112) will be considered and further the actuator<br />

forces are approximated by fa = ApdP o ApdP i, where Ap is the normalised operational<br />

area <strong>of</strong> the hydraulic actuator. Multiplying (2.112) by J ;T<br />

l�x , the desired pressure is given by<br />

dP i�d =(ApJl�x) ;T (Mt(xd + xc) +Cc( _x� sx) _x + Gc): (5.15)


176 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

In the first approach, filling in design values for these parameters like masses, inertias and<br />

centres-<strong>of</strong>-gravity already results in reasonable performance by a considerable reduction in<br />

interaction. This was further improved by taking identified model parameters. With the<br />

parameters <strong>of</strong> the kinematical model <strong>of</strong> (4.1) calibrated in Section 4.1.4 the jacobian, J l�x,<br />

can be calculated, as follows from (2.42). The identified platform mass matrix M t <strong>of</strong> (4.31),<br />

including actuator intertia constant w.r.t. the platform, is taken instead <strong>of</strong> the more limited<br />

one body mass matrix Mc <strong>of</strong> (2.112). By using the identified center <strong>of</strong> gravity in (4.30) as<br />

origin <strong>of</strong> the platform coordinates, the model parameters become relatively simple. E.g. as<br />

seen in (4.26), the gravity vector force, G c only consists <strong>of</strong> the identified mass, mg, times<br />

gravitational acceleration g in z-direction.<br />

The inverse jacobian calculation was performed at 1kHz twice as required for the implicit<br />

state measurement. Inversion was implemented through a LU-factorisation with full<br />

pivoting as given in [112].<br />

Gravity compensation is easily made explicit for the actuators by use <strong>of</strong> (the third column<br />

<strong>of</strong>) the inverse jacobian information as<br />

G = ;J ;T<br />

l�x ( � 3)mg� (5.16)<br />

Easily derived from (2.111).<br />

For coriolis and centripetal forces, only the nonlinear term given in (2.111) at platform<br />

force level by<br />

fc =! (Ic!) = ~ Ic! (5.17)<br />

was taken into account. Thereby neglecting the (small) actuator related nonlinear inertial<br />

terms.<br />

5.5.3 Implicit state measurement requirements<br />

As the iteration (5.13) reconstructing the state is part <strong>of</strong> the feedback loop, it has to converge<br />

at all times and moreover sufficiently fast in order to prevent the <strong>system</strong> from going<br />

unstable. This problem has been considered in detail in Section 2.3. Summarising, the convergence<br />

properties <strong>of</strong> the implicit state measurement by the NR-scheme can be looked into<br />

by considering the weak Newton-Kantovorich theorem given by Stoer [141], stating that a<br />

NR-scheme results in a well-defined sequence in a limited area with a solution (limit point)<br />

to which it converges with a guaranteed speed if the following three properties hold.<br />

1. The Lipschitz condition on the jacobian This condition implies that a limited difference<br />

between two platform poses should result in a limited difference between the<br />

two jacobians at those points.<br />

kJl�sx(sx1) ; Jl�sx(sx2)k ksx1 ; sx2k (5.18)<br />

2. ’Away’ from singular positions This means that the smallest gain <strong>of</strong> the jacobian should<br />

be sufficiently far away from zero:


5.5 Multivariable feedback linearisation 177<br />

kJ ;1<br />

l�sx (sx)k (5.19)<br />

With Stewart Platforms this is not trivially true. Singular points exist in the construction<br />

if the actuator lengths are not constrained by limited stroke [97]. For example, if<br />

a platform is moving down until all legs are in the XY-plane, no forces in Z direction<br />

or moments around X and Y axis are possible.<br />

Singularity can be excluded by considering a nonsingular point and the Lipschitz<br />

condition. This excludes a (possibly small) volume from singularity, and the whole<br />

workspace can be proved to be free <strong>of</strong> singularities by gridding. In this way the<br />

workspace <strong>of</strong> the Simona <strong>motion</strong> <strong>system</strong> was proved to be free <strong>of</strong>, and even far enough<br />

from singular points.<br />

3. ’Near’ to solution The third and last condition requires an initial guess which is not too<br />

far away from the real solution:<br />

kJ ;1<br />

l�sx (lmeasured ; lk=0)k : (5.20)<br />

With the described <strong>control</strong> structure implemented on a digital computer with given<br />

sample rate (> 100 Hz) and limited speed <strong>of</strong> the <strong>system</strong>, the previous solution can<br />

be used as an initial guess which can be proved to be close enough. The hydraulic<br />

actuators are physically limited in speed.<br />

Since the three conditions can be proved to be true for the Simona <strong>motion</strong> <strong>system</strong>, as was<br />

shown in detail in Chapter 2, the proposed <strong>control</strong> structure can be used without stability<br />

problem concerning the iterative part <strong>of</strong> the <strong>control</strong>ler. Given the speed <strong>of</strong> convergence, the<br />

solution is close enough (within sensor accuracy) after two steps. The update frequency is<br />

easily attained by calculating two iterations at 1 kHz.<br />

5.5.4 Outer loop <strong>control</strong><br />

Considering Fig. 5.8, the transfer function from our input, the desired accelerations, x d, to<br />

the platform pose without the feedback path K(s), does contain unstabilized double integrator<br />

paths. The outer loop <strong>control</strong>ler, K(s), will have to stabilise the <strong>simulator</strong> pose to prevent<br />

the actuators from running out-<strong>of</strong>-stroke. As the feedback linearising <strong>control</strong>ler decouples<br />

the mechanics into separate double integrators, the outer loop can generate correction accelerations<br />

resulting from a PD-structure (kpos + kvels) to stabilise these integrators.<br />

The output <strong>of</strong> the outer loop <strong>control</strong>ler, K(s), can be considered as an extra desired<br />

acceleration, additional to xd, to correct the errors in platform pose. These correction accelerations,<br />

xc, should not exceed human sensory thresholds [54] i.e. generate no noticeable<br />

false cues. Therefore the correction should ideally be sufficiently smooth (filtered) and<br />

only requires limited bandwidth (well below 1Hz). In practice, a bandwidth <strong>of</strong> 2Hz was<br />

chosen necessary to achieve sufficient suppression <strong>of</strong> the disturbances and for most the unmodelled<br />

dynamics. With the normalised signals, the feedback gains are easily chosen at


178 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

kpos =158=(2 2) 2 and kvel =17:5 for a damping <strong>of</strong> =0:7 for each platform direction.<br />

The measured platform velocity is constructed from the actuator velocity estimation<br />

multiplied by the inverse jacobian. The angular velocity, which follows from this multiplication<br />

can then be used to calculate euler parameter velocities.<br />

The outer loop <strong>control</strong>ler are split up into translational and rotational platform coordinate<br />

<strong>control</strong>lers. The outer loop for the translational directions is given by<br />

cc = kpos(cd ; c) +kvel(_cd ; _c) (5.21)<br />

For the rotational directions, the nonlinear structure for the parameters used for the description<br />

<strong>of</strong> the rotation should be used. In this case the euler parameters are taken and as<br />

= G T s _!=2, equivalent to (2.22), a double integrator structure <strong>of</strong> the euler parameters can<br />

also be considered in the platform structure <strong>of</strong> Fig. 5.8. The appropriate corrective angular<br />

acceleration to generate a corrective euler parameter acceleration can be found through<br />

where the corrective euler parameter acceleration, T<br />

<strong>control</strong> structure on the reduced euler parameters<br />

and the reconstruction <strong>of</strong> 0�c is found by<br />

_!c =2G( e) e�c� (5.22)<br />

e�c = [ 0�c<br />

T<br />

c ], is taken as the P(I)D<br />

c = kpos( d ; )+kvel(_ ; _) (5.23)<br />

0�c = ;( T c + _ T<br />

e _ e)= 0� (5.24)<br />

twice differentiating (2.19). Of course, using this strategy, rotations, , should satisfy<br />

< . A disadvantage <strong>of</strong> this structure is the nonlinear structure <strong>of</strong> the euler parameters.<br />

This results in different response on similar rotational errors in different poses. In [45]<br />

an alternative method is proposed, which <strong>control</strong>s difference rotation matrices but it is not<br />

clear yet how to incorporate integral <strong>control</strong> in this structure.<br />

As the pressure generators could not be made an order <strong>of</strong> a magnitude faster than the<br />

outer loop, the references for the outer loop (the desired velocity and position) were made<br />

to lag by using a second order filter with the estimated bandwidth <strong>of</strong> the acceleration generation<br />

(12Hz). Otherwise the outer loop would start to compensate errors, which were<br />

already taken into account.<br />

Although a P(I)D-structure <strong>of</strong> the outer loop <strong>control</strong>ler robustifies the <strong>system</strong> [113],<br />

explicit robust <strong>control</strong> will have to be used to more accurately deal with the varying <strong>system</strong><br />

conditions one encounters working within a real-time environment. A first step towards<br />

such a <strong>control</strong>ler was done as part <strong>of</strong> a master thesis research [122], but this did not lead to<br />

<strong>control</strong>lers, which could be implemented safely. So also at this point many research aspects<br />

are still open.<br />

5.6 Reference model <strong>based</strong> feed forward<br />

The <strong>motion</strong> <strong>control</strong> computer is provided with desired <strong>simulator</strong> accelerations ( xd) and<br />

poses (sxd) by a host, which <strong>control</strong>s the over all simulation (including visual, instrumental


5.6 Reference model <strong>based</strong> feed forward 179<br />

and acoustic stimuli). These signals are generated by a model <strong>of</strong> the vehicle to be simulated<br />

and the sub<strong>system</strong> with the wash-out filters which translate vehicle <strong>motion</strong> into feasible<br />

<strong>simulator</strong> <strong>motion</strong>. Although the host comes up with new set points at a relative low update<br />

frequency (ca. 60 Hz), the fact that the <strong>system</strong>s being simulated are known, will enable a<br />

reasonable prediction <strong>of</strong> the next set point.<br />

The reference model <strong>based</strong> <strong>control</strong> has the task to deal with the set points and future<br />

predictions in a proper way. Using knowledge <strong>of</strong> the set points supplied (mainly the fact<br />

that the signals do not contain information at frequencies higher than 30 Hz) a smooth<br />

interpolation filter provides a suitable reference acceleration to the feedback linearisation<br />

level together with a smooth jerk (derivative acceleration) signal which can be used as lead<br />

signal in the same feed forward channel.<br />

With the reference model <strong>based</strong> <strong>control</strong> considerable improvement can be achieved<br />

w.r.t. phase lag or delay in simulating on-set <strong>of</strong> abrupt (e.g. landing bump) and fast varying<br />

<strong>motion</strong> (e.g. turbulence) [111].<br />

The objective <strong>of</strong> predictive reference model <strong>based</strong> feed-forward <strong>control</strong> is to reduce<br />

latencies to effectively zero. This is achieved by using knowledge <strong>of</strong> the vehicle to be simulated,<br />

known as the reference model, to guide the <strong>simulator</strong> by feed-forward and feedback<br />

<strong>control</strong>, which also accounts for the <strong>motion</strong> <strong>system</strong> dynamics.<br />

Providing the <strong>motion</strong> <strong>control</strong>ler solely with the desired <strong>system</strong>’s output results in latencies,<br />

since pure feedback has limited bandwidth. This observation led to the key idea<br />

that predictive knowledge from the fully available simulation model should be obtained and<br />

used in an appropriate way.<br />

To obtain a solution that can be implemented on and properly integrated with the available<br />

sub<strong>system</strong>s, namely the host and <strong>motion</strong> computer, the problem was split into the<br />

following:<br />

reference model prediction<br />

reference model <strong>based</strong> feed forward <strong>control</strong> design<br />

<strong>system</strong>s interconnection<br />

These problems will be treated subsequently now.<br />

5.6.1 Reference model <strong>based</strong> predictors<br />

The simulation model can be used as a reference signal generator. The only uncertain factor<br />

is the pilot. However, due to the relatively slow response by the human operator, it is quite<br />

easy to predict with reasonable accuracy the future accelerations over a short period <strong>of</strong> time<br />

(30-50 ms). The major problem is due to the complexity <strong>of</strong> the vehicle model, making it<br />

difficult to calculate future values within the real-time environment. Approximate models<br />

can however be used, and several methods have been studied and evaluated [29].


180 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Approaches<br />

The model approximation methods ranged from the very general method <strong>of</strong> series expansion<br />

(with little knowledge <strong>of</strong> the simulation model), to the fast-time modelling approach, in<br />

which the structure <strong>of</strong> the simulation model is taken explicitly into account.<br />

Taylor series expansion.<br />

With the most recent data points <strong>of</strong> the simulated signals, a polynomial is fitted. Extrapolation<br />

<strong>of</strong> this polynomial results in a prediction <strong>of</strong> the future values <strong>of</strong> the signal.<br />

Identified linear model predictor.<br />

A linear state-space model is identified from the non-linear time-variant vehicle model.<br />

With the identified model and its state, a future prediction can be made.<br />

Self tuning predictor.<br />

The previous model parameters can be updated by an adaptation using a stochastic<br />

model and correlation methods.<br />

Fast time modelling (FTM).<br />

The simplified model <strong>of</strong> the equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> the aircraft can be updated by the<br />

coefficients calculated by the original (and more complex) simulation model. In the<br />

construction <strong>of</strong> the predicted <strong>motion</strong>, these coefficients are assumed to be constant.<br />

The advantage <strong>of</strong> the first three methods is that they can be used more independent <strong>of</strong><br />

the specific vehicle to be simulated. However, the expansion method and self tuning predictor<br />

appeared to suffer heavily from high frequent distortion like a turbulence model. One<br />

constant linear approximation model was not able to come up with high quality predictions<br />

at all operation points. The FTM method proved to be best suited for prediction over limited<br />

time horizons. It revealed constant performance under different test conditions.<br />

The parameters transferred to a linear model <strong>of</strong> the equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> the aircraft<br />

are:<br />

Aerodynamic coefficients,<br />

Thrust,<br />

Mass parameters,<br />

Aircraft state vector.<br />

With these parameters together with the wash out filter parameters (which do not have much<br />

effect on the short horizon as they merely act as high pass filters) the required acceleration<br />

<strong>of</strong> the <strong>simulator</strong> can be obtained<br />

The total <strong>of</strong> forces and moments considered acting on the aircraft is a summation <strong>of</strong><br />

those generated by the engine model, aerodynamic model, landing gear model and turbulence<br />

model. As forces immediately result in accelerations considering the equations <strong>of</strong><br />

<strong>motion</strong>, the prediction horizon should be obtained by the ability to predict the future forces


5.6 Reference model <strong>based</strong> feed forward 181<br />

smoothly over some time frame. As thrust, aileron, elevation and rudder steering and landing<br />

gear model have reasonable time constants, this can be done with respect to the engine,<br />

aerodynamic and landing gear forces. The turbulence related forces vary considerably faster<br />

but as the dependence on the pilots actions can be assumed constant over short horizons, the<br />

stochastic nature can be simulated in advance to deliver future predictions.<br />

5.6.2 Construction <strong>of</strong> the reference model <strong>based</strong> <strong>control</strong>ler<br />

An important aspect with respect to the reference model <strong>based</strong> <strong>control</strong> is that there is no<br />

direct response <strong>of</strong> the acceleration (or force generation) resulting from the <strong>control</strong> inputs,<br />

u. This is due to the presence <strong>of</strong> the valve dynamics and the finite oil stiffness in the<br />

column. With the application <strong>of</strong> model-<strong>based</strong> <strong>control</strong>, the desired <strong>motion</strong> outputs (namely,<br />

the specific forces and angular accelerations) can be driven more directly. Therefore the<br />

design <strong>of</strong> a feed forward signal becomes more straight forward.<br />

First, the requirements for the reference model <strong>based</strong> <strong>control</strong> task will be given. Then,<br />

the different approaches will be outlined.<br />

Requirements<br />

In the ideal case, the multiple-level <strong>control</strong>ler would result in a transfer function from the desired<br />

platform accelerations, xd to actual accelerations x equal to a fast first-order response,<br />

determined by the time constant <strong>of</strong> the pressure feedback.<br />

In practice, however, several other aspects lead to deficiencies in the <strong>system</strong>. First <strong>of</strong> all,<br />

the velocity compensation is not perfect and, as a result, the pressure feedback gains cannot<br />

be increased to arbitrarily high values. Nonetheless, the hydraulic actuators can be turned<br />

into ’force generators’ with a bandwidth <strong>of</strong> about 2-3 times the lowest natural frequency <strong>of</strong><br />

the rigid-mass <strong>system</strong> with finite oil spring stiffness. In the case <strong>of</strong> the SRS, this natural<br />

frequency, with a design load <strong>of</strong> 4000 kg, is 4-7 Hz. Finally, the limited bandwidth <strong>of</strong> the<br />

servo valves accounts for an additional latency <strong>of</strong> about 5 to 10 ms as was shown in Fig. 3.7.<br />

Furthermore, the reference accelerations should not have a frequency content higher<br />

than 20-30 Hz: Undesirable deformations <strong>of</strong> the <strong>simulator</strong> result in parasitic resonances<br />

slightly above this frequency range. The pilot’s visual <strong>system</strong> is highly sensitive to vibrations<br />

<strong>of</strong> the visual display <strong>system</strong> optics.<br />

Approach<br />

A study <strong>of</strong> literature resulted in three options for a trajectory tracking <strong>control</strong> approach.<br />

Servo Control. If little use can be made <strong>of</strong> knowledge <strong>of</strong> a reference model or <strong>of</strong> the<br />

<strong>system</strong> to be <strong>control</strong>led, the servo <strong>control</strong> <strong>of</strong> Desoer [33] still results in the robust<br />

asymptotic tracking <strong>of</strong> a reference signal. Transient response quality, which is considerably<br />

relevant in <strong>motion</strong> simulation where onsets play an important role, cannot<br />

be guaranteed however. The hydraulically driven <strong>motion</strong> <strong>system</strong> has tracking quality<br />

w.r.t. step signals <strong>of</strong> the position if position feedback is applied due to the physical<br />

structure given in Fig. 5.2.


182 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

<strong>Model</strong> Matching. Vehicle simulation can also be seen as trying to match the dynamics<br />

<strong>of</strong> the ’washed out’ vehicle model by the dynamics <strong>of</strong> the <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>.<br />

<strong>Model</strong> matching [156] can be applied with feed-forward, feed back, or both.<br />

Preview Control. With preview <strong>control</strong> [144], predicted references can be used in a<br />

<strong>control</strong>ler in order to yield limited phase lags for <strong>system</strong>s not having stable inverses.<br />

The scheme, given in Fig. 5.1, results in tracking for the long term <strong>of</strong> the platform pose<br />

which is stabilized by the outer loop position feedback. Further, as already pointed out, the<br />

inner loop feedback and feedback linearising <strong>control</strong> result in first order responses <strong>of</strong> the<br />

<strong>system</strong>, from desired to actual accelerations nominally. This <strong>system</strong> can be described by<br />

Gnom(s) =<br />

1<br />

I� (5.25)<br />

s +1<br />

where the time constant is determined by one over the product <strong>of</strong> pressure feedback and<br />

the oil stiffness times valve gain, CV . The oil stiffness, C, varies slightly with the position<br />

<strong>of</strong> the actuator[124]. The valve gain also shows some non-linearity.<br />

The immediate unit (step) response <strong>of</strong> a <strong>system</strong> can be obtained by precompensation<br />

with an inverse <strong>system</strong>. Because the <strong>system</strong> Gnom(s) is strictly proper, the inverse is not a<br />

proper stable <strong>system</strong>. Two approaches can be taken to overcome this problem.<br />

First, knowledge <strong>of</strong> the reference <strong>system</strong> can be used to determine the derivative acceleration,<br />

the ’jerk’ signal d=dt(xd). If<br />

xr = d<br />

dt (xd) +xd<br />

(5.26)<br />

is used as feed forward, both the phase lag and high frequent amplitude decrease are<br />

compensated for. As high frequency components tend to be amplified in this way, one<br />

needs to be sure that the reference signal does not contain high frequent components<br />

which could excite the parasitic resonances <strong>of</strong> the <strong>motion</strong> <strong>system</strong>. This could be<br />

reduced somewhat by filtering the reference acceleration with first order filter with<br />

a shorter time constant than , which results in a faster but still lagging response.<br />

However, as is not even exactly known and is not constant, the compensation can<br />

be wrong in the mid-frequency area i.e. the bandwidth area <strong>of</strong> the <strong>motion</strong> <strong>system</strong> <strong>of</strong><br />

10-20 Hz.<br />

Alternatively, using a prediction <strong>of</strong> the future acceleration over a horizon equal to<br />

the time constant <strong>of</strong> the <strong>control</strong>led <strong>motion</strong> <strong>system</strong> can be used to compensate for the<br />

phase lag over the relevant frequency area without unnecessarily amplifying the high<br />

frequency area. The reference to the feedback linearising <strong>control</strong>ler would then be<br />

xr(t) =xd(t + ): (5.27)<br />

Also with this approach high frequency reference signal will not be compensated for<br />

properly, but at least will not be amplified.


5.6 Reference model <strong>based</strong> feed forward 183<br />

In either case, the connection <strong>of</strong> the host to the <strong>motion</strong> computer has to be such that<br />

the reference signal does not have a significant high frequency content. This is especially<br />

relevant since the host is operating at a relative low sample rate w.r.t. the <strong>motion</strong> computer.<br />

Caution has to be taken to prevent aliasing effects from mapping into this area. This will be<br />

discussed next.<br />

Host/Motion computer connection<br />

The host computer generates the <strong>simulator</strong> feasible trajectories on a sample frequency which<br />

is usually set at 60 Hz. In earlier days even 30 Hz was seen to be used. In near future, one<br />

is likely to be able to set this frequency at 120 Hz.<br />

According to Shannon’s theorem, reference signal frequencies higher than half the sample<br />

frequency, > 0:5fn, can not be discriminated from the area 0 :::0:5fn. In practice,<br />

aircraft model time constants should be well below this frequency. This alone motivates a<br />

higher sampling frequency as also flexible effects <strong>of</strong> large aircraft or rotorcraft need to be<br />

taken into account in simulation models in the future.<br />

Extensive literature [26] towards signal processing and filtering in the area <strong>of</strong> multirate<br />

signal processing exists. Since the <strong>motion</strong> <strong>system</strong> <strong>control</strong> computer requires (smooth)<br />

signals at 1-5 kHz, the lower-frequency sampled signals <strong>of</strong> the host computer will have to<br />

be interpolated. Ideally, if the signal x(k) to be interpolated would be known from time, t,<br />

in the interval ;1 to 1, filtering with the well known anti-causal filter h(t),<br />

xc(t) =<br />

=<br />

1X<br />

h(t ; k=fn)x(k)<br />

k=;1<br />

1X<br />

k=;1<br />

sin( (fnt ; k))<br />

x(k)� (5.28)<br />

(fnt ; k)<br />

would exactly reconstruct a continuous signal x c(t) with only frequency content in relevant<br />

area without any deformation in this area. In practice none (or at best only part) <strong>of</strong> the future<br />

is known; different techniques exist to approximate h(t). Every filtering technique then<br />

typically not only reduces high frequency signal contents, but also introduces phase lags.<br />

There are always shortcomings, which the designer must be aware <strong>of</strong>.<br />

Choice <strong>of</strong> polynomial interpolation technique<br />

In this research, several filtering techniques like Finite Impulse Response (FIR) filters, Infinite<br />

Impulse Response (IIR) filters and Cubic Polynomial Reconstruction (CPR), have been<br />

considered. Unlike FIR and IIR filters which are designed in the frequency domain, polynomial<br />

interpolation techniques can be designed in time domain. This has some advantages:<br />

time domain criteria can be used,<br />

non-linear design is possible,<br />

interpolation functions are continuous time functions, so they are independent <strong>of</strong> the<br />

interpolation factor.


184 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

The last point is considered important because in case <strong>of</strong> inter-connecting <strong>system</strong>s with<br />

large (> 50) ratios between their sampling rate (as in the case <strong>of</strong> host-<strong>motion</strong> computer<br />

integration), CPR interpolation techniques can still be implemented as low order filters.<br />

This is unlike the FIR-filters. Further, no redesign has to done as the interpolation factor<br />

changes, unlike the IIR-filter design.<br />

In polynomial interpolation on the interval t k to tk+1 <strong>of</strong> order p, a function<br />

f ( )=<br />

pX<br />

i=0<br />

ai i<br />

(5.29)<br />

is defined on the local time interval from 0 (tk) to1(tk+1). The coefficients ai <strong>of</strong> the<br />

polynomial are chosen such that the value <strong>of</strong> the signal at the sample time points, and<br />

possibly the time derivatives there<strong>of</strong>, satisfy certain constraints. With a third-order CPR,<br />

which was used to filter the acceleration signal, the four coefficients can for example be<br />

specified by defining the value f (0) = x(k ; 1) and f (1) = x(k) and their first time<br />

derivative _<br />

f(1) = x(k) ; x(k ; 1) and _<br />

f(0) = x(k ; 1) ; x(k ; 2). If a prediction <strong>of</strong> the<br />

signal x(k +1)is known. the interpolation can be shifted.<br />

With the position signal interpolation <strong>of</strong> a fifth-order function, (C)PR can be used to<br />

allow the accelerations (second order time derivatives) to correspond to the desired values<br />

also. In case signals from the host would arrive late, or not at all, precaution should be taken<br />

not to extrapolate the polynomials. This could run the signal <strong>of</strong>f-line very quickly. Here,<br />

by taking the next sample equal to the previous one for the position the <strong>system</strong> gradually<br />

comes to a stop.<br />

With CPR, the host and <strong>motion</strong> computer communicate the reference signals with a simple<br />

low-order method, which has relatively little high-frequency contents. The method can<br />

easily be combined with the use <strong>of</strong> the time-derivative <strong>of</strong> the acceleration for feed forward.<br />

This ’jerk’-signal can be extracted explicitly at every time point. Predicted time points can<br />

be incorporated directly to reduce the phase lag.<br />

Resume concerning the reference model <strong>based</strong> <strong>control</strong><br />

In this section a number <strong>of</strong> approaches to reduce time delays or lags in the response <strong>of</strong><br />

simulation <strong>motion</strong> <strong>system</strong>s were presented. A synergistic procedure was proposed in which<br />

<strong>system</strong> knowledge, <strong>of</strong> both the vehicle model to be simulated and the <strong>motion</strong> <strong>system</strong> to<br />

be accelerated, is used to have virtually immediate response to the desired <strong>motion</strong>. By a<br />

division <strong>of</strong> sub-tasks, the modular simulation structure can be maintained.<br />

Further research with a full operational <strong>simulator</strong> will be needed in order to point out<br />

exactly what delays are allowed while having no influence on the perceived realism <strong>of</strong> simulation.<br />

Having a <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> as the SRS, which is able to operate nominally<br />

with low latencies, will help enabling this research.<br />

Resume concerning the multi level <strong>control</strong> structure<br />

Motion <strong>control</strong> <strong>of</strong> a complex non-linear multivariable <strong>system</strong> like the Simona Research<br />

Simulator can be performed by using a <strong>control</strong> strategy in which the <strong>control</strong>ler is structured<br />

in multiple levels. Each level has its own specifications in close relation with the level it


5.7 Implementational issues 185<br />

communicates with. In this way a set <strong>of</strong> various <strong>control</strong> problems can be solved more or<br />

less separately. Further, the <strong>control</strong> structure enables implementation on a multi-processor<br />

dsp-<strong>system</strong>.<br />

5.7 Implementational issues<br />

The multiple level <strong>control</strong>ler has been implemented on a real-time multi-processor DSP<br />

<strong>motion</strong> computer [37] connected to the <strong>motion</strong> <strong>system</strong> with configuration C., a temporary<br />

dummy platform (ca. 4tons). In this set-up, as shown in Fig. 5.9, one C40-processor has to<br />

perform all communication with the outside world (bottle-neck w.r.t. sampling frequency)<br />

and could be run at 5 kHz. Also the coprocessor, which calculates the inner loop <strong>control</strong>,<br />

runs at 5 kHz, necessary to deal with the relevant fast actuator dynamics.<br />

In this respect the multiple level structure pays-<strong>of</strong>f since the other levels, especially the<br />

feedback linearising <strong>control</strong> ran on yet another coprocessor at 1 kHz which is just sufficient<br />

to go through all the algorithms involved.<br />

Design <strong>of</strong> the <strong>control</strong> structure was performed in the user friendly environment <strong>of</strong> Matlab/Simulink<br />

1 from which c-code can be generated automatically and connected with userwritten<br />

code in so called S-functions. The four processor c-code for the full model <strong>based</strong><br />

<strong>control</strong>ler consisted <strong>of</strong> about 9500 lines automatically generated code from the simulink<br />

model and 1850 lines <strong>of</strong> user written code (<strong>of</strong> which 600 lines came from the standard templates).<br />

Though probably not optimal in efficiency, a very user friendly environment has<br />

been enabled, which allows reasonable complex <strong>control</strong> structures to be implemented. In<br />

this way, rapid prototyping <strong>of</strong> complex <strong>control</strong>lers, as presented in this research, becomes<br />

feasible. Going from a Simulink model to a <strong>control</strong>ler running in a real time environment<br />

takes about 10 minutes.<br />

5.8 Performance quantification<br />

The evaluation procedure <strong>of</strong> the multiple level <strong>control</strong>ler implemented on the <strong>flight</strong> <strong>simulator</strong><br />

<strong>motion</strong> <strong>system</strong> SRS is presented in this section. The results, as partly also described<br />

in [73], were obtained through experimental measurements with the dummy platform with<br />

additional weights added, which amounted to the predicted final weight <strong>of</strong> 4 tons as in<br />

Fig. 1.5.<br />

The <strong>motion</strong>-base must be able to guarantee a high level <strong>of</strong> performance throughout its<br />

workspace. To do this, first a test procedure and a framework was defined whereby the<br />

dynamic characteristics could be quantified. A basis for the procedure is the long existing<br />

AGARD Advisory Report AR-144 [1], which quantifies several independent tests including<br />

the measurement <strong>of</strong> describing functions, dynamic thresholds, noise levels and the hysteresis.<br />

This set <strong>of</strong> tests gives insight into several linear and non-linear properties <strong>of</strong> a <strong>control</strong>led<br />

<strong>motion</strong> <strong>system</strong> in both the time and frequency domains. The AGARD tests had to be modified<br />

to include the high-frequency dynamics, and extended to enable characterisation <strong>of</strong><br />

1 Matlab and Simulink are registered trademarks <strong>of</strong> the MathWorks, Inc.


186 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Fig. 5.9: Motion <strong>control</strong>ler hardware setup: The model <strong>based</strong> <strong>control</strong>ler is implemented on<br />

a multi-C40-processor board. A master processor has to deal with all the I/O and<br />

forms the bottle-neck w.r.t. the necessary 5 kHz sampling time for the hydraulic<br />

loops. Apart from I/O the master only performs some safety checks on the outgoing<br />

signals. The two slave processors calculate through the multi level <strong>control</strong><br />

structure given in Fig. 5.1 On the first coprocessor (slave 1), the six inner loop<br />

<strong>control</strong>lers (module 1 <strong>of</strong> Fig. 5.1) are calculated at 5 kHz. They receive their reference<br />

(pressure) from the second coprocessor (slave 2), whose task is to process<br />

the mechanical model <strong>based</strong> <strong>control</strong> (modules 2 to 4 <strong>of</strong> Fig. 5.1). Finally, a third<br />

co-processor (slave 3), was used to generate the host signal (at 100 Hz).<br />

the <strong>system</strong> throughout its operating area. In order to arrive at a standardised approach, it is<br />

proposed that the latter be achieved by performing a set <strong>of</strong> path tracking tests. Furthermore,<br />

it was noticed that in order to enable an accurate characterisation <strong>of</strong> a high-performance<br />

<strong>motion</strong> <strong>system</strong>, both the experimental set up and test method have to be critically observed.<br />

Since the goal <strong>of</strong> the SRS simulation facility is to develop and validate new simulation<br />

technologies and to investigate human-machine interface design concepts, the performance<br />

demands on the <strong>motion</strong> cueing capability are higher than current devices. The high level <strong>of</strong><br />

performance has to be guaranteed throughout the workspace <strong>of</strong> the <strong>motion</strong> <strong>system</strong>.<br />

Qualifying or even quantifying the performance <strong>of</strong> a non-linear, multivariable, physical<br />

<strong>system</strong> is complex, especially if one wishes to compare the results with other configurations<br />

or <strong>simulator</strong>s. Ideally, first a model structure is chosen which describes the characteristics<br />

<strong>of</strong> such <strong>system</strong>s. Then, the parameters <strong>of</strong> this model structure are identified, which requires<br />

experiments or testing. Finally, the models are classified by assignment <strong>of</strong> a measure or cost<br />

function. The performance classification will at best produce results as good as the method<br />

and devices used to measure it. The identification <strong>of</strong> rigorous models, which describe all


5.8 Performance quantification 187<br />

<strong>system</strong>s characteristics, generally does not lead to accurate results [84]. Therefore, <strong>motion</strong><br />

<strong>system</strong>s are <strong>of</strong>ten classified by performing several tests to identify specific parameters,<br />

sometimes assigned within independent simplified model structures.<br />

In case <strong>of</strong> <strong>motion</strong> simulation, the objective cueing performance should preferably be<br />

measured by the ability to properly stimulate the human vestibular <strong>system</strong> [54]. This consists<br />

<strong>of</strong> two organs; the otoliths and the semicircular canals. Consider the perception models<br />

given in Section 1.2 and the parameters provided in Table 1.1. The otholiths are primarily<br />

and almost proportional sensitive to linear specific forces, the so-called ”regular” unit. The<br />

time derivatives <strong>of</strong> the specific forces, called the ”irregular” unit or ”jerk”, are also detected.<br />

The semicircular canals are proportionally sensitive to rotational acceleration in some frequency<br />

area (1:44 ; 30Hz) but are also proportional to rotational speed at lower frequency<br />

(0:026;1:44Hz). These sensitivities form the basis <strong>of</strong> the use <strong>of</strong> acceleration as the primary<br />

metric to measure the performance <strong>of</strong> a <strong>motion</strong> <strong>system</strong>.<br />

5.8.1 Motion <strong>system</strong> evaluation methods and requirements<br />

At present, only one standard method to characterise the performance <strong>of</strong> a <strong>motion</strong> <strong>system</strong><br />

is known to exist. This is described in the AGARD Advisory Report 144 [1]. Although this<br />

report dates back to the end <strong>of</strong> the seventies, the technique is still up to date. Additional<br />

aviation requirements, like JAR-STD-1A8 [118] and FAA 120-40C5 [3] generally, overlap<br />

AR-144 [1] and are basically less stringent. In robotics, even though rapid developments are<br />

underway in design and testing parallel robots, these have not led to a generally accepted<br />

characterisation standard. The AGARD report defines tests to measure<br />

- the describing function as a frequency domain evaluation,<br />

- dynamic threshold as the time domain response,<br />

- noise levels to characterise parasitic <strong>motion</strong>,<br />

- and hysteresis to identify hard non-linearities.<br />

It uses these parameters to describe the performance in one prescribed point within the<br />

workspace. As this is an international standard, which has been applied to several <strong>motion</strong><br />

<strong>system</strong>s [46], [123], it will be used as a basis in further expansion to evaluate the performance<br />

<strong>of</strong> our mechanism.<br />

The AGARD method does have some limitations.<br />

- First <strong>of</strong> all, it operates in only one point <strong>of</strong> the workspace, the neutral point. The dexterity<br />

in this point is <strong>of</strong>ten minimal [7], implying that the excursion forces are well within<br />

the operational limits <strong>of</strong> the <strong>motion</strong> <strong>system</strong>, and the limited non-linearity in this point<br />

implies that simple <strong>control</strong>lers can still perform reasonably well.<br />

- Secondly, the <strong>simulator</strong> upper-gimbal centroid is taken as the point <strong>of</strong> reference, but this<br />

point is almost never the pilot’s head reference location.<br />

- Furthermore, only a limited bandwidth <strong>of</strong> up to 10 Hz is tested in the AR-144 procedures.


188 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

- No evaluation <strong>of</strong> the limitations <strong>of</strong> the experimental set up is included. Especially in<br />

measuring the noise levels and hysteresis in high-performance <strong>motion</strong> <strong>system</strong>s, the<br />

discrimination between measurement noise and <strong>system</strong>-generated parasitic <strong>motion</strong><br />

should be possible.<br />

- Finally, the separate tests do not guarantee a level <strong>of</strong> performance <strong>of</strong> the <strong>system</strong> being<br />

operated during actual <strong>flight</strong> <strong>motion</strong> simulation. This evaluation is lacking in AR-<br />

144.<br />

To improve the performance characterisation this section will therefore propose some<br />

modifications to the original AGARD method in order to enable the calibration <strong>of</strong> highperformance<br />

<strong>motion</strong> <strong>system</strong>s. The newly proposed test should not be unnecessary extensive<br />

to remain economically feasible.<br />

The objective <strong>of</strong> this section on performance quantification is to characterise the performance<br />

<strong>of</strong> a high-fidelity <strong>flight</strong> simulation <strong>motion</strong> <strong>system</strong> throughout its workspace in an<br />

efficient manner. This can not be achieved with existing methods. Therefore, a first attempt<br />

is made to modify the test procedure in order to overcome the current limitations. For the<br />

sake <strong>of</strong> practicality and standardisation purposes, the experimental set up is considered to be<br />

part <strong>of</strong> the procedure. Application to the Simona <strong>motion</strong> <strong>system</strong> will lead to a preliminary<br />

characterisation and an evaluation <strong>of</strong> the procedure.<br />

Although tests exist which are independent <strong>of</strong> the <strong>system</strong> at hand, it is preferred to<br />

take the specific characteristics <strong>of</strong> a <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> into account in designing a<br />

procedure.<br />

- A <strong>motion</strong> <strong>system</strong> is a hydraulically or electrically-driven mechanism in which the un<strong>control</strong>led<br />

dynamics usually have resonance frequencies with a low level <strong>of</strong> damping.<br />

A test procedure should point out whether the degree <strong>of</strong> damping <strong>of</strong> the <strong>control</strong>led<br />

<strong>system</strong> is sufficient.<br />

- A purely mechanical <strong>system</strong> is considered to have immediate force generation and, as<br />

a result, direct acceleration. If the dynamics <strong>of</strong> the actuators is relevant, and this is<br />

certainly the case with hydraulics (no direct feed-through), then there will be a limited<br />

bandwidth to be quantified.<br />

- The use <strong>of</strong> a synergistic Stewart Platform implies that all actuators have to move to perform<br />

pure <strong>motion</strong> <strong>of</strong> one platform degree-<strong>of</strong>-freedom. Interaction in moving along<br />

the different degrees <strong>of</strong> freedom easily results if not compensated for. This should be<br />

taken into account in testing. Further, the interaction and reflected masses depend on<br />

the platform pose.<br />

- The hydraulic actuators are symmetric, enabling similar force generation in both directions.<br />

Asymmetricity can still occur if there is a pre-load, as is usually the case if<br />

gravitational forces exist. Hydrostatic bearings are applied to reduce friction, thereby<br />

reducing hysteresis, turn-around bump and parasitic <strong>motion</strong>. As a result, there are<br />

several sources <strong>of</strong> possible non-linear behaviour <strong>of</strong> the <strong>system</strong>, which should be considered.


5.8 Performance quantification 189<br />

In the design <strong>of</strong> a modernised test procedure an attempt will be made to take these<br />

characteristics into account.<br />

5.8.2 New test procedure<br />

In setting up the test procedure to describe the dynamic properties <strong>of</strong> a <strong>control</strong>led <strong>flight</strong><br />

<strong>simulator</strong> <strong>motion</strong> <strong>system</strong>, the AGARD Advisory Report 144 [1] is taken as a basis. In<br />

the following it is discussed, which part <strong>of</strong> AR-144 is considered important, what relevant<br />

modifications were made, and what extensions were taken into account.<br />

Describing function<br />

With the describing function method, the response <strong>of</strong> a <strong>system</strong> is given along two axes, in<br />

frequency and amplitude. As extensively discussed in Van Schothorst [124], the describing<br />

function method can be used to characterise a large class <strong>of</strong> non-linear <strong>system</strong>s in the<br />

frequency domain. It is stated there that hydraulically-driven mechanical <strong>system</strong>s belong to<br />

this class even if hard non-linearities such as Coulomb friction are relevant. If no effects<br />

like Coulomb friction are present, as should be the case with high performance <strong>motion</strong> <strong>system</strong>s<br />

with hydrostatic bearings, then the sinusoidal input describing function converges to<br />

the frequency response <strong>of</strong> the linear dynamics for small input signals. This assumption has<br />

to be checked, e.g. by the hysteresis and threshold tests described below.<br />

Considering the frequency response <strong>of</strong> a linear <strong>system</strong>, several properties can be deduced.<br />

Bandwidth can be calculated, stated as the -3 dB and/or the ;45 point. The ”peaking”<br />

or degree <strong>of</strong> damping <strong>of</strong> the <strong>control</strong>led <strong>system</strong> can be found, and cross talk can be<br />

analyzed. By taking a sinusoidal input amplitude <strong>of</strong> about 10 percent <strong>of</strong> the <strong>system</strong>’s positional,<br />

speed or acceleration limits, the non linear effects are observed to have a relatively<br />

minor influence.<br />

By using an analog frequency analyzer, a broad frequency spectrum can be taken into<br />

account and highly accurate frequency responses can be measured. This is considered to be<br />

preferable to the method described in AGARD in which an extensive number <strong>of</strong> discretefrequency<br />

measurements have to be taken. The limitations <strong>of</strong> this technique are however<br />

that one should measure within the (amplitude) range where the <strong>system</strong> can be considered<br />

linear, and that no higher harmonic response is taken into account.<br />

Threshold response<br />

With step response measurements, the response <strong>of</strong> the <strong>system</strong> can be analyzed in the time<br />

domain. It is considered an easily applicable test to infer the non-linearity <strong>of</strong> the <strong>system</strong><br />

with respect to different amplitudes. Originally, these kinds <strong>of</strong> tests were used to measure<br />

the lowest amplitude or threshold acceleration to which the <strong>system</strong> still responded. With<br />

modern <strong>motion</strong> <strong>system</strong>s having virtually no friction, the <strong>motion</strong>-base will respond to any<br />

amplitude and the problem at very low amplitude responses mainly results from the accuracy<br />

<strong>of</strong> the measurement apparatus.<br />

By considering a very simple first-order linear <strong>system</strong> response model with time delay,


190 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

acceleration (m/s2)<br />

position (m)<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

0.1<br />

0.05<br />

0<br />

0 0.5 1 1.5 2 2.5 3<br />

0 0.5 1 1.5<br />

time (s)<br />

2 2.5 3<br />

Fig. 5.10: Representative input signal, acceleration and position, for the dynamic threshold<br />

test.<br />

the following structure results:<br />

G(s) =<br />

K<br />

s +1 e- ds<br />

(5.30)<br />

This is <strong>of</strong>ten a convenient structure to approximate the <strong>motion</strong> <strong>system</strong> accelerational response.<br />

If e.g. a feedback linearisation and inner loop cascade-dp is applied, the <strong>motion</strong><br />

<strong>system</strong> acceleration roughly follows the first order presssure <strong>control</strong>led dynamics. The dynamics<br />

at higher frequencies and computational implementation mostly influences phase<br />

and can therefore <strong>of</strong>ten appropriately be described by one time delay.<br />

The parameters <strong>of</strong> this model are identified from the different step responses. The time<br />

delay, d, is the time that the <strong>system</strong> takes to respond, and the time constant, ,isgivenby<br />

the time taken from this point to reach 63% <strong>of</strong> the final value.<br />

With <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s, the acceleration step response is considered important<br />

for good cueing. To measure this response, a trajectory as given in Fig. 5.10 can<br />

be applied. One should notice that this cycled trajectory actually consists <strong>of</strong> eight different<br />

acceleration step responses. In this research, a response starting with zero velocity is<br />

considered.<br />

Path tracking<br />

To evaluate the <strong>system</strong> in its normal operating mode, and to analyze whether the properties<br />

measured at one operating point can be extended to a relevant part <strong>of</strong> the workspace, a new


5.9 Experimentally evaluating performance 191<br />

test is introduced. Advani [7] already introduced a set <strong>of</strong> benchmark manoeuvres to check<br />

the design <strong>of</strong> the kinematics <strong>of</strong> a <strong>motion</strong> <strong>system</strong>. These include standard manoeuvres during<br />

<strong>flight</strong> simulation training which are considered critical in the utilisation <strong>of</strong> the <strong>simulator</strong>.<br />

Some 30 manoeuvres, responses to heavy turbulence, rejected take-<strong>of</strong>f, landing, etc., are<br />

included. The ability <strong>of</strong> the <strong>control</strong>led <strong>system</strong> to track these manoeuvres is considered. At<br />

this moment, the mean and standard deviation <strong>of</strong> the acceleration error, after being compensated<br />

by the time delay, and the time delay itself are taken as the parameters to be measured.<br />

With this test, two aforementioned limitations <strong>of</strong> the AGARD procedure are accounted for.<br />

Further tests<br />

The AGARD procedure also includes noise levels and hysteresis tests. With high performance<br />

<strong>system</strong>s, the acceleration noise levels are very close to perceivable thresholds <strong>of</strong> 1%<br />

<strong>of</strong> gravity, and become hard to measure. This is because the measurement apparatus also<br />

introduces measurement noise, which should be discriminated from <strong>system</strong>-generated parasitic<br />

<strong>motion</strong>. With the gyro measurement taken with respect to rotation, the situation even<br />

becomes more severe as this signal has to be differentiated. With the hysteresis test, in<br />

which the <strong>system</strong> is made to track very low-frequency sinusoids, the situation is the same.<br />

It was concluded that at this moment these tests can (and should) only be used to check<br />

whether the <strong>system</strong>’s noise and hysteresis are below measurable thresholds e.g. to investigate<br />

if turn-around-bumps can be noticed.<br />

5.9 Experimentally evaluating performance<br />

In this section the results <strong>of</strong> the model <strong>based</strong> <strong>control</strong>ler will be evaluated using the proposed<br />

test method and with the reference <strong>of</strong> a conventional <strong>control</strong>ler with a decentralised<br />

pressure/position feedback for each actuator.<br />

5.9.1 Experimental set up<br />

As explained in the previous sections, the test procedure consists <strong>of</strong> several trials in which<br />

the <strong>control</strong>led <strong>motion</strong> <strong>system</strong> is made to perform a series <strong>of</strong> manoeuvres. During these<br />

tests, the response <strong>of</strong> the <strong>system</strong> has to be measured. First, the design and calibration <strong>of</strong> the<br />

experimental set up will be considered.<br />

Performance test set up<br />

The experimental configuration is given in Fig. 5.11. It consists <strong>of</strong> the Simona <strong>motion</strong><br />

<strong>system</strong>, which is <strong>control</strong>led by the <strong>motion</strong> computer. With the actuator extensions being<br />

measured, together with the kinematics <strong>of</strong> the <strong>system</strong>, the platform pose (position and orientation)<br />

can be calculated on-line. In fact the <strong>system</strong>’s response with respect to the accelerations<br />

can also be calculated from these measurements. However, as the signals then<br />

need to be differentiated twice, the high-frequency component <strong>of</strong> the reconstructed acceleration<br />

becomes corrupted by measurement noise. Therefore, an independent data acquisition<br />

<strong>system</strong> is used to measure the <strong>system</strong>s accelerating response.


192 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Fig. 5.11: Motion <strong>control</strong> evaluation test setup: Independent <strong>of</strong> the <strong>motion</strong> <strong>control</strong> <strong>system</strong><br />

given in Fig. 5.9 implemented on the <strong>motion</strong> computer, the accelerations and<br />

angular velocities <strong>of</strong> the <strong>system</strong> are measured at the <strong>motion</strong> platform by another<br />

data acquisition <strong>system</strong> (also a dSpace computer). The <strong>motion</strong> <strong>control</strong>ler and<br />

data acquisition <strong>system</strong> are synchronised by a frequency generator which is used<br />

to trigger the start <strong>of</strong> each run and detect possible discrepancies between the<br />

<strong>system</strong> clocks.<br />

With a reference measurement package consisting <strong>of</strong> three accelerometers and three<br />

rate gyros attached to the dummy platform, the local specific forces and angular velocities<br />

are measured. This still requires one differentiation step in order to calculate the angular<br />

acceleration.<br />

As measurements <strong>of</strong> the <strong>motion</strong> computer and the data acquisition computer will be<br />

used in the evaluation <strong>of</strong> the test procedures, the time scale <strong>of</strong> both <strong>system</strong>s has to be synchronised.<br />

To achieve this, an external reference signal from a frequency generator was fed<br />

to both <strong>system</strong>s. This sinusoid was used to trigger the start <strong>of</strong> each run. The test setup is<br />

depicted in Fig. 5.11.<br />

Calibration measurement <strong>system</strong><br />

Reconstruction <strong>of</strong> the actual acceleration <strong>of</strong> the platform in all six degrees-<strong>of</strong>-freedom cannot<br />

be done with the measurements taken by the reference package alone. Some additional<br />

information has to be known in order to process the data properly. The gain and orientation<br />

<strong>of</strong> the measurement devices, and also their relative position, are required. Some (but not all)<br />

<strong>of</strong> these parameters can be calibrated by <strong>of</strong>f-line measurements. As part <strong>of</strong> this research,


5.9 Experimentally evaluating performance 193<br />

another approach was taken.<br />

In the low-frequency region, the accelerations <strong>of</strong> the platform can be reconstructed with<br />

the actuator extensions and the kinematics <strong>of</strong> the platform as also done in Section 4.2.2.<br />

By performing a manoeuvre at a specific frequency (e.g. at f=4 Hz), a high signal-to-noise<br />

ratio can be obtained. Especially if only this component <strong>of</strong> the signal is taken into account,<br />

the non-linear components can be neglected. This frequency should be chosen such that<br />

both the accelerometers and the position transducers have an approximately equal signalto-noise<br />

ratio, taking into account the accuracy and resolution <strong>of</strong> the measurement devices.<br />

Manoeuvring in six independent platform directions assures spanning the whole <strong>motion</strong><br />

space. Here, we assume only rigid-body <strong>motion</strong> <strong>of</strong> the platform, and a rigidly-attached<br />

base. This assumption is not valid in the case <strong>of</strong> the current placement <strong>of</strong> the Simona<br />

<strong>motion</strong> <strong>system</strong> (on the floating concrete floor plate) at frequencies higher than 8 Hz.<br />

The position and orientation <strong>of</strong> the reference package in platform coordinates can be reconstructed<br />

taking the procedure presented in the previous chapter (4.46) for the accelerometers<br />

used in identifying the flexible dynamics. The identification was improved further by<br />

incorporating the dynamics <strong>of</strong> the anti-aliasing filters.<br />

This procedure was applied to the reference package attached to the dummy platform<br />

moved by the Simona <strong>motion</strong> <strong>system</strong>. The accelerometers and gyros appeared almost perfectly<br />

aligned with the axes <strong>of</strong> the moving reference frame. The gains also corresponded<br />

to the earlier <strong>of</strong>f-line calibrated values (approx. 3.6V/g). The position <strong>of</strong> the lines <strong>of</strong> measurement<br />

for the accelerometers were also very much the same with respect to their relative<br />

design positions. The absolute position <strong>of</strong> the reference package was identified at<br />

683mm 1mm below the centre <strong>of</strong> gravity (which is 233mm below centre upper gimbal<br />

point). The x-y coordinates (forward and to the side) depend on the accelerometer at hand<br />

which should be compensated for in the following test procedure. As this easy to apply<br />

procedure was successful, it could be used in the future to reconstruct the relative position<br />

in the interior <strong>of</strong> the <strong>simulator</strong>.<br />

5.9.2 Characteristics <strong>of</strong> the Simona <strong>motion</strong> <strong>system</strong><br />

Given the description <strong>of</strong> the test procedure and the experimental set up, the resulting performance<br />

characteristics <strong>of</strong> the Simona <strong>motion</strong> <strong>system</strong> will be outlined.<br />

Frequency response<br />

First, the application <strong>of</strong> the describing function test will be discussed. In Fig. 5.12, both<br />

the amplitude and phase frequency characteristics <strong>of</strong> the Simona <strong>motion</strong> <strong>system</strong> with the<br />

multiple level <strong>control</strong>ler from 0.5 to 50 Hz are shown. All plots are separately shown in<br />

Fig. A.9 and Fig. A.10 <strong>of</strong> Appendix A with direct comparison <strong>of</strong> the model <strong>based</strong> <strong>control</strong>ler<br />

to a conventional <strong>control</strong>ler and in Fig. A.11 <strong>of</strong> Appendix A also the error response <strong>of</strong> both<br />

<strong>control</strong>lers to reference signals is given.<br />

It can be seen that the -3 dB point can be found at 13 to 15 Hz for the ’non horizontal’<br />

directions <strong>of</strong> pitch (omy), roll (omx) and heave (z). In surge (x), sway (y) and yaw (om z),<br />

the bandwidth is lower due to the <strong>motion</strong> <strong>of</strong> the current floor foundation. Except for yaw,


194 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Direction Response time (ms) Direction Response time (ms)<br />

Surge 32 Roll 35<br />

Sway 32 Pitch 35<br />

Heave 33 Yaw 34<br />

Table 5.1: The Simona Research Simulator latency with a model <strong>based</strong> <strong>control</strong>ler.<br />

the ;45 bandwidth can be found at approx. 5 Hz. The peak amplitudes remain well below<br />

2 dB. The most relevant parasitic <strong>motion</strong> is measured between pitch due to surge and vice<br />

versa, and roll due to sway and vice versa.<br />

For reference purposes, a conventional approach was used to <strong>control</strong> the same <strong>motion</strong><br />

<strong>system</strong>. The frequency response is given in Fig. 5.13. The same pressure feedback gain<br />

is applied, however, since there is no compensation for the different natural frequencies,<br />

some responses, like surge and sway, demonstrate peaks much more than others (like pitch<br />

and roll, which are over damped). The bandwidth is twice as low, and the peaking and<br />

interaction are twice as high with this <strong>control</strong> approach. The describing function enables<br />

the characterisation <strong>of</strong> some <strong>of</strong> the parameters, which are considered most important in<br />

<strong>control</strong>, like bandwidth, damping and interaction in a multivariable <strong>system</strong>.<br />

Threshold response<br />

Threshold tests have been performed for all platform directions for four different amplitudes<br />

in acceleration (1, 0.4, 0.1, 0.05 m/s2) for both negative as positive directions. The response<br />

in the surge direction is shown in Fig. 5.14. Note that the signal amplitudes have been scaled<br />

to a desired final value <strong>of</strong> 1. As can be seen the response becomes less regular for the lower<br />

amplitudes, but the time delay (approx. 10 ms) and time constant (approx. 22 ms) do not<br />

vary too much and correspond with the ;45 bandwidth ( =1=(2 fbandwidth)) measured<br />

with the frequency response. Also, the smooth response without too much peaking was<br />

predicted by the frequency response.<br />

The characteristic parameters for the other platform directions are given in Table 5.1. In<br />

all cases, a time delay <strong>of</strong> about 10 ms can be observed. Given the sample frequency <strong>of</strong> the<br />

host processor <strong>of</strong> 100 Hz, such values can be expected. It should be noted that the responses<br />

in the rotational directions were much more corrupted by measurement noise, as the signal<br />

had to be differentiated. Especially for the low amplitude responses <strong>of</strong> 0.05 and 0.1 rad/s2,<br />

no exact time constant or delay could be extracted.<br />

Path tracking<br />

To check whether the characteristics as measured in one operating point would still be preserved<br />

while performing relevant <strong>flight</strong> manoeuvres, the path tracking test was introduced.<br />

Testing Characteristic Manoeuvres<br />

At the KLM <strong>flight</strong> crew training centre, 31 <strong>simulator</strong> training-critical manoeuvres were<br />

flown in a Boeing 747-400 <strong>simulator</strong> and the aircraft model responses registered. These


5.9 Experimentally evaluating performance 195<br />

Amplitude m/s2−>m/s2<br />

Phase (deg)<br />

10 0<br />

10 −1<br />

0<br />

−50<br />

−100<br />

−150<br />

10 0<br />

10 0<br />

om_z<br />

z<br />

x y<br />

Frequency (Hz)<br />

om_y<br />

om_x<br />

om_z<br />

Fig. 5.12: Actual model <strong>based</strong> <strong>control</strong> <strong>motion</strong> <strong>system</strong> frequency response. The Bode<br />

plot <strong>of</strong> the closed loop <strong>system</strong> from required to actual (translational and rotational)<br />

accelerations applying the model <strong>based</strong> <strong>control</strong>ler in the neutral position.<br />

The most important interactions are also given (dashed) and rotations are multiplied<br />

by the upper gimbal radius. Highest bandwidths are attained with pitch<br />

(omy), roll (omx) and somewhat lower with heave (z). For the directions surge<br />

(x), sway (y) and yaw (omz), where the foundation is flexible and not included<br />

in the model used to design the <strong>control</strong>ler, some energy does not accelerate the<br />

platform above 4 Hz. The phase <strong>of</strong> the yaw direction does lag the others. Maybe<br />

the gyro filter was not compensated for properly. All responses are reasonably<br />

flat without peaking more than a few percent above 10 0 . Interaction, especially<br />

between pitch to surge and roll to sway and vice versa is unexpectedly high up<br />

till 30 % above 4Hz. The six by six Bode response plots directly comparing the<br />

model <strong>based</strong> and conventional <strong>control</strong>ler can be found in Fig. A.9 and Fig. A.10<br />

<strong>of</strong> Appendix A<br />

10 1<br />

10 1


196 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Amplitude m/s2−>m/s2<br />

Phase (deg)<br />

10 0<br />

10 −1<br />

0<br />

−50<br />

−100<br />

−150<br />

−200<br />

10 0<br />

10 0<br />

y<br />

x<br />

x<br />

y om_z<br />

z<br />

om_z<br />

Frequency (Hz)<br />

10 1<br />

10 1<br />

om_y<br />

om_x<br />

Fig. 5.13: Actual conventionally <strong>control</strong>led <strong>motion</strong> <strong>system</strong> frequency response. The<br />

Bode plot <strong>of</strong> the closed loop <strong>system</strong> from required to actual (translational and rotational)<br />

accelerations applying the basically conventional <strong>control</strong> strategy in the<br />

neutral position. The most severe interactions, between pitch (om y) and surge (x)<br />

and roll (omx) and sway (y) are given (dashed) and the rotations are multiplied<br />

by the upper gimbal radius. As expected, with the conventional <strong>control</strong>ler it is not<br />

possible to tune the bandwidths <strong>of</strong> the platform accelerational loops w.r.t. each<br />

other. Further, a compromise has to be found between peaking (up till 6 dB)<br />

<strong>of</strong> the lowest bandwidth loops <strong>of</strong> surge (x) and sway (y) and the overdamped<br />

response <strong>of</strong> pitch (omy) and roll (omx). The 4 Hz bandwidth <strong>of</strong> surge (x) and<br />

sway (y) is also a maximum attainable bandwidth using this <strong>control</strong> structure, the<br />

amount <strong>of</strong> peaking given and the lowest eigenfrequency <strong>of</strong> the platform. The six<br />

by six Bode response plots directly comparing the model <strong>based</strong> and conventional<br />

<strong>control</strong>ler can be found in Fig. A.9 and Fig. A.10 <strong>of</strong> Appendix A


5.9 Experimentally evaluating performance 197<br />

1.5<br />

1<br />

0.5<br />

0<br />

−0.5<br />

−1<br />

−1.5<br />

+/−1.0 m/s2<br />

+/−0.4 m/s2<br />

+/−0.1 m/s2<br />

0.05 m/s2<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14<br />

Fig. 5.14: Short response times <strong>of</strong> the model <strong>based</strong> <strong>control</strong>ler <strong>of</strong> a large range <strong>of</strong> accelerational<br />

amplitudes. The normalised threshold or acceleration step response<br />

measurements are given over a range <strong>of</strong> amplitudes for the most difficult surge<br />

direction.<br />

manoeuvres included a wide range <strong>of</strong> dynamic conditions, including take-<strong>of</strong>fs, normal and<br />

hard landings, engine seizures, and response to heavy turbulence. These manoeuvres were<br />

used for research into kinematical design <strong>of</strong> <strong>motion</strong> platforms by Advani [7] and reference<br />

model <strong>based</strong> feedforward <strong>control</strong> within a Masters project as part <strong>of</strong> this research by Piatkiewitz<br />

[111].<br />

It was decided to use a set <strong>of</strong> these manoeuvres, which were dynamically most challenging,<br />

as a benchmark for the SRS <strong>motion</strong> platform. Five <strong>of</strong> the thirty-one manoeuvres were<br />

selected, since these represented the most <strong>motion</strong>-demanding conditions.<br />

These were:<br />

1. Response to maximum clear air turbulence<br />

The <strong>motion</strong> <strong>system</strong> is made to move in a stochastic manner experiencing vibrations<br />

up to 10 Hz. Also an airpocket is included which requires almost full speed <strong>of</strong> the<br />

actuators (:9 m/s with vmax=1 m/s).<br />

2. Taxiing<br />

Fast vibrations introduced through the landing gear have to be experienced in <strong>simulator</strong><br />

during which it has to move gradually through surge, sway and yaw <strong>motion</strong>.<br />

3. Landing with cross wind<br />

In the air, the <strong>simulator</strong> has to experience the stochastic nature <strong>of</strong> cross wind. At


198 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

landing three instances <strong>of</strong> landing gear touch down bumps require fast acceleration<br />

peak generation after which the landing slide introduces even faster vibrations up to<br />

15 Hz.<br />

4. Rejected take <strong>of</strong>f<br />

Immediate negative surge together with tilt coordination (in which gravity is made<br />

to introduce long term breaking force) move the <strong>simulator</strong> through a wide area <strong>of</strong><br />

operating conditions w.r.t the platform pose.<br />

5. Rotation during take <strong>of</strong>f roll<br />

Limited amount <strong>of</strong> <strong>motion</strong> also has to be introduced smoothly in the asymmetric<br />

(roll) <strong>motion</strong>. Faster resonances (before take <strong>of</strong>f) during the first seconds have to be<br />

replaced by the more smooth vibrations in flying.<br />

To test the <strong>system</strong> with the load conditions <strong>of</strong> a full operational SRS the <strong>control</strong> strategy<br />

proposed was evaluated with a dummy platform <strong>of</strong> 4000 kg on top <strong>of</strong> the SRS <strong>motion</strong> <strong>system</strong><br />

depicted in Fig. 1.4. The host computer was replaced by an additional processor on the<br />

<strong>motion</strong> computer running at a lower sampling rate <strong>of</strong> 100 Hz. In the ’host’-processor the<br />

tracks for platform pose and acceleration are made available and send to the <strong>motion</strong> computer<br />

<strong>control</strong> processors at demand. The original tracks were aircraft model output. By a<br />

plain wash out these trajectories were made fit for the SRS. Actuator stroke had to be within<br />

1.25 m and velocity within 1 m/s. To test the <strong>motion</strong> <strong>system</strong> with the <strong>control</strong> strategy these<br />

signals were considered the feasible trajectories which had to be tracked as close as possible.<br />

Evaluation <strong>of</strong> the path tracking<br />

Each manoeuvre has a duration <strong>of</strong> 16 seconds, however, due to the limited hydraulic power<br />

supply (temporarily) available at the Central Workshop <strong>of</strong> Mechanical Engineering this duration<br />

is just manageable. In Fig. 5.15, the measured heave response together with the<br />

reference acceleration, are depicted.<br />

The measured heave signals correspond very closely to the desired acceleration. If it<br />

is considered on a short time frame an almost constant lag <strong>of</strong> approximately 30 m can be<br />

observed. The main contribution <strong>of</strong> the lag in response is due to the smoothing filter at 30 Hz<br />

(63 % rise time <strong>of</strong> 20 ms). This motivates the use <strong>of</strong> a prediction horizon <strong>of</strong> 20-30 ms, in<br />

which case the lag in response could very well be reduced to a situation <strong>of</strong> ’virtual zero time<br />

delay’. In this chapter a number <strong>of</strong> approaches to achieve this were presented.<br />

In Fig. 5.15 part <strong>of</strong> the response during a landing with cross wind is shown. Acceleration<br />

peaks can be observed as the three landing bumps occur. These are well represented by the<br />

<strong>control</strong>led <strong>motion</strong> <strong>system</strong>. The frequency content <strong>of</strong> the signal is clearly changing as the<br />

wheels start to roll over the landing track (more high frequency rumble required). The<br />

errors are well below 2% <strong>of</strong> gravity. As the high-frequency part <strong>of</strong> the error signal can be<br />

significant (due to measurement noise and also anti-alias filtering, starting at 30 Hz), it is<br />

hard to draw conclusions with respect to this frequency region.<br />

The figure shows that the <strong>motion</strong> <strong>system</strong> reacts without gain decrease on the stochastic<br />

nature <strong>of</strong> wheel landing track contact at least up to vibrations <strong>of</strong> ca. 8 Hz (which can be seen<br />

during the time span 8.05 s:::8.45 s). The ’noise’ in the measured signal is mainly caused


5.9 Experimentally evaluating performance 199<br />

acceleration (m/s2)<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

−3<br />

−4<br />

landing with cross wind, desired/actual heave<br />

−5<br />

0 2 4 6 8 10<br />

time (s)<br />

12 14 16 18 20<br />

Fig. 5.15: Path tracking in landing with cross wind manoeuvre with the model <strong>based</strong> <strong>control</strong>ler<br />

considering accelerations in the platform heave (z-direction). Upper plot<br />

gives the desired accelerations in heave direction and the lower one gives the<br />

actually measured accelerations (<strong>of</strong>f-set 2m=s 2 brought in on purpose for comparison)<br />

by the application <strong>of</strong> servo valve dither. As said, the top-top values remain however well<br />

below 0.02 g, which almost unnoticeable.<br />

The other manoeuvres, not depicted here, show comparable characteristics. This proves<br />

that this <strong>motion</strong> <strong>system</strong> <strong>control</strong> and design strategy resulted in a high-quality <strong>system</strong>. The<br />

use <strong>of</strong> a predictive reference model <strong>based</strong> <strong>control</strong> will even enable a synergistic connection<br />

between host and <strong>motion</strong> computer, which will considerably reduce latencies.<br />

The two aforementioned <strong>control</strong>lers, (1) the multiple level <strong>control</strong>ler and (2) the conventional<br />

<strong>control</strong>ler, were compared using the manoeuvres. The characteristic parameters,<br />

given in Table 5.2, show again that the multiple level <strong>control</strong>ler has higher performance with<br />

respect to the conventional one.<br />

Further tests<br />

Further, as already pointed out, noise level and hysteresis tests are also part <strong>of</strong> this procedure.<br />

As both were near or below measurable or noticeable levels, no exact values can be<br />

given here. The hysteresis tests showed values less than 0.3 mm. This value was measured<br />

going through a sinusoid with 0.4 m <strong>of</strong> amplitude and period time <strong>of</strong> 5 minutes, and was<br />

still decreasing as the period time was increased. It was observed that higher values <strong>of</strong> hys-


200 5 <strong>Model</strong> <strong>based</strong> <strong>control</strong> <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong><br />

Manoeuvre Control Delay error Mean error Std error<br />

(ms) (m=s 2 ) (m=s 2 )<br />

Landing <strong>Model</strong> <strong>based</strong> 32 0.032 0.037<br />

Conventional 62 0.042 0.041<br />

Taxiing <strong>Model</strong> <strong>based</strong> 34 0.022 0.019<br />

Conventional 62 0.041 0.033<br />

Table 5.2: Comparison path tracking results in heave (delay, mean error, standard deviation<br />

(std)).<br />

teresis appear if the dither - a high-frequency vibration in the pilot valve <strong>of</strong> the actuator - is<br />

not properly tuned. Therefore, this test should be part <strong>of</strong> the procedure to guarantee some<br />

upper bound.<br />

The noise levels measured with respect to the translational acceleration are acceptable,<br />

though close to measurable levels <strong>of</strong> 0.01 to 0.02 g. It should be noted that with the experimental<br />

set-up used, no predictions with respect to high-frequencies ( 50 Hz) can be<br />

given. The noise levels in the rotational directions could not be measured as there was too<br />

much measurement noise caused by differentiating the angular velocity. Therefore, more<br />

accurate gyros should be used here. The frequency responses measured in Fig. 5.12 do not,<br />

however, predict higher sensitivity in the rotational directions compared to the translations.<br />

5.10 Chapter Resume<br />

Almost invariantly, in every test performed, the model <strong>based</strong> <strong>control</strong>ler did a far better job<br />

than the reference conventional <strong>control</strong>ler. Higher bandwidths (8 Hz for 90 phase lag)<br />

were measured in the frequency domain along with less peaking (max: 2 dB). Further,<br />

higher suppression <strong>of</strong> interaction was also attained, though somewhat less than expected. In<br />

the time domain, fast responses (response time between 30 and 35 ms) were observed over<br />

a large range <strong>of</strong> amplitudes including 5 mg.<br />

In the conventional <strong>control</strong>ler, a compromise had to be settled for in which some loops<br />

have high bandwidth (still not more than 4 to 5 Hz) and too severely damped response<br />

and other loops show already relatively high peaking up till 6 dB. Only in heave direction,<br />

in the neutral position, performance is almost equal to the model <strong>based</strong> <strong>control</strong> approach.<br />

Fortunately, for the conventional <strong>control</strong> approach, most accelerations in the simulated manoeuvres<br />

were required along the heave direction.<br />

In this chapter also a test procedure was proposed to characterise the performance <strong>of</strong> a<br />

high-fidelity <strong>flight</strong> simulation <strong>motion</strong> <strong>system</strong> throughout its workspace in an efficient manner.<br />

The procedure was evaluated with the Simona <strong>motion</strong> <strong>system</strong>, which, although in preliminary<br />

form, already shows favourable properties due to its design and the model <strong>based</strong><br />

<strong>control</strong> approach. The main parameters <strong>of</strong> such a <strong>system</strong> appear to be given by its linear<br />

dynamics. Measuring the frequency response is an efficient way to characterise this. Other<br />

tests such as threshold step responses, path tracking and hysteresis, can be used to quantify


5.10 Chapter Resume 201<br />

to what extent these properties are maintained throughout the full operating area. Further<br />

research has to be done to identify the noise levels with higher accuracy.<br />

Some modifications to the current standard outlined in AGARD-144 can yield a better<br />

quantification <strong>of</strong> high-performance <strong>motion</strong> <strong>system</strong>s.<br />

Performance <strong>of</strong> the <strong>motion</strong> <strong>system</strong> was defined by the degree <strong>of</strong> <strong>motion</strong> realism attained.<br />

There are no measures known which exactly quantify this. More research into human perception<br />

has to point out how this has to be done. The Simona Research Simulator could<br />

play a role in attaining this goal.


Chapter 6<br />

Review and discussion on the<br />

results<br />

The previous chapters went over the full <strong>system</strong>/<strong>control</strong> design process <strong>of</strong> <strong>flight</strong> <strong>simulator</strong><br />

<strong>motion</strong> <strong>system</strong>s. Before drawing the final conclusions towards this research, the main<br />

considerations in this process will be discussed.<br />

It has been investigated to what extent the use <strong>of</strong> relevant <strong>system</strong> knowledge in the <strong>motion</strong><br />

<strong>control</strong> strategy improves the <strong>control</strong>led dynamics <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>.<br />

This required structuring the research in several subproblems. Defining quality i.e. setting<br />

specifications, obtaining <strong>system</strong> knowledge through theoretical modelling, quantifying and<br />

verifying these models and the relevant dynamics by experiments. Definition <strong>of</strong> a model<strong>based</strong><br />

but implementable <strong>control</strong> strategy and a test to validate and compare the obtained<br />

closed loop dynamics with conventionally <strong>control</strong>led <strong>motion</strong> <strong>system</strong>s. In literature various<br />

solutions to many <strong>of</strong> these subproblems are proposed. Integratibility in the full design<br />

scheme and actual applicability have been the main arguments in this research in the choice<br />

for the most suitable available alternative or, in some cases, newly proposed variant.<br />

6.1 Flight simulation<br />

Flight simulation is about creating a complex environment, which can be used to train pilots,<br />

evaluate <strong>flight</strong> <strong>system</strong> characteristics, etc. in a <strong>system</strong> which is not actually flying.<br />

Using a <strong>flight</strong> <strong>simulator</strong> is cheaper, safer, environmentally less harmful than flying an aircraft<br />

and training or evaluation can more easily be defined, standardised, modularised and<br />

repeated. Since these advantages become more and more important, there is a continuing<br />

market driven tendency towards the use <strong>of</strong> <strong>flight</strong> <strong>simulator</strong>s. However, requiring higher standards<br />

in simulation quality. Major breakthrough in technology, mainly due to the increase<br />

in computer power, concern the visual <strong>system</strong> and the ability to take complex <strong>flight</strong> characteristics<br />

into account in the aircraft (environment) modelling. Motion is one the stimuli<br />

which can never be perfect given the limited dexterity <strong>of</strong> <strong>motion</strong> <strong>system</strong>s kinematics. This<br />

foremost constrains the ability to duplicate the low frequency part <strong>of</strong> manoeuvres. This can<br />

partly be compensated for by current visual <strong>system</strong>s. The attainable quality <strong>of</strong> simulating<br />

203


204 6 Review and discussion on the results<br />

high frequent vibration or onset <strong>motion</strong> is determined by the <strong>control</strong>led dynamics <strong>of</strong> the<br />

<strong>motion</strong> <strong>system</strong> together with the simulation model. By design <strong>of</strong> the Simona <strong>motion</strong> <strong>system</strong><br />

construction, expectedly favourable properties such as low mass and inertia, low centre <strong>of</strong><br />

gravity and high rigidity were strived for. Design <strong>of</strong> a suitable <strong>control</strong>ler to exploit this, is<br />

the next step. This has been investigated in this research.<br />

6.2 Motion <strong>system</strong> specifications<br />

Perceived <strong>motion</strong> or realism <strong>of</strong> <strong>motion</strong> simulation is a subjective measure which can not<br />

exactly be specified. Research into <strong>motion</strong> perception models is one <strong>of</strong> the arguments in<br />

trying to attain a relatively high performance <strong>motion</strong> <strong>system</strong>. Flight <strong>simulator</strong> users explicitly<br />

ask for quantification <strong>of</strong> required <strong>motion</strong> cues and accuracy necessary to attain a certain<br />

training quality. Literature provides for quite some indications in what range specifications<br />

should be tighter. First <strong>of</strong> all, predictable characteristics <strong>of</strong> the <strong>control</strong>led <strong>motion</strong> <strong>system</strong><br />

over its full working area should be strived for to be able to perform well defined <strong>simulator</strong><br />

training experiments. This alone motivates the use <strong>of</strong> as much relevant <strong>system</strong> knowledge<br />

as possible.<br />

Control specifications should result from the requirements set, the internal structure <strong>of</strong><br />

the <strong>system</strong> to be <strong>control</strong>led and the properties <strong>of</strong> the external signals which are expected<br />

to perturb this <strong>system</strong>. A fundamental difference in <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s w.r.t. usual<br />

robotics, is the fact that acceleration instead <strong>of</strong> positional accuracy is most important. In the<br />

low frequency range (preferably below :1 Hz) position stabilisation to prevent the actuators<br />

running out <strong>of</strong> stroke should prevail, but as from this frequency on, acceleration should<br />

be tracked. Since the <strong>simulator</strong> is running free, tracking <strong>of</strong> acceleration reference signals<br />

is the major task, which has to be compromised with unmodelled dynamics, uncertainties<br />

and measurement noise. Disturbance rejection is less evidently necessary. If the <strong>system</strong> is<br />

known, feed forward tracking can be used. As uncertainty grows, feedback will have to be<br />

applied.<br />

Although there is no difference in the ratio between desired and attained position or<br />

acceleration, errors or sensitivity to external signals are to be weighted with the squared<br />

frequency (! 2 ) if one evaluates accelerational w.r.t. positional accuracy in the frequency<br />

domain. This means sensitivity to noise is much more important in <strong>simulator</strong>s than it is in<br />

robotics. As a <strong>system</strong> is typically sensitive to noise around the bandwidth, it will be more<br />

difficult to obtain a high bandwidth <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>. Use <strong>of</strong> <strong>system</strong> knowledge<br />

through model-<strong>based</strong> <strong>control</strong>, can prevent from unnecessary amplification <strong>of</strong> noise.<br />

In the usual setting, the <strong>control</strong>led <strong>motion</strong> <strong>system</strong> is in serial closed loop connection<br />

with the aircraft model and pilot. Both the aircraft and pilot are not exactly specified, however,<br />

an aircraft typically has its highest relevant frequency modes at 2-5Hz and pilots will<br />

close the loop in order to obtain a bandwidth <strong>of</strong> .3 to 1Hz. This is attained by providing<br />

lead over a decade around this frequency (so up till 3 Hz). To have minimal influence on<br />

the pilot-aircraft(-model) loop, phase lag <strong>of</strong> the <strong>control</strong>led <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> should<br />

be virtually zero in this frequency area, which otherwise would influence relative stability<br />

<strong>of</strong> this loop. This requires a bandwidth <strong>of</strong> an order <strong>of</strong> a magnitude higher ( 10 Hz).<br />

The frequency response <strong>of</strong> an aircraft evaluated w.r.t. position can be misleading as this


6.3 Theoretical modelling 205<br />

response quickly drops above the highest mode while acceleration response remains constant.<br />

Apart from these loop characteristics, realism might also be improved by simulation<br />

<strong>of</strong> vibrations well above these frequencies such as turbulence, rotorcraft resonance, taxiing<br />

wheel to ground rumble and impact on landing bump. Amplitude characteristics should be<br />

unchanged in simulation <strong>of</strong> these kind <strong>of</strong> manoeuvres.<br />

Finally the multivariable <strong>system</strong> should not introduce parasitic coupling in the other<br />

degrees <strong>of</strong> freedom and as this depends on the relative position, the response in each degree<br />

<strong>of</strong> freedom should be similar.<br />

To summarise, a <strong>control</strong>led <strong>motion</strong> <strong>system</strong> should preferably be able to simulate references<br />

with frequency content up to 1 Hz with virtually zero time delay and an order <strong>of</strong><br />

a magnitude higher without amplitude attenuation. Use <strong>of</strong> <strong>system</strong> knowledge can help in<br />

attaining a predictable response without unnecessary noise amplification.<br />

6.3 Theoretical modelling<br />

<strong>Model</strong>ling is the first step towards model <strong>based</strong> <strong>control</strong> but also helps in analysing the characteristics<br />

<strong>of</strong> the open loop <strong>system</strong>, the response <strong>of</strong> the conventionally <strong>control</strong>led <strong>system</strong><br />

and picking the relevant dynamics. Theoretic physical modelling enables analysis before<br />

actually building a <strong>system</strong> and helps in attaining an experimentally fitted model <strong>of</strong> limited<br />

complexity describing the relevant dynamical features and pointing at the physics from<br />

which these features result. In describing the dynamics <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>,<br />

the most relevant dynamics can be found in the mechanical <strong>system</strong>, hydraulic drives and the<br />

connection between the two.<br />

The mechanical <strong>system</strong> is, unlike the usual robotic manipulators, a fully parallel <strong>system</strong>.<br />

This leads to some dual properties if compared with the serial connected <strong>system</strong>s. In<br />

singular points, the manipulator becomes unsupported in one or more degrees <strong>of</strong> freedom,<br />

while in serial connected <strong>system</strong>s these become fixed. Kinematically, the parallelism poses<br />

problems in describing the position as an explicit function <strong>of</strong> the actuated coordinates, while<br />

in series connections these problems can occur using end effector coordinates. In describing<br />

the dynamics, the equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> a mechanical <strong>system</strong>, not choosing appropriate coordinates<br />

then leads to combined differential algebraic equations which results in additional<br />

problems in simulation or model <strong>based</strong> <strong>control</strong>. In the most general mechanical <strong>system</strong>s<br />

both problems occur together with possible nonholonomy, in which case it is not clear if<br />

or which coordinates can be chosen to have a global description in explicit form. In the<br />

Stewart Platforms, i.e. the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s, only parallelism has to be taken<br />

care <strong>of</strong>f and its dynamics can be described explicitly as a function <strong>of</strong> platform coordinates.<br />

Of course still choices have to be made in describing the spatial rotation. The use <strong>of</strong> quaternions<br />

or Euler-parameters leads to a computationally simple description also favourable in<br />

<strong>system</strong> analysis although an additional but easy to handle, constraint equation results.<br />

Symbolic multibody equation <strong>of</strong> <strong>motion</strong> solvers are suitable to generate a simulation<br />

model in short time and without to much difficulties. These solvers, however, as they handle<br />

the most general form <strong>of</strong> mechanical <strong>system</strong>s, do not easily recognise special properties, like<br />

the possible explicit form <strong>of</strong> certain <strong>system</strong>s like the Stewart Platforms. Large numbers <strong>of</strong><br />

scalar variables result with <strong>system</strong>s <strong>of</strong> limited complexity like the <strong>motion</strong> <strong>system</strong>. They are


206 6 Review and discussion on the results<br />

not suited to represent automatically appropriate vector and matrix equations or to take into<br />

account equivalence between the different actuator legs. This makes <strong>system</strong> analysis more<br />

difficult.<br />

Using a projection method, the equations <strong>of</strong> <strong>motion</strong> <strong>of</strong> hexapod <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong><br />

<strong>system</strong> can be written down preserving insight. Further, this shows that in this case both<br />

Euler-Newton as Lagrange, whatever is appropriate, methods can be used in stating the<br />

local equations <strong>of</strong> <strong>motion</strong>.<br />

System analysis <strong>of</strong> multibody mechanical <strong>system</strong> shows the following. The strong coupling<br />

in the <strong>system</strong> evaluated along the platform coordinates is mainly due to the coupled<br />

transposed jacobian w.r.t. the centre <strong>of</strong> gravity mapping the actuator forces onto the platform<br />

with specific mass and inertial properties. The jacobian varies with the platform position<br />

and forms the main non-linearity in the mechanical <strong>system</strong>. Choice <strong>of</strong> appropriate kinematical<br />

parameters, mass and inertial properties in the design <strong>of</strong> the construction should result in<br />

reasonable conditioning <strong>of</strong> the mass matrix evaluated at the actuators. Inertial properties <strong>of</strong><br />

actuators are not very relevant in conventional <strong>system</strong>s due to high mass payload. With the<br />

Simona <strong>motion</strong> <strong>system</strong> the one body mass matrix structure is not enough. Varying relative<br />

actuator inertia is, however, still neglectable. The centripetal and coriolis forces will remain<br />

relatively small due to the limited velocity <strong>of</strong> the actuators and reasonably low centre <strong>of</strong><br />

gravity.<br />

Connection <strong>of</strong> hydraulics and mechanics results in the most relevant modes, the platform<br />

pose dependent rigid modes. Direction <strong>of</strong> eigenvectors <strong>of</strong> these modes only depend<br />

on the mass matrix <strong>of</strong> mechanical <strong>system</strong> in which the jacobian plays an important part.<br />

Eigenfrequencies typically vary three to four times in magnitude as the eigenvalues <strong>of</strong> the<br />

mass matrix vary ten to twenty times.<br />

The collocated input/output pair <strong>of</strong> valve position and valve pressure difference <strong>of</strong> a<br />

basic set <strong>of</strong> hydraulic actuator models connected to a general non-linear mechanical <strong>system</strong><br />

is passive. This is not the case with taking resulting actuator output pressure or platform<br />

acceleration as an output. The property is also lost if valve dynamics can not be neglected<br />

as is the case with the range where transmission line dynamics plays a role.<br />

Passive feedback <strong>of</strong> pressure difference will not cause stability problems with any flexible<br />

mode (e.g. position dependent rigid modes, fundament or <strong>simulator</strong> parasitic dynamics,<br />

(transmission lines dynamics)). Direct feedback <strong>of</strong> pressure to flow e.g. leakage provides<br />

for an additional dissipative term like viscous friction as feedback from velocity to force<br />

does in purely mechanical <strong>system</strong>s. Total energy consists <strong>of</strong> both the mechanical kinetic<br />

part as the hydraulic pressure oil stiffness part.<br />

6.4 Experimental modelling and validation<br />

The kinematical model forms a basic part <strong>of</strong> the dynamics. Validation <strong>of</strong> the kinematics<br />

and identification <strong>of</strong> the parameters is the first step in identifying a full model. Redundant<br />

measurements are required to identify the parameters. It is important to distinguish measurements<br />

in calibration, which should be accurate vs. those which do not matter. Combining<br />

two existing methods with limited adaptation, first redundant measurement <strong>of</strong> platform pose


6.4 Experimental modelling and validation 207<br />

and secondly identification <strong>of</strong> the kinematic parameters, results in closely fit (within evaluation<br />

accuracy) kinematics. This improves positional accuracy by an order <strong>of</strong> a magnitude<br />

and can be attained by low cost length measuring devices. The location <strong>of</strong> the devices and<br />

their connection on moving and rigid base is basically not important though it can influence<br />

the sensitivity <strong>of</strong> the method. Important are the absolute lengths between the devices and<br />

the absolute measurement <strong>of</strong> length itself.<br />

In evaluating <strong>simulator</strong> acceleration response, the location <strong>of</strong> the accelerometers is highly<br />

relevant. Gain, direction and location can be identified using the kinematical model <strong>of</strong> the<br />

<strong>system</strong> with reasonable accuracy. Conditioning filters should also be known and taken into<br />

account in evaluating measurements.<br />

In this research, the <strong>motion</strong> <strong>system</strong> has been tested with three different load conditions.<br />

The unloaded dummy platform (2250 kg), the empty shuttle (1680 kg) and finally the<br />

dummy platform with full operational <strong>flight</strong> <strong>simulator</strong> load (4000 kg).<br />

Already with a basic hydraulic/mechanical <strong>system</strong> model, the most relevant open loop<br />

dynamics can quite closely be predicted. This was confirmed by measuring the result <strong>of</strong><br />

sine sweeps with a frequency analyser. With the Simona <strong>motion</strong> <strong>system</strong> this rigid body<br />

dynamics typically showed around five to thirty Hertz. Dynamics, earlier obtained from a<br />

single actuator and resulting from the valve and transmission lines is still valid, confirmed,<br />

in the multivariable <strong>system</strong> but appears for most in the high frequency area (larger than fifty<br />

Hertz) .<br />

With the dummy platform additional resonances could be observed in the mid frequency<br />

area, eight to fifteen Hertz, which resulted from a non rigid foundation at the test site as<br />

confirmed by additional measurements with accelerometers attached to the ground. By<br />

assuming a planar mass/spring/damper <strong>system</strong> in the fundament this additional dynamics<br />

could be modelled, which led to a reasonable fit to the measurements taken at both the<br />

pressure dynamics in the actuator and the acceleration measurements at the ground. With<br />

the loaded dummy platform the dynamics <strong>of</strong> the foundation changes since a relatively higher<br />

dissipation (possibly relatively more work done by friction) can be observed leading to<br />

overdamped characteristics. Design <strong>of</strong> the foundation at a <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> test site<br />

should be done in careful consideration and always be evaluated properly after the <strong>motion</strong><br />

<strong>system</strong> has been set up.<br />

Accelerometer measurements in the newly designed shuttle reveal deformation at the<br />

frequencies where resonances occur in the pressure dynamics. Though stiff (> 40 Hz), the<br />

modes <strong>of</strong> the <strong>system</strong> can be very lightly damped and occur in a frequency area, which was<br />

previously used to attenuate pressure feedback (introducing phase lag losing the passivity<br />

property). Further, one <strong>of</strong> the modes <strong>of</strong> the shuttle (at 65 Hz) appeared to be almost unobservable<br />

in the valve pressure dynamics as the transmission line dynamics block (pair <strong>of</strong><br />

complex zeros) at this point. The number <strong>of</strong> measurements and the ability <strong>of</strong> the data acquisition<br />

<strong>system</strong> should be enlarged to identify such mode shapes and other characteristics<br />

<strong>of</strong> a flexible <strong>system</strong> more closely. Flexible deformations can be expected to occur at much<br />

lower frequencies with a large projection screen attached. Further, a less exciting experiment<br />

should be designed in case <strong>of</strong> a full operational <strong>simulator</strong> to prevent such a <strong>system</strong><br />

from any harm done.<br />

If flexibility <strong>of</strong> the fundament can be modelled precisely, it can be compensated for,


208 6 Review and discussion on the results<br />

but this is doubtful. The flexibility <strong>of</strong> the <strong>simulator</strong> results in a principle compromise since<br />

only six coordinates can be functionally <strong>control</strong>led. Problems with these kinds <strong>of</strong> parasitic<br />

modes should for most be dealt with in design <strong>of</strong> the <strong>system</strong>. Not afterwards in <strong>control</strong>.<br />

6.5 Control strategy and evaluation<br />

With a conventional <strong>control</strong> strategy one uses pressure feedback to introduce extra damping<br />

to the rigid mode resonances. With additional position feedback, a bandwidth close to this<br />

resonance frequency can be obtained. As in this case, decentralised feedback is used, no<br />

explicit compensation for each direction, which typically has a range <strong>of</strong> the rigid mode<br />

frequencies, is taken into account and worst case compensation, usually overdamping the<br />

<strong>system</strong> with respect to the variation in platform pose, will have to be settled for. Feed<br />

forward can at best locally provide for adequate amplification <strong>of</strong> desired high frequency<br />

references. If the <strong>system</strong> is operated and testing is only performed in or close to any neutral<br />

position, this can be sufficient but still can require quite some tuning. Tuning can be reduced<br />

if a model is taken into account in design.<br />

With a model-<strong>based</strong> <strong>control</strong> strategy, one tries to compensate for the directionally different<br />

rigid modes varying as a function <strong>of</strong> the position by taking into account the varying<br />

jacobian and platform mass matrix. Taking into account the integration <strong>of</strong> hydraulics and<br />

mechanics will <strong>of</strong> course form the basic part <strong>of</strong> this strategy. To attain a <strong>control</strong>ler, which is<br />

structured and can be implemented in different modules, it was split up in the following sub<br />

tasks.<br />

- Inner loop decentralised actuator pressure <strong>control</strong><br />

- Partial feedback linearisation<br />

- Coordinate reconstruction<br />

- Reference model <strong>based</strong> feed forward<br />

- Outer loop position stabilisation<br />

As acceleration in the mid-frequency area is important, the hydraulic actuators should<br />

be able to deliver the required force to attain this acceleration. With the so-called cascadedp<br />

<strong>control</strong> one compensates for the oil necessary to move at actual velocity and feeds back<br />

the pressure necessary to attain required force. With this principle, a bandwidth <strong>of</strong> two to<br />

three times the rigid mode frequency can be obtained. Velocity compensation principally<br />

decouples the influence <strong>of</strong> the mechanics on the hydraulics.<br />

The lowest rigid mode frequency can be found at 5 Hz with the fully loaded Simona<br />

<strong>motion</strong> <strong>system</strong> in the neutral position. It will typically change to lower values moving to<br />

other positions.<br />

As the velocity is not measured directly, it has to be reconstructed by an observer. Both<br />

position and pressure measurements and the desired velocity can be used to calculate the<br />

required compensation. Differentiation <strong>of</strong> position is simple but very limited in frequency<br />

due to measurement errors. Pressure can be used with higher frequencies but requires the


6.5 Control strategy and evaluation 209<br />

full multivariable mechanical model which is not exact, adds multiple uncertainties and is<br />

further very sensitive to friction which is small but existent in the <strong>motion</strong> <strong>system</strong>. Even with<br />

exact velocity reconstruction, velocity compensation will not be exact due to uncertainty in<br />

the valve characteristics.<br />

Also platform state reconstruction has to be performed since the model is only explicit<br />

in platform coordinates. An iterative scheme is used to calculate the platform pose from<br />

actuator position measurement. Use in feedback requires convergence at all time for stability<br />

and guaranteed accuracy in a very limited amount <strong>of</strong> iterations if applied in real time.<br />

Sufficient conditions on the convergence (speed) <strong>of</strong> a general Newton-Raphson scheme can<br />

be translated to the general Stewart Platform. The jacobian used in the scheme is shown to<br />

be Lipschitz under mild conditions. With the Simona <strong>motion</strong> <strong>system</strong>, convergence in two<br />

iterations within measurement device accuracy can be guaranteed given the limited actuator<br />

speed and a sufficiently high update frequency (100 Hz), which was easily attainable.<br />

With partly feedback linearising <strong>control</strong> <strong>of</strong> the mechanical <strong>system</strong>, one strives for two<br />

objectives. First, from feed forward <strong>of</strong> desired accelerations, required forces are calculated<br />

taking the varying mass matrix, non-linear velocity terms and gravity into account. Further,<br />

feedback <strong>of</strong> positional and velocity error is performed in a theoretically decoupled set <strong>of</strong><br />

coordinates. Translation <strong>of</strong> desired platform forces to actuator forces requires inversion <strong>of</strong><br />

the jacobian, which, however, can also be used with the coordinate reconstruction.<br />

In the conventional structure <strong>of</strong> <strong>motion</strong> simulation, the series connection <strong>of</strong> <strong>control</strong>ler<br />

and <strong>motion</strong> <strong>system</strong> ideally requires an infinitely fast <strong>control</strong>led <strong>motion</strong> <strong>system</strong>. Since the<br />

aircraft model and wash-out filters are incorporated in the simulation program, reference<br />

model knowledge can be used to generate a predicted acceleration over a certain amount <strong>of</strong><br />

time. 30-50 ms would be sufficient to attain a situation in which the acceleration onset can<br />

be felt exactly at the required instant <strong>of</strong> time with a finitely fast <strong>motion</strong> <strong>system</strong>. This will be<br />

possible since the model <strong>based</strong> <strong>control</strong>led <strong>motion</strong> <strong>system</strong> behaves in a predictable manner.<br />

As positional accuracy is not the main objective <strong>of</strong> the <strong>control</strong>ler, the position feedback<br />

should only be used in preventing the <strong>system</strong> from running out <strong>of</strong> stroke and attaining the<br />

right accelerations in the low frequency area. Especially at the bandwidth <strong>of</strong> the positional<br />

feedback, care should be taken not to distort the appropriately generated accelerations.<br />

A model <strong>based</strong> <strong>control</strong>ler is able to enlarge the bandwidth <strong>of</strong> the accelerations to be simulated<br />

by a factor two w.r.t. the directions <strong>of</strong> the rigid mode with the lowest frequency and<br />

also compared to the conventional <strong>control</strong>led <strong>system</strong>. This was confirmed in this research.<br />

By limiting the bandwidth in the other directions to attain equal response in any direction,<br />

unnecessary excitation <strong>of</strong> unmodelled dynamics, amplification <strong>of</strong> noise and coupling between<br />

rotation and translation is avoided.<br />

The dynamics <strong>of</strong> the foundation at a test site, which is not explicitly taken into account,<br />

limits the performance <strong>of</strong> the <strong>system</strong>. Success <strong>of</strong> compensation is doubtful since the effect<br />

is seen to vary a unpredictable manner w.r.t. payload characteristics.<br />

System feedback <strong>control</strong> enforces reasonable linear response. With this assumption,<br />

the characteristics <strong>of</strong> the <strong>system</strong> can be tested by measuring the frequency response. The<br />

assumption can be verified by performing some additional tests. In case <strong>of</strong> a <strong>motion</strong> <strong>system</strong>,<br />

the linearity can be tested with respect to a set <strong>of</strong> different amplitude acceleration<br />

step responses and a standardised number <strong>of</strong> realistic manoeuvres representative for <strong>motion</strong>


210 6 Review and discussion on the results<br />

through the workspace <strong>of</strong> the <strong>system</strong>. Exact quantification <strong>of</strong> the noise would require more<br />

sophisticated methods and measurement devices. Moreover, the noise seems to vary over<br />

time.<br />

A modern standardised test method is required to evaluate the performance <strong>of</strong> a <strong>simulator</strong><br />

<strong>motion</strong> <strong>system</strong>. In this research some building blocks to set up such a test were provided.<br />

Main aspect in such a test should be the predictability <strong>of</strong> a <strong>system</strong> i.e. guaranteed simulation<br />

quality. The model <strong>based</strong> approach taken in this research helps to attain this.


Chapter 7<br />

Conclusion and recommendations<br />

7.1 Conclusion<br />

Central to this conclusion is the main problem statement to investigate what relevant <strong>system</strong><br />

knowledge should be used in a model <strong>based</strong> <strong>control</strong> strategy for a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong><br />

<strong>system</strong> and to what extent this results in improved <strong>control</strong>led dynamics.<br />

In this thesis, the path <strong>of</strong> choice became physical modelling through balance equations<br />

<strong>of</strong> the hydraulically driven mechanical <strong>system</strong>, identification <strong>of</strong> the model parameters, experimental<br />

evaluation <strong>of</strong> the model structure and explicit use <strong>of</strong> the model in the <strong>control</strong>ler.<br />

An important first observation, considering the mechanical construction <strong>of</strong> Stewart platform<br />

like <strong>motion</strong> <strong>system</strong>, is their limited complexity if compared to general higher degree<br />

<strong>of</strong> freedom manipulators. The dynamics can explicitly be described as a function <strong>of</strong> a fixed<br />

set <strong>of</strong> platform coordinates unlike many combined serial/parallel constructions. This also<br />

limits the complexity <strong>of</strong> a model <strong>based</strong> <strong>control</strong> strategy.<br />

The mapping between the variations along the platform coordinates and actuator coordinates,<br />

the jacobian, is an important parameter in <strong>control</strong> <strong>of</strong> multi degree <strong>of</strong> freedom manipulators<br />

such as the Stewart platform. Most <strong>of</strong> the calculations in the model <strong>based</strong> <strong>control</strong>ler<br />

presented for the <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> are directed towards this parameter.<br />

In parallel <strong>system</strong>s, <strong>control</strong>lability is lost if the jacobian loses rank in some position,<br />

unlike serial robotic <strong>system</strong>s. In this thesis, a method was presented to exclude these kind<br />

<strong>of</strong> singular positions from the reachable envelope <strong>of</strong> a parallel <strong>system</strong> considering limited<br />

stroke actuators. For the Simona Research Simulator (SRS) it was proven that such points,<br />

although not far away, can just not be reached.<br />

If the position <strong>of</strong> a parallel <strong>system</strong> is only measured along the actuator coordinates,<br />

an inverse problem reconstructing the platform coordinates has to be solved in real time<br />

using the model <strong>based</strong> <strong>control</strong> strategy. In this thesis a method was presented to prove<br />

convergence <strong>of</strong> the specific Newton-Rapshon iteration chosen anywhere in the reachable<br />

envelope <strong>of</strong> the platform. It was proven that it does and it does sufficiently fast for the<br />

SRS. The platform coordinate reconstructor could therefore be made a central module in<br />

the model <strong>based</strong> <strong>control</strong>ler, which was implemented.<br />

The platform pose dependent jacobian together with the platform coordinate related<br />

mass matrix <strong>of</strong> the <strong>system</strong> recover the specific masses as seen through the actuators. In<br />

211


212 7 Conclusion and recommendations<br />

most <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s, and also the SRS, large variations <strong>of</strong> the specific<br />

masses, more than an order <strong>of</strong> a magnitude, can be expected. Therefore, only misdirecting<br />

a force 1% easily results in a ten times higher accelerational error <strong>of</strong> 10%. Feedforward<br />

<strong>control</strong> therefore requires highly accurate models.<br />

The kinematical model structure <strong>of</strong> the <strong>system</strong> was shown to be accurate within tens<br />

<strong>of</strong> mm. Calibration <strong>of</strong> the model parameters improved the positional accuracy an order <strong>of</strong><br />

a magnitude. The structural model properties <strong>of</strong> the dynamics <strong>of</strong> the hydraulically driven<br />

mechanical <strong>motion</strong> <strong>system</strong> were also confirmed. However, estimation <strong>of</strong> the <strong>system</strong> mass<br />

properties can probably be improved despite the scheme developed to robustly derive the<br />

mass properties <strong>of</strong> such <strong>system</strong>s in the presence <strong>of</strong> dissipation or some nonlinear effects.<br />

The specific mass directions are almost equal to the rigid body modal directions connecting<br />

the hydraulic actuators. In analysis and identification <strong>of</strong> hydraulically driven <strong>motion</strong><br />

<strong>system</strong>s, knowledge <strong>of</strong> these directions proved to be very helpful to decouple the, at first<br />

sight, highly interactive dynamics. Un<strong>control</strong>led hydraulically driven <strong>system</strong>s are usually<br />

lightly damped and this increases the already large difference along the rigid body modal<br />

directions in the neighbourhood <strong>of</strong> the resonance frequencies. The rigid body resonance<br />

modes form the main dynamics in the most relevant frequency area for <strong>flight</strong> simulation up<br />

till 30Hz.<br />

The conventional <strong>control</strong> structure robustifies the decentralised positional feedback through<br />

dissipating energy by pressure feedback thereby damping the resonance peaks. In this thesis<br />

this strategy proved right by showing that feedback <strong>of</strong> the input valve pressure difference<br />

to valve flow is feedback <strong>of</strong> a passive transfer function, also in the presence <strong>of</strong> parasitic<br />

dynamics e.g. due to transmission line resonances or mechanical flexibilities. However,<br />

no directionality can be accounted for in this structure. It was shown that this leads to<br />

limited bandwidth <strong>system</strong>s, which still have to compromise overdamped directions with<br />

some peaking <strong>of</strong> others.<br />

With the model <strong>based</strong> strategy, the already known cascade dp structure was chosen,<br />

which maintained local decentralised dissipating pressure feedback by inner loops and decoupled<br />

hydraulics from mechanics. This leaves the directionality to be dealt with by the<br />

multivariable feedback <strong>of</strong> the mechanics. Through implementation <strong>of</strong> this <strong>control</strong>ler on<br />

the SRS it was shown that this indeed leads to a <strong>system</strong> with higher bandwidth and more<br />

equalised response in each degree <strong>of</strong> freedom.<br />

The attainable improvement through the model <strong>based</strong> <strong>control</strong>ler chosen is limited in several<br />

ways. Not all dynamics was or could be made part <strong>of</strong> the model used in <strong>control</strong>. In testing<br />

the SRS, <strong>of</strong> course, the dynamics <strong>of</strong> the foundation appeared in the relevant frequency<br />

area due to the non ideal experimental environment. But this should not be a problem in a<br />

specific <strong>simulator</strong> building. However, it can be concluded that the absence <strong>of</strong> such dynamics<br />

should be confirmed, tested, in any place where <strong>motion</strong> <strong>system</strong>s are to be used or evaluated.<br />

The flexible deformations <strong>of</strong> the <strong>system</strong> are a more severe problem. This causes loss <strong>of</strong><br />

functional <strong>control</strong>lability, strived for in simulation. Adding additional objects, e.g. projection<br />

screens, will further limit the attainable performance.<br />

The phase lag <strong>of</strong> the hydraulic valve usually limits the attainable feedback pressure<br />

bandwidth. With fast valves one can choose this bandwidth in between the rigid body<br />

modes and the transmission lines resonances but exactly in this frequency area the flexible


7.2 Recommendations 213<br />

resonances show up if velocity is not precisely compensated for, as is impossible in practice.<br />

The physical modelling in this research enabled a limited complexity model <strong>based</strong> <strong>control</strong><br />

structure for <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s by choosing only the most relevant dynamical<br />

effects to be incorporated. Also the parameter identification procedure could be made<br />

very compact in this way. The model <strong>based</strong> <strong>control</strong>ler proved better in almost all respects if<br />

compared to a conventional <strong>control</strong> structure implemented on the actual <strong>motion</strong> <strong>system</strong>.<br />

Summarising, the full design process proposed was shown to be applicable and leads to<br />

more predictable characteristics and higher performance <strong>of</strong> <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s.<br />

7.2 Recommendations<br />

This thesis <strong>system</strong>atically went through the full design process <strong>of</strong> <strong>motion</strong> <strong>control</strong> design <strong>of</strong><br />

<strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>s but it still leaves many points <strong>of</strong> research open for more<br />

thorough investigation. Some at the high abstract level <strong>of</strong> the over all simulation, some at<br />

the lower level <strong>of</strong> detail though in practise <strong>of</strong>ten the most troublesome.<br />

First <strong>of</strong> all, specifications towards the design <strong>of</strong> <strong>motion</strong> <strong>control</strong>lers in simulation should<br />

become more precise in stating in what way the fundamentally non exact <strong>motion</strong> simulation<br />

characteristics should be compromised for. E.g. how much more accelerational noise can<br />

be settled for in attaining one additional Hertz <strong>of</strong> bandwidth? What class <strong>of</strong> signals will<br />

have to be tracked by the <strong>motion</strong> <strong>system</strong>? Usually there is much information in this area,<br />

structure in the trajectories, but not yet very well specified. Short time spectral properties<br />

<strong>of</strong> the reference signals is relevant in this sense since otherwise the important aspects, high<br />

frequency content, <strong>of</strong> e.g. a landing bump, would not be revealed.<br />

These specifications should form the basis for a standardised test to measure the performance<br />

and predictability i.e. robustness <strong>of</strong> a <strong>motion</strong> <strong>system</strong>. In this research a number<br />

<strong>of</strong> tests were introduced to measure the characteristics <strong>of</strong> a <strong>control</strong>led <strong>motion</strong> <strong>system</strong> in a<br />

more modernised manner as compared to the current <strong>flight</strong> simulation standards. But only<br />

providing more specific guidelines, weighting factors w.r.t. desired properties <strong>of</strong> a <strong>motion</strong><br />

<strong>system</strong>, will enable the design and analysis <strong>of</strong> the right test and can lead the way towards<br />

the most effective <strong>control</strong> strategy.<br />

Information from the <strong>flight</strong> simulation model can also be used more conveniently in the<br />

<strong>control</strong> scheme itself. Limited short response time <strong>of</strong> the required pressure results in lagged<br />

accelerational response <strong>of</strong> the <strong>motion</strong> <strong>system</strong> as compared to what was required in simulation.<br />

As the spectrum <strong>of</strong> the desired accelerations has almost always limited bandwidth,<br />

predictive information theoretically is sufficient to exactly attain the required response. In<br />

this thesis it was shown how this information can be used. Flight simulation models should<br />

be set up in such a way that this predictional information can be communicated to the <strong>motion</strong><br />

<strong>system</strong> <strong>control</strong>ler. It is the most viable way to attain virtual zero time delay response in<br />

simulation.<br />

Dynamics, not taken into account by the model <strong>based</strong> <strong>control</strong>ler, occurred in practise.<br />

The most important part was shown to be caused by the flexible deformations <strong>of</strong> the <strong>system</strong>,<br />

which in future will become even more relevant. Procedures to identify the specific structure<br />

<strong>of</strong> these parasitic resonances will have to be set up and applied. With more specifically


214 7 Conclusion and recommendations<br />

directed identification techniques in general, parameter estimation <strong>of</strong> the <strong>system</strong> model can<br />

probably be improved further.<br />

The <strong>control</strong> strategy should be extended in order to deal effectively with the uncertain<br />

or identified flexible dynamics. One could even consider applying additional actuators,<br />

to dissipate energy stemming from local resonant modes, to solve the current problem <strong>of</strong><br />

functional <strong>control</strong>lability in the presence <strong>of</strong> deformations. Also additional sensors, such<br />

as the accelerometers applied in identification, could be considered in an extended <strong>control</strong><br />

structure. After all, acceleration felt in the cockpit is what really matters in <strong>motion</strong> cues for<br />

<strong>flight</strong> simulation.<br />

Dealing with uncertainty in general in <strong>control</strong> requires robust feedback/feed forward<br />

schemes. <strong>Model</strong>s can never be exact. Unfortunately many modern robust <strong>control</strong> techniques<br />

are not mature enough at the moment to deal efficiently with nonlinear six d.o.f. hydraulically<br />

driven <strong>motion</strong> <strong>system</strong>s. In the side line <strong>of</strong> this research some first steps towards robust<br />

<strong>control</strong> <strong>of</strong> such <strong>system</strong>s were taken but this did not lead to satisfying results yet. Still, attaining<br />

robustly guaranteed properties in simulation is important in <strong>flight</strong> training and makes<br />

more effort in this direction worthwhile.<br />

W.r.t. the specific model <strong>based</strong> <strong>control</strong> structure chosen, the interaction between the inner<br />

and outer loops appeared to be more severe than expected. More exact velocity compensation,<br />

central in the decoupling <strong>of</strong> mechanics and hydraulics, would directly enhance the<br />

proposed <strong>control</strong> structure. This requires more accurate velocity estimation first. E.g. due<br />

to a structural deviation in the position measurement the increased resolution could not yet<br />

be used in more precise velocity reconstruction though it can probably be compensated for.<br />

Next, enhanced <strong>control</strong> over the valve, the oil flow to and from the actuators, would be <strong>of</strong><br />

help. Especially the nonlinear characteristics <strong>of</strong> the valves <strong>of</strong>ten appeared to be troublesome<br />

in practise. E.g. each valve has to be activated at all time by a high frequent dither signal to<br />

prevent stick <strong>of</strong> one <strong>of</strong> the spools. In setting the dither amplitude one had to be very careful<br />

not to hit transmission line dynamics (too high) or otherwise let sticking effects occur (too<br />

low). An adaptive setting can improve this.<br />

Over all this thesis provides on one hand many leads towards more scientific investigations<br />

and on the other hand laid down an improved design procedure, which can be applied<br />

directly by the practioneer. And that’s what’s research is all about.


Appendix A<br />

Frequency domain measurements<br />

215


hdp_i1<br />

hdp_i2<br />

hdp_i3<br />

hdp_i4<br />

hdp_i5<br />

hdp_i6<br />

216 A Frequency domain measurements<br />

hu_1<br />

hu_2 hu_3 hu_4 hu_5 hu_6<br />

Fig. A.1: Bode amplitude plots dummy platform from valve voltage to valve pressure<br />

measured and modelled (dashed). Frequency runs form 5Hz to 50Hz with<br />

ticks at 10Hz, 20Hz and 30Hz. Amplitude from 0:5 to 50 with ticks at 1 and 10.


hdp_i1<br />

hdp_i2<br />

hdp_i3<br />

hdp_i4<br />

hdp_i5<br />

hdp_i6<br />

hu_1<br />

hu_2 hu_3 hu_4 hu_5 hu_6<br />

217<br />

Fig. A.2: Bode phase plots dummy platform from valve voltage to valve pressure measured<br />

and modelled (dashed). Frequency runs form 5Hz to 50Hz with ticks at<br />

10Hz, 20Hz and 30Hz. Phase from ;700 to 100 with ticks from ;135 to<br />

90 every 45 .


hdp_i1<br />

hdp_i2<br />

hdp_i3<br />

hdp_i4<br />

hdp_i5<br />

hdp_i6<br />

218 A Frequency domain measurements<br />

hu_1<br />

hu_2 hu_3 hu_4 hu_5 hu_6<br />

Fig. A.3: Bode amplitude plots dummy platform from valve voltages to valve pressures<br />

along the rigid body modal directions measured and modelled (dashed). Frequency<br />

runs form 5Hz to 50Hz with ticks at 10Hz, 20Hz and 30Hz. Amplitude<br />

from 0:5 to 50 with ticks at 1 and 10.


hdp_i1<br />

hdp_i2<br />

hdp_i3<br />

hdp_i4<br />

hdp_i5<br />

hdp_i6<br />

hu_1<br />

hu_2 hu_3 hu_4 hu_5 hu_6<br />

219<br />

Fig. A.4: Bode phase plots dummy platform from valve voltages to valve pressures<br />

along the rigid body modal directions measured and modelled (dashed). Frequency<br />

runs form 5Hz to 50Hz with ticks at 10Hz, 20Hz and 30Hz. Phase<br />

from ;700 to 100 with ticks from ;135 to 90 every 45 .


hdp_i1<br />

hdp_i2<br />

hdp_i3<br />

hdp_i4<br />

hdp_i5<br />

hdp_i6<br />

220 A Frequency domain measurements<br />

hu_1<br />

hu_2 hu_3 hu_4 hu_5 hu_6<br />

Fig. A.5: Bode amplitude plots shuttle from valve voltage to valve pressure measured<br />

and modelled (dashed). Frequency runs form 5Hz to 500Hz with ticks at 10Hz<br />

and 100Hz. Amplitude from 0:03 to 30 with ticks at :1, 1 and 10.


hdp_i1<br />

hdp_i2<br />

hdp_i3<br />

hdp_i4<br />

hdp_i5<br />

hdp_i6<br />

hu_1<br />

hu_2 hu_3 hu_4 hu_5 hu_6<br />

221<br />

Fig. A.6: Bode amplitude plots shuttle from valve voltage to valve pressure measured<br />

and modelled (dashed). Frequency runs form 5Hz to 500Hz with ticks at 10Hz<br />

and 100Hz. Phase from ;700 to 100 with ticks from ;180 to 90 every 90 .


hdp_i1<br />

hdp_i2<br />

hdp_i3<br />

hdp_i4<br />

hdp_i5<br />

hdp_i6<br />

222 A Frequency domain measurements<br />

hu_1<br />

hu_2 hu_3 hu_4 hu_5 hu_6<br />

Fig. A.7: Bode phase plots shuttle from valve voltages to valve pressures along the<br />

rigid body modal directions measured and modelled (dashed). Frequency<br />

runs form 5Hz to 500Hz with ticks at 10Hz and 100Hz. Amplitude from 0:03<br />

to 30 with ticks at :1, 1 and 10.


hdp_i1<br />

hdp_i2<br />

hdp_i3<br />

hdp_i4<br />

hdp_i5<br />

hdp_i6<br />

hu_1<br />

hu_2 hu_3 hu_4 hu_5 hu_6<br />

223<br />

Fig. A.8: Bode phase plots shuttle from valve voltages to valve pressures along the<br />

rigid body modal directions measured and modelled (dashed). Frequency<br />

runs form 5Hz to 500Hz with ticks at 10Hz and 100Hz. Phase from ;700 to<br />

100 with ticks from ;180 to 90 every 90 .


ax<br />

ay<br />

az<br />

aex<br />

aey<br />

aez<br />

224 A Frequency domain measurements<br />

ax_d<br />

ay_d az_d aex_d aey_d aez_d<br />

Fig. A.9: Bode amplitude plots closed loop around heavy weight dummy platform<br />

from desired to actual ccelerations model <strong>based</strong> <strong>control</strong>ler (-) and conventional<br />

<strong>control</strong>ler (dashed). Frequency runs form 0:5Hz to 50Hz with ticks at<br />

1� 2� 3� 5� 10� 20� 30Hz. Amplitude from 0:025 (;32dB)to2:5 (8dB) with ticks<br />

at ;30� ;20� ;6� ;3� 0� 3� 6dB.


ax<br />

ay<br />

az<br />

aex<br />

aey<br />

aez<br />

ax_d<br />

ay_d az_d aex_d aey_d aez_d<br />

Fig. A.10: Bode phase plots closed loop from desired to actual accelerations model<br />

<strong>based</strong> <strong>control</strong>ler (-) and conventional <strong>control</strong>ler (dashed). Frequency runs<br />

form 0:5Hz to 50Hz with ticks at 1� 2� 3� 5� 10� 20� 30Hz. Phase from ;180<br />

to 180 with ticks from ;180 to 180 every 45 .<br />

225


e_ax<br />

e_ay<br />

e_az<br />

e_aex<br />

e_aey<br />

e_aez<br />

226 A Frequency domain measurements<br />

ax_d<br />

ay_d az_d aex_d aey_d aez_d<br />

Fig. A.11: Bode amplitude plots closed loop from desired to error accelerations model<br />

<strong>based</strong> <strong>control</strong>ler (-) and conventional <strong>control</strong>ler (dashed). Frequency runs<br />

form 0:5Hz to 50Hz with ticks at 1� 2� 3� 5� 10� 20� 30Hz. Amplitude from<br />

;32dB to 14dB with ticks at ;20� ;6� ;3� 0� 3� 6dB.


Appendix B<br />

Derivation <strong>of</strong> actuator inertial<br />

properties<br />

In this appendix essentially the derivation <strong>of</strong> (2.119) and (2.120) is performed. Further, it<br />

is shown how to choose a factor N in the quadratic velocity term to make _ M ; 2N skew<br />

symmetric. This last property is important in passivity <strong>based</strong> <strong>control</strong>.<br />

First the derivation <strong>of</strong> the derivative mass matrix equation (2.119).<br />

and<br />

With<br />

d<br />

dt (Mia�ib )=; (ia + ib)<br />

j l j3 (2v T a lnPln + Plnval T n + lnv T a Pln ) (B.1)<br />

d d p l<br />

j l j= lT l =<br />

dt dt<br />

T va<br />

j l j = lT n va<br />

_<br />

ln = d l<br />

dt j l j =<br />

l<br />

_ j l j;l d dt j l j<br />

j l j2 the required steps can be taken<br />

d<br />

(Mia�ib )<br />

dt<br />

=<br />

d<br />

dt ( (ia + ib)<br />

j l j2 Pln )=(ia + ib)<br />

(B.2)<br />

= (I ; lnlT n )<br />

va =<br />

j l j<br />

1<br />

j l j Plnva: (B.3)<br />

d<br />

dt<br />

1<br />

j l j2 I + d<br />

dt<br />

lnl T n<br />

j l j 2<br />

= (ia + ib) ;2lT n va j l j<br />

j l j4 1<br />

(<br />

jlj<br />

I ; PlnvalT 1<br />

n +<br />

jljlnvT a Pln ) j l j2 ;2lT n va j l j lnlT n<br />

j l j4 = ; (ia + ib)<br />

j l j3 (2v T a lnPln + Plnval T n + lnv T a Pln ): (B.4)<br />

N is defined as a factor in the factorisation <strong>of</strong> the nonlinear velocity term C(x� _x). But<br />

this factorisation is not uniquely defined by requiring that C(x� _x) =N (x� _x) _x.<br />

227


228 B Derivation <strong>of</strong> actuator inertial properties<br />

Taking<br />

f a ia�ib = Mia�ib _va ; 2(ia + ib)<br />

j l j3 (l T n va)Plnva: (B.5)<br />

The nonlinear velocity term, ; 2(ia+ib)<br />

jlj3 (lT n va)Plnva can be factored into N (x� va)va in an<br />

infinite number <strong>of</strong> ways. Two possible factors N are in this case:<br />

and<br />

N1 = ; 2(ia + ib)<br />

j l j3 (l T n va)Pln<br />

(B.6)<br />

N2 = ; 2(ia + ib)<br />

j l j 3 Plnval T n (B.7)<br />

And all factorisations N1 +(1; )N2, with 2 IR are valid. Only one factorisation<br />

results in a skew _ M ; 2N.<br />

Take<br />

Then<br />

N = 1<br />

2 (N1 + N2): (B.8)<br />

_M ; 2N = 2(ia + ib)<br />

j l j3 (Plnval T n ; lnv T a Pln ): (B.9)<br />

Check through _ M ; 2N = ;( _ M ; 2N ) T that the skew symmetry is there (P T<br />

ln<br />

= Pln).<br />

Now the derivation <strong>of</strong> the partial derivative <strong>of</strong> the kinetic energy actuator inertia given<br />

by<br />

@Kia�ib<br />

@pa<br />

= ; (ia + ib)<br />

j l j3 Define a partial derivative <strong>of</strong> a vector, x to a vector, y n 1 as:<br />

@x<br />

@y =<br />

h @x<br />

@y1<br />

@x<br />

@y2<br />

l T n (vT a Plnva) +v T a Pln (lT n va) : (B.10)<br />

:::<br />

@x<br />

@yi<br />

:::<br />

@x<br />

@yn<br />

Kinetic energy <strong>of</strong> the actuator intertia was derived as the scalar term:<br />

with<br />

Kia�ib<br />

i : (B.11)<br />

1<br />

=<br />

2 vT (ia + ib)<br />

a<br />

j l j2 Plnva (B.12)<br />

= 1<br />

2 vT a<br />

(ia + ib)<br />

j l j2 va ; 1<br />

2 (ia + ib)<br />

v T a l<br />

j l j 2<br />

v T a l<br />

j l j 2<br />

� (B.13)<br />

Pln =(I ; lnl T n )� ln = l= j l j : (B.14)


The upper actuator gimbal point, pa, depends on the variables c and in the same way as l:<br />

Therefore<br />

Now<br />

229<br />

pa =c + T ( )a� (B.15)<br />

l =c + T ( )a ; b: (B.16)<br />

@l<br />

= I: (B.17)<br />

@pa<br />

= 1<br />

2 (ia + ib)v T a va<br />

h<br />

1<br />

@<br />

jlj2 i<br />

; (ia + ib)<br />

@pa<br />

h<br />

vT a l<br />

jlj2 i<br />

@Kia�ib<br />

@pa<br />

vT a l<br />

j l j2 @<br />

@pa<br />

: (B.18)<br />

So two partial derivatives to scalars have be examined further. The first<br />

and the second<br />

@<br />

h v T a l<br />

jlj 2<br />

@pa<br />

i<br />

@<br />

h 1<br />

jlj 2<br />

@pa<br />

= vT a j l j 2 ;2v T a ll T<br />

j l j 4<br />

i<br />

= @ 1<br />

l T l<br />

@pa<br />

= vT a<br />

j l j2 ; 2vT a lnlT n<br />

j l j2 = ;2lT<br />

j l j4 = ;2lT n<br />

� (B.19)<br />

j l j3 Filling in these equations in (B.18) results in, using (x T y)=(y T x),<br />

@Kia�ib<br />

@pa<br />

=vT 1<br />

a Pln<br />

j l j2 ; vT a lnlT n<br />

: (B.20)<br />

j l j2 = ; 1<br />

2 (ia + ib)v T 2l<br />

a va<br />

T n<br />

j l j3 ; (ia + ib) v T 1<br />

a Pln<br />

j l j2 ; vT a lnlT n<br />

j l j2 = ; (ia + ib)<br />

j l j 3<br />

= ; (ia + ib)<br />

j l j 3<br />

And this was what was to be determined.<br />

v T a Ival T n ; v T a lnl T n val T n +vT a Pln (vT a ln)<br />

v T a ln<br />

j l j<br />

l T n (vT a Plnva) +v T a Pln (lT n va) : (B.21)


Bibliography<br />

[1] AGARD Advisory Report No. 144. Dynamic characteristics <strong>of</strong> <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong>-bases.<br />

Technical report, NATO, 1979.<br />

[2] AGARD Advisory Report No. 159. Fidelity <strong>of</strong> simulation for pilot training. Technical report,<br />

NATO, 1980.<br />

[3] Federal Aviation Administration. Faa advisory circular 120-40c. Technical report, U.S. Dept.<br />

<strong>of</strong> Transport, 1996.<br />

[4] P. Adorian, W.N. Staynes, and M. Bolton. The evolution <strong>of</strong> the <strong>flight</strong> <strong>simulator</strong>. In Proc. 50<br />

Years <strong>of</strong> Flight Simulation, pages 1–24. The Royal Aeronautical Society, 1979.<br />

[5] Advanced Robotics, 88(6), 1994. Special issue on parallel mechanisms.<br />

[6] S.K. Advani. The development <strong>of</strong> simona: A <strong>simulator</strong> facility for advanced research into<br />

simulation techniques, <strong>motion</strong> <strong>system</strong> <strong>control</strong> and navigation <strong>system</strong>s technologies. In Proc.<br />

AIAA conference on Flight Simulation, pages 156–166, 1993.<br />

[7] S.K. Advani. The kinematic design <strong>of</strong> <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong>-bases. PhD thesis, Delft University<br />

<strong>of</strong> Technology, 1998.<br />

[8] S.K. Advani, M. van Tooren, and S. de Winter. The design <strong>of</strong> a high-performance all-composite<br />

<strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> platform. In Proceedings AIAA Flight Simulation Technologies Conference,<br />

1995.<br />

[9] R.J. Anderson. Passive computed torque algorithms for robots. In Proc. 28th IEEE Conf. on<br />

Decision and Control, pages 1638–1644, 1989.<br />

[10] S.B. Anderson. Historical review <strong>of</strong> piloted simulation at nasa ames. In AGARD Conference<br />

on Flight Simulation - Where are the challenges, pages 1.1–1.13, 1995.<br />

[11] S. Arimoto. Control theory <strong>of</strong> non-linear mechanical <strong>system</strong>s, a passivity-<strong>based</strong> circuittheoretic<br />

approach. Clarendon-Press-Oxford, 1996.<br />

[12] M. Baarspul. Lecture notes <strong>of</strong> <strong>flight</strong> simulation techniques. Technical report, Delft University<br />

<strong>of</strong> Technology, 1989.<br />

[13] M. Baret. Six degrees <strong>of</strong> freedom large <strong>motion</strong> <strong>system</strong> for <strong>flight</strong> <strong>simulator</strong>s. In Proc. AGARD<br />

Conf. on Flight Simulation, pages 22.1–22.8, 1978.<br />

[14] H. Berghuis. <strong>Model</strong>-<strong>based</strong> robot <strong>control</strong>: from theory to practice. PhD thesis, Twente University,<br />

1993.<br />

231


232 Bibliography<br />

[15] R. Bernhardt and S.L. Albright. Robot Calibration. Chapman-Hall, 1993.<br />

[16] S. Bettendorf. Static calibration <strong>of</strong> the simona <strong>motion</strong> <strong>system</strong>. Master’s thesis, Delft University<br />

<strong>of</strong> Technology, 1997.<br />

[17] S.P. Boyd and C.H. Barrat. Linear <strong>control</strong>ler design; limits <strong>of</strong> performance. Prentice-Hall,<br />

1991.<br />

[18] S.P. Boyd, L. El Ghaoui, E. Feron, and V. Balekrishnan. Linear Matrix Inequalities in System<br />

and Control Theory. SIAM, 1994.<br />

[19] K.E. Brenan, S.L. Campbell, and L.R. Petzold. Numerical Solution <strong>of</strong> Initial-Value Problems<br />

in Differential-Algebraic Equations. Elsevier Science Publishers B.V., 1989.<br />

[20] J.A. Carmein and A.J. Clark. New methods for increased fidelity and safety <strong>of</strong> simulation<br />

<strong>motion</strong> platforms. In Proc. AIAA conference on <strong>Model</strong>ling and Simulation, pages 295–304,<br />

1998.<br />

[21] O. Chételat. Algorithme numerique pour les changements de coordonnées des méchanismes<br />

articulés. PhD thesis, École Polytechnique Fédérale De Lausanne, 1997.<br />

[22] W.W. Chung, D.J. Robinson, J. Wong, and D. Tran. Investigation <strong>of</strong> roll-lateral coordinated<br />

<strong>motion</strong> requirements with a conventional hexapod <strong>motion</strong> platform. In Proc. AIAA conference<br />

on <strong>Model</strong>ling and Simulation, pages 91–101, 1998.<br />

[23] C. C. Clark. Human <strong>control</strong> performance and tolerance under severe complex wave form<br />

vibration with a preliminary historical review <strong>of</strong> <strong>flight</strong> simulation. In National Meeting on<br />

Manned Space Flight, pages pp 176–194, 1962.<br />

[24] C. Conijn. Developments in high quality <strong>motion</strong> <strong>system</strong>s for <strong>flight</strong> and driving <strong>simulator</strong>s used<br />

in r&d and training. In CEAS Symposium on Simulation Technologies - Making it real, pages<br />

MSy03/1–12, 1995.<br />

[25] E. Cook. Flight simulation boundaries - barriers to progress or guidlines to succes. In Proc.<br />

Flight Simulation - Expanding the boundaries, pages 4.1–4.10. The Royal Aeronautical Society,<br />

1997.<br />

[26] D.E. Crochiere and L.R. Rabiner. Multirate Digital Signal Processing. Prentice Hall, 1983.<br />

[27] B. Dasgupta and T.S. Mruthyunjaya. The stewart platform manipulator: a review. Mechanism<br />

and Machine Theory, 35(1):15–40, 2000.<br />

[28] R.A. de Callafon, D. de Roover, and P.M.J. van den H<strong>of</strong>. Multivariable least squares frequency<br />

domain identification using polynomial matrix fraction descriptions. In Proc. 35th IEEE Conf.<br />

on Decision and Control, pages 2030–2035, 1996.<br />

[29] L. de Nijs. Evaluation <strong>of</strong> predictors for implementation in a feed-forward <strong>control</strong>ler <strong>of</strong> a <strong>motion</strong><br />

platform. Master’s thesis, Delft University <strong>of</strong> Technology, 1997.<br />

[30] D. de Roover. Motion Control <strong>of</strong> a Wafer Stage; a design approach for speeding up IC production.<br />

PhD thesis, Delft University <strong>of</strong> Technology, 1997.<br />

[31] C. Canudas de Wit, B. Siciliano, and Georges Bastin. Theory <strong>of</strong> Robot Control. Springer,<br />

1996.


Bibliography 233<br />

[32] V. der Kevorkian. Beoordeling van een bewegingsregeling voor de vlucht<strong>simulator</strong>. Master’s<br />

thesis, University <strong>of</strong> Technology Delft, 1996. In Dutch.<br />

[33] C.A. Desoer and Y.T. Wang. Linear time-invariant robust servomechanism problem: a selfcontained<br />

exposition. Control and Dynamic Systems, pages 81–129, 1980.<br />

[34] P. Dietmaier. The stewart-gough platform <strong>of</strong> general geometry can have 40 real postures. In<br />

J. Lenarcic and M.L. Husty, editors, Advances in Robot Kinematics: Analysis and Control,<br />

pages 7–16. 1998.<br />

[35] J.E. Dieudonne, R.V. Parrish, and R.E. Bardusch. An actuator extension transformation for a<br />

<strong>motion</strong> <strong>simulator</strong> and an inverse transformation applying newton-raphson’s method. Technical<br />

Note NASA TN D-7067, NASA Langley Research Center, 1972.<br />

[36] W.Q.D. Do and D.C.H. Yang. Inverse dynamics analysis and simulation <strong>of</strong> a platform type <strong>of</strong><br />

robot. Journal <strong>of</strong> Robotic Systems, 5(3):209–227, 1988.<br />

[37] dSpace GmbH. RTI-MP Real-Time Multi-Processor interface to Simulink, document version<br />

1.0. dSpace GmbH, Paderborn, Germany, 1996.<br />

[38] S. Dutré, H. Bruyninckx, and J. de Schutter. The analytical jacobian and its derivative for a<br />

parallel manipulator. In Proc. IEEE Int. Conf. on Robotics and Automation, pages 2961–2966,<br />

1997.<br />

[39] D.J. Ewins. Modal testing. Wiley, 1984.<br />

[40] B.A. Francis and W.M. Wonham. The internal model principle <strong>of</strong> <strong>control</strong> theory. Automatica,<br />

12:457–465, 1976.<br />

[41] M.D. Franken. Parasitaire trillingen van de simona vlucht<strong>simulator</strong>. Master’s thesis, University<br />

<strong>of</strong> Technology Delft, 1998. In Dutch.<br />

[42] Z. Geng and L.S. Haynes. Six-degree-<strong>of</strong>-freedom active vibration isolation using a stewart<br />

platform mechanism. Journ. <strong>of</strong> Robotic Systems, 10(5):725–744, 1993.<br />

[43] Z. Geng and L.S. Haynes. A 3-2-1 kinematic configuration <strong>of</strong> a stewart platform and its<br />

application to six degrees <strong>of</strong> freedom pose measurements. Robotics and Computer Integrated<br />

Manufacturing, 11(1):23–34, 1994.<br />

[44] F. Ghorbel, O. Chetelat, and R. Longchamp. On the modeling and <strong>control</strong> <strong>of</strong> closed chain<br />

mechanisms. In Proc. IEEE Systems, Man and Cybernetics IMACS Conf. CESA’98, pages<br />

714–720, 1998.<br />

[45] B.B. Goeree and B. Shucker. Geometric attitude <strong>control</strong> <strong>of</strong> a small satellite for ground tracking<br />

maneuvers. In 13 th AIAA Conference on Small Satellites, pages 1–12, 1999.<br />

[46] P.R. Grant. Motion characteristics <strong>of</strong> the utias <strong>flight</strong> research simulation <strong>motion</strong> base. Technical<br />

report, UTIAS Technical Note 261, 1986.<br />

[47] J.R. Hall. Motion versus visual cues in piloted <strong>flight</strong> simulation. In Proc. AGARD Conf. on<br />

Flight Simulation, pages 17.1–17.13, 1978.<br />

[48] J. Heintze. Design and Control <strong>of</strong> a Hydraulically Actuated Industrial Brick Laying Robot.<br />

PhD thesis, Delft University <strong>of</strong> Technology, 1997.


234 Bibliography<br />

[49] J. Heintze and A.J.J. van der Weiden. Inner-loop design and analysis for hydraulic actuators,<br />

with application to impedance <strong>control</strong>. Control Engineering Practice, pages 1323–1330, 1995.<br />

[50] J. Hesselbach, A. Kusiek, and M. Frindt. Parallelkinematiken in der robotertechnik; einsatz<br />

von neuen kinematischen strukturen. ZWF Robotertechniek, 93(6):259–262, 1998.<br />

[51] R. H<strong>of</strong>fman. Dynamics and <strong>control</strong> <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>. In Proc. Canadian<br />

Conference <strong>of</strong> Automatic Control, 1979.<br />

[52] J.M. Hollerbach and C.W. Wampler. The calibration index and taxonomy for robot kinematic<br />

calibration methods. The Int. Journ. <strong>of</strong> Robotics Research, pages 573–591, 1996.<br />

[53] R.J.A.W. Hosman. Pilot’s perception and <strong>control</strong> <strong>of</strong> aircraft <strong>motion</strong>s. PhD thesis, Delft University<br />

<strong>of</strong> Technology, 1996.<br />

[54] R.J.A.W. Hosman and J.C. van der Vaart. Thresholds <strong>of</strong> <strong>motion</strong> perception and parameters <strong>of</strong><br />

vestibular models obtained from tests in a <strong>motion</strong> <strong>simulator</strong>. Effects <strong>of</strong> vestibular and visual<br />

<strong>motion</strong> perception on task performance-Memorandum M-372. Delft University <strong>of</strong> Technology-<br />

Department <strong>of</strong> Aerospace Engineering, 1980.<br />

[55] H. Huisman. Digitaal interface met verhoogde resolutie voor temposonics opnemer n476.<br />

Technical report, Delft University <strong>of</strong> Technology, 1995.<br />

[56] R.L. Huston. Multibody dynamics formulations via kane’s equations. In John L. Junkins,<br />

editor, Mechanics and Control <strong>of</strong> Large Flexible Structures, number 129 in Progress in Astronautics<br />

and Aeronautics, chapter 4, pages 71–86. AIAA, 1990.<br />

[57] M. Husty. An algorithm for solving the direct kinematics <strong>of</strong> general stewart-gough platforms.<br />

Mech. Mach. Theory, 31(4):365–380, 1996.<br />

[58] M. Idan, M.A. Nahon, and D. Sahar. A comparison <strong>of</strong> classical and robust <strong>flight</strong> <strong>simulator</strong><br />

<strong>motion</strong> <strong>control</strong>. In Proc. AIAA conference on <strong>Model</strong>ling and Simulation, pages 98–4366, 1998.<br />

[59] C. Innocenti. Algorithms for kinematic calibration <strong>of</strong> fully-parallel manipulators. In J.P. Merlet<br />

and B. Ravani, editors, Computational Kinematics, pages 241–249. Kluwer, 1995.<br />

[60] C. Innocenti. Polynomial solution <strong>of</strong> the spatial burmester problem. ASME Journal <strong>of</strong> mechanical<br />

design, 117(1):64–68, 1995.<br />

[61] Zhiming Ji. Dynamics decomposition for stewart platforms. Journal <strong>of</strong> Mechanical Design,<br />

116:67–69, December 1994.<br />

[62] B. Jonker. A finite element dynamic analysis <strong>of</strong> flexible manipulators. Int. Journ. <strong>of</strong> Robotics<br />

Res., 9(4):59–74, 1990.<br />

[63] J.L. Junkins and Y. Kim. Introduction to Dynamics and Control <strong>of</strong> Flexible Structures. AIAA<br />

Education Series, 1993.<br />

[64] Thomas R. Kane and David A. Levinson. Dynamics: Theory and Applications. McGraw-Hill<br />

Book Company, 1985.<br />

[65] Thomas R. Kane and David A. Levinson. Dynamics Online, Theory and Implementation with<br />

Autolev. OnLine Dynamics, Inc., 1996.


Bibliography 235<br />

[66] A. Kelkar and Suresh Joshi. Control <strong>of</strong> Nonlinear Multibody Flexible Space Structures.<br />

Springer, 1996.<br />

[67] P.T. Kemmerling. Dynamic characteristics <strong>of</strong> <strong>flight</strong> simulation <strong>motion</strong> <strong>system</strong>s. In Proc.<br />

AGARD Conf. on Flight Simulation, pages 20.1–20.19, 1978.<br />

[68] D. Kim and W. Chung. Analytic singularity equation and analysis <strong>of</strong> six-d<strong>of</strong> parallel manipulators<br />

using local structurization method. IEEE Trans. on Robotics and Automation, 15(4):612–<br />

622, 1999.<br />

[69] D.L. Kleinman, S. Baron, and W.H. Levison. A <strong>control</strong> theoretic approach to manned-vehicle<br />

<strong>system</strong>s analysis. IEEE Trans. on automatic <strong>control</strong>, 16(6):231–249, 1971.<br />

[70] S.H. Koekebakker. Convergence <strong>of</strong> coordinate reconstruction in real-time model <strong>based</strong> <strong>control</strong>.<br />

In Proc. IFAC Workshop on Motion Control, pages 117–122, 1998.<br />

[71] S.H. Koekebakker. Feedback linearisation <strong>of</strong> a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>. Journal A,<br />

Benelux Quarterly Journal on Automatic Control, 39(2):28–34, 1998.<br />

[72] S.H. Koekebakker, Ph. Piatkiewitz, S.K. Advani, and L.J.J. de Nijs. Predictive reference model<br />

<strong>based</strong> <strong>control</strong> <strong>of</strong> the simona <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>. In Proc. AIAA conference on<br />

<strong>Model</strong>ling and Simulation, pages 262–271, 1998.<br />

[73] S.H. Koekebakker, A.J. Scheffer, and S.K. Advani. The dynamic calibration <strong>of</strong> a high performance<br />

<strong>motion</strong> <strong>system</strong>. In Proc. AIAA conference on <strong>Model</strong>ling and Simulation, pages 98–<br />

4365, 1998.<br />

[74] S.H. Koekebakker, P.C. Teerhuis, and A.J.J. van der Weiden. Alternative parametrization in<br />

modelling and analysis <strong>of</strong> a stewart platform. Sel. Topics in Identification, <strong>Model</strong>ling and<br />

Control, 9:59–68, 1996.<br />

[75] S.H. Koekebakker, P.C. Teerhuis, and A.J.J. van der Weiden. <strong>Model</strong>ling and identification <strong>of</strong><br />

a <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>. In Proc. IASTED Conf. on <strong>Model</strong>ling, Identification and<br />

Control, pages 171–174, 1998.<br />

[76] S.H. Koekebakker, P.C. Teerhuis, and A.J.J. van der Weiden. Multiple level <strong>control</strong> <strong>of</strong> a hydraulically<br />

driven <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong> <strong>system</strong>. In Proc. IEEE Systems, Man and Cybernetics<br />

IMACS Conf. CESA’98, pages 886–891, 1998.<br />

[77] K. Kreutz. On manipulator <strong>control</strong> by exact linearization. IEEE Trans. on Automatic Control,<br />

34(7):763–767, 1989.<br />

[78] M. Krstic, I. Kanellakopoulos, and P. Kokotovic. Nonlinear and Adaptive Control Design.<br />

Wiley, 1995.<br />

[79] A. Kurdila, J.G. Papstravides, and M.P. Kamat. Role <strong>of</strong> maggi’s equations in computational<br />

methods for constrained multibody <strong>system</strong>s. Journal <strong>of</strong> Guidance, Control and Dynamics,<br />

13(1):113–120, 1990.<br />

[80] J.D. Lee and Z. Geng. A dynamic model <strong>of</strong> a flexible stewart platform. Computers and<br />

Structures, 48(3):367–374, 1993.<br />

[81] F. Lewis, C. Abdallah, and D. Dawson. Control <strong>of</strong> Robot Manipulators. Macmillan, 1993.


236 Bibliography<br />

[82] D. Li and S.E. Salcudean. <strong>Model</strong>ling, simulation, and <strong>control</strong> <strong>of</strong> a hydraulic stewart platform.<br />

In Proc. IEEE Conf. on Robotics and Automation, pages 3360–3366, 1997.<br />

[83] K. Liu, M. Fitzgerald, D.M. Dawson, and F.L. Lewis. <strong>Model</strong>ing and <strong>control</strong> <strong>of</strong> a stewart<br />

platform manipulator. In N. Sadegh and Y.-H. Chen, editors, Control <strong>of</strong> Systems with Inexact<br />

Dynamic <strong>Model</strong>s, pages 83–89. ASME, 1991.<br />

[84] L. Ljung. System Identification; Theory for the User. Prentice-Hall, 1987.<br />

[85] A. De Luca. Dynamic <strong>control</strong> <strong>of</strong> robots with joint elasticity. In IEEE ROBOT’88, pages<br />

152–158, 1988.<br />

[86] C-M Luh, F.A. Adkins, E.J. Haug, and C.C. Qiu. Working capability analysis <strong>of</strong> stewart<br />

platforms. Trans. ASME: Journal <strong>of</strong> Mechanical Design, pages 220–227, 1996.<br />

[87] O. Ma and J. Angeles. Architecture singularities <strong>of</strong> platform manipulators. In Proc. IEEE<br />

International Conference on Robotics and Automation, pages 1542–1547, 1991.<br />

[88] E. Mach. Grundlinien der lehre von den Bewegungsempfinden. Engelmann, Leipzig, 1875.<br />

[89] J.M. Maciejowski. Multivariable feedback design. Addison-Wesley, 1989.<br />

[90] N.M.M. Maia and J.M.M. Silva. Theoretical and Experimental Modal Analysis. Wiley, 1997.<br />

[91] HP Manual. Dynamic Signal Analyzer <strong>Model</strong> 3562A User Manual. Hewlett-Packard Corp.,<br />

1985.<br />

[92] E.A. Martin. Motion and force simulation <strong>system</strong>s i, whole-body <strong>motion</strong> <strong>simulator</strong>s. Flight<br />

Simulation Update, 1995.<br />

[93] J.E. McInroy. Dynamic modeling and decoupling force <strong>control</strong> <strong>of</strong> a precision hexapod. In<br />

Proc. IFAC Workshop on Motion Control, pages 251–256, 1998.<br />

[94] D.T. McRuer and H.R. Jex. A review <strong>of</strong> quasi-linear pilot models. IEEE Trans. on human<br />

factors in electronics, 8(3):231–249, 1967.<br />

[95] J.-P. Merlet. Parallel manipulators: state <strong>of</strong> the art and perspectives. Advanced Robotics,<br />

8(6):589–596, 1994.<br />

[96] J.P. Merlet. Force-feedback <strong>control</strong> <strong>of</strong> parallel manipulators. In Proc. IEEE Int. Conf. on<br />

Robotics and Automation, pages 1484–1489, 1988.<br />

[97] J.P. Merlet. Singular configurations <strong>of</strong> parallel manipulators and grassmann geometry. International<br />

Journal <strong>of</strong> Robotics Research Vol. 8. No. 5, pages 45–56, 1989.<br />

[98] J.P. Merlet. Determination <strong>of</strong> the presence <strong>of</strong> singularities in a workspace volume <strong>of</strong> a parallel<br />

manipulator. In Computational methods in mechanisms; Workshop <strong>of</strong> the NATO Advanced<br />

study institute, pages 289–295, 1997.<br />

[99] H.E. Merrit. Hydraulic Control Systems. Wiley, 1967.<br />

[100] R.G. Moore, C.N. Pope, and E. Foxlin. Toward minimal latency simulation <strong>system</strong>s. In Proc.<br />

AIAA conference on <strong>Model</strong>ling and Simulation, pages 129–137, 1998.


Bibliography 237<br />

[101] B.W. Mooring, Z.S. Roth, and M. Driels. Fundamentals <strong>of</strong> manipulator calibration. Wiley,<br />

1991.<br />

[102] R.M. Murray, Z. Li, and S.S. Sastry. A Mathematical Introduction to Robotic Manipulation.<br />

CRC, 1994.<br />

[103] C. C. Nguyen, S.S. Antrazi, J.Y. Park, and Z.L. Zhou. Trajectory planning and <strong>control</strong> <strong>of</strong><br />

a stewart platform-<strong>based</strong> end-effector with passive compliance for part-assembly. Journ. <strong>of</strong><br />

intelligent and robotic <strong>system</strong>s, (6):263–281, 1992.<br />

[104] Charles C. Nguyen and Farhad J. Pooran. Dynamic analysis <strong>of</strong> a 6 d<strong>of</strong> ckcm robot end-effector<br />

for dual-arm telerobot <strong>system</strong>s. Robotics and Autonomous Systems, 5(3):377–394, 1989.<br />

[105] H. Nijmeijer and A.J. van der Schaft. Nonlinear dynamical <strong>control</strong> <strong>system</strong>s. Springer Verlag<br />

Berlin, 1990.<br />

[106] P.E. Nikravesh. Computer aided analysis <strong>of</strong> mechanical <strong>system</strong>s. Prentice-Hall, 1988.<br />

[107] P.E. Nikravesh, R.A. Wehage, and O.K. Kwon. Euler parameters in computational kinematics<br />

and dynamics. Trans. <strong>of</strong> the ASME, 107:358–369, 1985.<br />

[108] J.F. O’Brien, J.E. McInroy, D. Brodtke, M. Bruch, and J.C Hamann. Lessons learned in nonlinear<br />

<strong>system</strong>s and flexible robots through experiments on a 6 legged platform. In Proc. <strong>of</strong> the<br />

American Control Conf., pages 868–872, 1998.<br />

[109] J.M. Ortega and W.C. Rheinboldt. Iterative solution <strong>of</strong> non-linear equations in several variables.<br />

In Computer science and applied mathematics. Academic press NY-London, 1970.<br />

[110] J.H. Park and H.C. Cho. Impedance <strong>control</strong> with varying stiffness for parallel-link manipulators.<br />

In Proc. <strong>of</strong> the American Control Conf., pages 478–482, 1998.<br />

[111] P. Piatkiewitz. Reference-model-<strong>based</strong> feedforward <strong>control</strong> for the simona <strong>motion</strong> platform.<br />

Master’s thesis, Delft University <strong>of</strong> Technology, 1997.<br />

[112] W.H. Press, S.A. Teukolsky, W.T. Vetterling, and B.P. Flannery. Numerical Recipes in C; The<br />

Art <strong>of</strong> Scientific Computing; Second Edition. Cambridge, 1992.<br />

[113] Z. Qu and D.M. Dawson. Robust Tracking Control <strong>of</strong> Robot Manipulators. IEEE-Press, 1996.<br />

[114] M. Raghavan. The stewart platform <strong>of</strong> general geometry has 40 configurations. In Proc. ASME<br />

Des Tech. Conf.-Advances in design automation-DE Vol 32-2, 1991.<br />

[115] M.C. Readman. Flexible Joint Robots. CRC, 1994.<br />

[116] C. Reboulet and R. Pigeyre. Hybrid <strong>control</strong> <strong>of</strong> a six-degree-<strong>of</strong>-freedom in-parallel actuated<br />

micro-manipulator mounted on a scara robot. In Robotics and manufacturing: recent trends<br />

in research, education and applications, volume 3, pages 293–298. ASME Press Series, 1990.<br />

[117] L.D. Reid. Technical evaluation report. In AGARD Conference on Flight Simulation - Where<br />

are the challenges, pages T1–T9, 1996.<br />

[118] Joint Aviation Requirements. Aeroplane <strong>flight</strong> <strong>simulator</strong>s jar-std 1a. Technical report, Westward<br />

Digital Limited, 1997.


238 Bibliography<br />

[119] B. Rijnten. Fysisch modelleren van een vlucht<strong>simulator</strong>. Master’s thesis, University <strong>of</strong> Technology<br />

Delft, 1995. In Dutch.<br />

[120] R.E. Roberson and R. Schwertassek. Dynamics <strong>of</strong> multibody <strong>system</strong>s. Springer-Verlag, 1988.<br />

[121] V.V. Rodchenko, L.E. Zaichik, and Y.P. Yashin. Motion cueing on research <strong>flight</strong> <strong>simulator</strong>s.<br />

In CEAS Symposium on Simulation Technologies - Making it real, pages MC02/1–12, 1995.<br />

[122] R. Al Saidi. Robust <strong>control</strong> <strong>of</strong> the simona <strong>flight</strong> <strong>simulator</strong>. Master’s thesis, Delft University <strong>of</strong><br />

Technology, 1999.<br />

[123] A.J. Scheffer. Dynamic calibration <strong>of</strong> the simona research <strong>simulator</strong> <strong>motion</strong> <strong>system</strong> (part 2).<br />

Master’s thesis, Delft University <strong>of</strong> Technology, 1998.<br />

[124] G. Van Schothorst. <strong>Model</strong>ling <strong>of</strong> long-stroke hydraulic servo-<strong>system</strong>s for <strong>flight</strong> <strong>simulator</strong> <strong>motion</strong><br />

<strong>control</strong> and <strong>system</strong> design. PhD thesis, Delft University <strong>of</strong> Technology, 1997.<br />

[125] G. Van Schothorst, P.C. Teerhuis, and A.J.J. van der Weiden. Stability analysis <strong>of</strong> a hydraulic<br />

servo-<strong>system</strong> including transmission line effects. In Proceedings <strong>of</strong> the Third International<br />

conference on Automation, Robitics and Computer Vision, Singapore, pages 1919–1923, 1994.<br />

[126] J.A. Schroeder. Yaw <strong>motion</strong> cues in helicopter simulation. In AGARD Conference Flight<br />

Simulation - Where are the challenges, pages 5.1–5.15, 1995.<br />

[127] N. Sepehri, G.A.M. Dumont, P.D. Lawrence, and F. Sassani. Cascade <strong>control</strong> <strong>of</strong> hydraulically<br />

actuated manipulators. Robotica, pages 207–216, 1990.<br />

[128] R. Sepulchre, M. Jankovic, and P.V. Kokotovic. Constructive Nonlinear Control. Springer,<br />

1997.<br />

[129] A.A. Shabana. Dynamics <strong>of</strong> multibody <strong>system</strong>s. Wiley, 1989.<br />

[130] G.M. Siouris. Aerospace avionics <strong>system</strong>s. Academic Press Inc., 1993.<br />

[131] N.S. Skans and A.G. Barnes. Fifty years <strong>of</strong> succes and failure in <strong>flight</strong> simulation. In Proc. 50<br />

Years <strong>of</strong> Flight Simulation, pages 1–24. The Royal Aeronautical Society, 1979.<br />

[132] S. Skogestad and I. Postlethwaite. Multivariable feedback <strong>control</strong>. Wiley, 1996.<br />

[133] J.-J.E. Slotine and W. Li. Applied Nonlinear Control. Prentice-Hall, 1991.<br />

[134] M. Spong and M. Vidyasagar. Robot Dynamics and Control. Wiley, 1989.<br />

[135] M.W. Spong. <strong>Model</strong>ing and <strong>control</strong> <strong>of</strong> elastic joint robots. ASME Journal on Dyn., Sys.,<br />

Measurement and Control, 109:310–319, 1987.<br />

[136] M.W. Spong, F.L. Lewis, and C.T. Abdallah. Robot Control; Dynamics, Motion Planning and<br />

Analysis. IEEE-Press, 1992.<br />

[137] W. Stadler. Analytical robotics and mechatronics. McGraw-Hill, 1995.<br />

[138] ISO Standard. Iso 2631/1-1985 (e) evaluation <strong>of</strong> human exposure to whole-body vibration-part<br />

1: General requirements. Technical report, International Standard, 1985.


Bibliography 239<br />

[139] Military Standard. Six-degree <strong>of</strong> freedom <strong>motion</strong> <strong>system</strong> requirements for aircrewmember<br />

training <strong>simulator</strong>s. Technical report, MIL-STD-1558, 1974.<br />

[140] D. Stewart. A platform with six degrees <strong>of</strong> freedom. In Proc. Inst. Mech. Engrs., pages 371–<br />

386, 1965.<br />

[141] J. Stoer. Einfuhrung in die Numerische Mathematik 1. Springer-Verlag Berlin, 1983.<br />

[142] S. Tadokoro. Control <strong>of</strong> parallel mechanisms. Advanced Robotics, 8(6):559–571, 1994.<br />

[143] R.P.A.M. Teunissen. Ab-initio zft: Creating pilots without wings. In Proc. Flight Simulation<br />

- Expanding the boundaries, pages 13.1–13.8. The Royal Aeronautical Society, 1997.<br />

[144] M. Tomizuka. Feedforward digital tracking <strong>control</strong>lers for <strong>motion</strong> <strong>control</strong> applications. Advanced<br />

Robotics, pages 575–586, 1993.<br />

[145] J.S. Tyler. The characteristics <strong>of</strong> model-following <strong>system</strong>s as synthesized by optimal <strong>control</strong>.<br />

IEEE Trans. on Automatic Control, AC-15:485–498, 1964.<br />

[146] G.W. van der Linden. High performance robust manipulator <strong>motion</strong> <strong>control</strong>; theory, tools, and<br />

experimental implementation. PhD thesis, Delft University <strong>of</strong> Technology, 1997.<br />

[147] A. van der Schaft. L2-Gain and Passivity Techniques in Nonlinear Control. Springer, 1996.<br />

[148] H. van der Steen. The application <strong>of</strong> a human orientation and <strong>motion</strong> perception model in the<br />

visual and inertial <strong>motion</strong> <strong>of</strong> <strong>flight</strong> <strong>simulator</strong>s. In CEAS Symposium on Simulation Technologies<br />

- Making it real, pages MC01/1–6, 1995.<br />

[149] M.F.C. van Gool. Influence <strong>of</strong> <strong>motion</strong> wash-out filters on pilot tracking performance. In Proc.<br />

AGARD Conf. on Flight Simulation, pages 19.1–19.5, 1978.<br />

[150] J. van Hulzen. Ontwerp en analyse van een robuuste krachtregeling (in dutch). Master’s thesis,<br />

Delft University <strong>of</strong> Technology, 1997.<br />

[151] E. van Zuylen. Static calibration <strong>of</strong> the simona <strong>motion</strong> <strong>system</strong>. Master’s thesis, Delft University<br />

<strong>of</strong> Technology, 1997.<br />

[152] M. Vidyasagar. Control <strong>system</strong> synthesis: a factorization approach. MIT-Press, 1985.<br />

[153] T.J. Viersma. Analysis, synthesis and design <strong>of</strong> hydraulic servo<strong>system</strong>s and pipelines. Technical<br />

report, Delft University <strong>of</strong> Technology, 1980.<br />

[154] J. Wang and O. Masory. On the accuracy <strong>of</strong> a stewart platform - part 1; the effect <strong>of</strong> manufacturing<br />

tolerances. In IEEE Int. Conf. on Robotics and Automation, pages 114–120, 1993.<br />

[155] J. Wang and O. Masory. On the accuracy <strong>of</strong> a stewart platform - part 2; kinematic calibration<br />

and compensation. In IEEE Int. Conf. on Robotics and Automation, pages 725–731, 1993.<br />

[156] S. Wang and B. Chen. <strong>Model</strong> matching: optimal approximation approach. International<br />

Journal <strong>of</strong> Control, pages 1899–1907, 1987.<br />

[157] P.M.R. Wortelboer. Balanced reduction <strong>of</strong> high-dimensional mechanical <strong>system</strong>s: a block<br />

balancing approach. In Proc. 30th IEEE Conf. Decision and Control, pages 1974–1975, 1991.


240 Bibliography<br />

[158] W.C. Yang and W.E. Tobler. Dissipative modal approximation <strong>of</strong> fluid transmission lines using<br />

linear friction model. Journal <strong>of</strong> dynamic <strong>system</strong>s, measurement and <strong>control</strong>, 113:152–162,<br />

1991.<br />

[159] L.R. Young. Visually induced <strong>motion</strong> in <strong>flight</strong> simulation. In Proc. AGARD Conf. on Flight<br />

Simulation, pages 16.1–16.7, 1978.<br />

[160] K. Zhou, J. Doyle, and K. Glover. Robust and Optimal Control. Prentice-Hall, 1996.<br />

[161] H. Zhuang. Self-calibration <strong>of</strong> parallel mechanisms with a case study on stewart platforms.<br />

IEEE Trans. on Robotics and Automation, 13(3):387–397, 1997.<br />

[162] H. Zhuang and Z.S. Roth. Method for kinematic calibration <strong>of</strong> stewart platforms. Journal <strong>of</strong><br />

Robotic Systems, 10(3):391–405, 1993.


Glossary <strong>of</strong> symbols<br />

241


242 Glossary <strong>of</strong> symbols<br />

Arabic symbols (large)<br />

Variable Unit Explanation<br />

A [m 2 ] Actuator operational area<br />

[m] 3x6 matrix with upper gimbal (to c.o.g.) vectors stacked<br />

Ap [m 2 ] Actuator operational area<br />

B [m] 3x6 matrix with lower gimbal vectors stacked<br />

B,Bp [Ns=m�Ns] Viscous friction (matrix)<br />

C [kg=s,kgm 2 =s] Nonlinear coriolis and centripetal terms<br />

Ci [N=m 5 � ;] Oil column stifness chamber i<br />

dC [N=m 5 ] Difference oil column stifness chambers 1 and 2<br />

Cm [N=m 5 ] Mean oil column stifness chambers 1 and 2<br />

E [N=m 2 ] Effective bulk modulus <strong>of</strong> the oil<br />

G [ ] Ground frame or intertial frame<br />

[;] 3x4 transformation as part <strong>of</strong> rotation R = GL T<br />

G T s [;] 3x3 ! to transformation<br />

I [ ] Identity matrix<br />

Iz [kgm 2 ] 3x3 platform inertia matrix evaluated at c.o.g.<br />

Jy�x [ ] Jacobian matrix between two sets <strong>of</strong> variables y and x<br />

Jl�x [;�m] Jacobian from platform velocities to actuator velocities<br />

K [ips=( c irc=s 2 �m=s 2 )] Semicircular and vestibular organ model gain<br />

[m=(sA)] Hydraulic actuator model gain<br />

Kd As=m Valve main spool differential feedback gain<br />

Kp A=m Valve main spool proportional feedback gain<br />

Kh [nd] Crossover model pilot gain<br />

L [ ] 4x3 eul. par. transformation as part <strong>of</strong> rotation R = GL T<br />

Ln [;] 3x6 matrix with all unit actuator directions vectors stacked<br />

Llm [m 5 =N] Oil leakage constant due to pressure difference actuator chambers<br />

Lt [m] Transmission line length<br />

M [kg,kgm 2 ] Mass matrix<br />

[ ] Moving frame<br />

[-] Mobility index<br />

Mact [kg� kgm 2 ] Simulator mass matrix actuator coordinates<br />

Mt [kg� kgm 2 ] Simulator mass matrix platform coordinates


Glossary <strong>of</strong> symbols 243<br />

Variable Unit Explanation<br />

dP [N=m 2 ] Net generated actuator pressure<br />

Pi1 [N=m 2 ] Oil pressure at valve side transmission line 1<br />

Pi2 [N=m 2 ] Oil pressure at valve side transmission line 2<br />

Pm [N=m 2 ] Mean actuator chamber pressure<br />

Pm1�2 [N=m 2 ] Main valve spool pressures<br />

Pn1�2 [N=m 2 ] Pilot valve spool pressures<br />

Po1 [N=m 2 ] Oil pressure at actuator chamber 1<br />

Po2 [N=m 2 ] Oil pressure at actuator chamber 2<br />

Ps [N=m 2 ] Supply oil pressure<br />

Pt [N=m 2 ] Tank oil pressure<br />

Pxn [;] 3x3 projection matrix onto plane with normal x n<br />

R [;] 3x3 rotation matrix<br />

B R A [;] 3x3 rotation matrix from frame A to B<br />

T [;] 3x3 rotation matrix platform<br />

Tx�y�z [;] 3x3 rotation matrix around one axis <strong>of</strong> the frame<br />

U [;] Unitary (rigid body modal direction) matrix<br />

V [m 3 ] General volume<br />

V1 [m 3 ] Oil volume actuator chamber 1<br />

V2 [m 3 ] Oil volume actuator chamber 2<br />

Vm [m 3 ] Mean oil volume actuator chambers 1 and 2<br />

Va [m=s] 3x6 matrix with upper gimbal velocity vectors stacked<br />

X general matrix (capital)<br />

~X vector product matrix<br />

Arabic symbols (small)<br />

Variable Unit Explanation<br />

ac [;] Upper actuator body c.o.g.<br />

ai [m] Vector from c.o.g. <strong>simulator</strong> to i th upper gimbal point<br />

bi [Ns=m] Viscous friction coefficient actuator i<br />

bxf�yf� f [Ns=m� Nms=rad] Viscous friction coefficients foundation<br />

bc [;] Lower actuator body c.o.g.<br />

bi [m] Vector from ground frame origin to i th lower gimbal point<br />

c [N=m] Hydraulic actuator stifness<br />

c1�::: �10 [ ] Valve model parameters<br />

co [m=s] Wave propagation velocity<br />

cxf�yf� f [N=m� Nm=rad] Stifness coefficients foundation<br />

c [m] Vector from ground frame origin to c.o.g. <strong>simulator</strong><br />

d [ ] Finitely small variation<br />

dl [m] Lower gimbal spacing<br />

du [m] Upper gimbal spacing<br />

e [ ] Error vector<br />

[ ] General external signal<br />

dp [N=m 2 ] Pressure difference over hydraulic actuator compartments


244 Glossary <strong>of</strong> symbols<br />

Variable Unit Explanation<br />

f N�Nm Inertial force<br />

fa N 6x1 vector active actuator forces<br />

f (:) [ ] Vector function<br />

g(:) [ ] Vector function<br />

g [N,Nm] Gravitational terms<br />

fd [N] Actuator driving force<br />

frm [1=s] Rigid body mode eigenfrequency<br />

i [A] Valve input current<br />

ia [kgm2 ] Inertia upper actuator body around<br />

ib [kgm2 j<br />

]<br />

[;]<br />

Inertia lower actuator body around around gimbal point<br />

Imaginairy unit, p ;1<br />

kdp [Am2 =N, ] (Normalised) pressure feedback gain<br />

li [m] Length ith joint (serial), actuator (parallel) manipulator<br />

l [m] Vector with all actuator platform lengths<br />

li [m] Vector form the i th lower to upper gimbal<br />

ln [;] Normalised unit direction vector actuator<br />

nG�M x�y�z [m] Unit direction vectors<br />

direction in ground, G, and moving frame, M<br />

m [kg] Mass (platform)<br />

ma [kg] Mass upper actuator body<br />

mb [kg] Mass lower actuator body<br />

mf [kg] Mass <strong>of</strong> the foundation<br />

m [kgm=s� kgrad=s] Vector <strong>of</strong> generalised momenta<br />

n [ ] Euler paramter unit vector along axis <strong>of</strong> rotation<br />

pg [m] Vector to point p in frame G<br />

pm [m] Vector to point p in frame M<br />

q [m] Actuator extension<br />

qi [m] Rotation angle i th joint serial manipulator<br />

q [m,rad] Robotic manipulator positional coordinates<br />

qd [m,rad] Desired positional coordinates<br />

ra [m] Length from c.o.g. upper actuator body to upper gimbal<br />

rb [m] Length from c.o.g. lower actuator body to lower gimbal<br />

s [rad=s] Sloppy Laplace operator as j!<br />

[m] Scaling factor, ka jmax<br />

sx [m] Vector with platform coordinates<br />

t [s] Time<br />

t1�::: �5 [ ] Transmission line model parameters<br />

tg [m] Translational vector in frame G<br />

u [V ] Actuator input voltage<br />

[ ] General <strong>system</strong> input<br />

va [m=s] Upper gimbal point velocity vector<br />

vm m m=s Velocity vector <strong>of</strong> a point in a moving frame


Glossary <strong>of</strong> symbols 245<br />

Variable Unit Explanation<br />

x [ ] General scalar<br />

[m] Surge direction<br />

xf [m] Valve flapper position<br />

[m] Foundation surge displacement<br />

xm [m] Valve main spool position<br />

xs [m] Valve pilot spool position<br />

xa [m] Vector from ground origin to upper gimbal point<br />

xn [ ] General normalized vector (index n) with length one<br />

x [ ] General vector (bar on top)<br />

_x [m=s� rad=s] Vector with translational and angular velocity platform<br />

x g [m] Vector in inertial (ground) frame<br />

x m [m] Vector in platform (moving) frame<br />

xin [(rad� m)=s 2 ] Semicircular and vestibular input (rotational) acceleration<br />

xout [(rad� m)=s 2 ] Noticed semicircular and vestibular (rotational) acceleration<br />

y [m] Sway direction<br />

[ ] General <strong>system</strong> output<br />

z [m] Heave direction<br />

z [m� rad] Vector with generalised (positional) coordinates<br />

Greek symbols (small)<br />

Variable Unit Explanation<br />

[ ] Bound on first Newton Raphson iteration<br />

[ ] Relative damping factor<br />

[ ] Bound on inverse lowest singular value NR-iteration<br />

[rad] Vector <strong>of</strong> euler angles<br />

[] Infinitely small variation<br />

0�1�2�3 [;] The specific four euler parameters<br />

[;] Vector with last three euler parameters 1, 2, 3<br />

e [;] Vector with all four euler parameters<br />

m [V ] Valve spool position error<br />

~ [;] Cross product matrix with euler parameters<br />

[ ] Lipschitz constant<br />

[rad] Angle in general<br />

[rad] Last euler angle around airplane roll axis<br />

[rad] Euler parameter angle <strong>of</strong> rotation<br />

[rad] Second euler angle around intermediate pitch-axis<br />

[kg=m 3 ] Oil density<br />

[ ] Singular value


246 Glossary <strong>of</strong> symbols<br />

Variable Unit Explanation<br />

[N,Nm] Generalized torques, forces<br />

1 [s] Crossover model pilot lead time constant<br />

[s] First semicircular and vestibular lag time constant<br />

2 [s] Crossover model pilot lag time constant<br />

[s] Second semicircular and vestibular lag time constant<br />

L [s] Semicircular and vestibular lead time constant<br />

d [s] Crossover model pilot pure time delay<br />

Nm�N Generalised torques<br />

! [rad=s] Frequency<br />

!o [rad=s] Hydraulic actuator eigenfrequency<br />

! [rad=s] Angular velocity vector platform<br />

!a [rad=s Angularo velocity vector actuator orth. to actuator length<br />

1�::: �5 [ ] Transmission line model parameters<br />

[rad] First euler angle around ground frame yaw axis<br />

Greek symbols (large)<br />

Variable Unit Explanation<br />

i1 [m 3 =s] Oil flow from valve into transmission line 1<br />

i2 [m 3 =s] Oil flow from valve out <strong>of</strong> transmission line 2<br />

l1 [m 3 =s] Leakage oil flow out<strong>of</strong> actuator chamber 1<br />

l2 [m 3 =s] Leakage oil flow into actuator chamber 2<br />

lm [m 3 =s] Leakage oil flow from actuator chamber 2 into 1<br />

m [m 3 =s] Mean oil flow into chamber 1 and out <strong>of</strong> chamber 2<br />

d [m 3 =s] Difference oil flow o1 ; o2<br />

~ [rad=s] Cross product matrix with angular velocity<br />

Caligraphic symbols (large)<br />

Variable Unit Explanation<br />

G [N�Nm] Generalised gravitational forces<br />

H [kgm 2 =s 2 ] Hamiltonian function (addition kinetic and potential energy)<br />

K [kgm 2 =s 2 ] Kinetic energy function<br />

L [kgm 2 =s 2 ] Langrangian function (difference kinetic and potential energy)<br />

P [kgm 2 =s 2 ] Potential energy function


Samenvatting en CV<br />

De hoge veiligheidseisen in de luchtvaart vereisen een goed begrip van het gedrag van de<br />

piloot in extreme weersomstandigheden. Vernieuwend onderzoek op dit gebied moet gebruik<br />

kunnen maken van vlucht<strong>simulator</strong>en die in staat zijn om met hoge nauwkeurigheid<br />

kritische condities tijdens een vlucht te simuleren. Verschijnselen zoals turbulentie en zijwind<br />

in de lagere delen van de atmosfeer vertonen een snelheidsspectrum dat grote krachten<br />

introduceert over een breed scala aan frequenties. Het reproduceren van de benodigde beweging,<br />

het voelbaar maken van de inertiële versnellingen en het juist uitrichten van de<br />

zwaartekracht in geavanceerde vlucht<strong>simulator</strong>en, is dus een uitdagende regeltaak die het<br />

uiterste uit het systeem moet halen. Voor vergelijkbare <strong>system</strong>en, zoals robots, bestaan<br />

moderne modelgebaseerde regelstrategieën, die hogere prestaties kunnen behalen dan minder<br />

gestructureerde methoden, meer inzicht geven in de begrenzingen van het systeem<br />

en wellicht op constructieve eigenschappen van het systeem kunnen wijzen waar gebruik<br />

van kan worden gemaakt in het ontwerp. Deze regelstrategieën worden nog vrijwel niet<br />

toegepast bij bewegings<strong>system</strong>en van vlucht<strong>simulator</strong>en. Eén van de redenen zou kunnen<br />

zijn dat deze <strong>system</strong>en zo complex zijn dat exacte <strong>of</strong> zeer gedetailleerde modellen geen<br />

onderdeel van de regelaar konden en kunnen worden gemaakt. Toepassing ervan in dit<br />

onderzoek vereiste een tussenstap waarin een minder gedetailleerd model moest worden<br />

geëxtraheerd dat toch de meest relevante dynamica moest beschrijven.<br />

De vraag welke relevante systeemkennis gebruikt zou moeten worden in een modelgebaseerde<br />

regelstrategie en in welke mate deze strategie bijdraagt aan het geregelde gedrag<br />

van een bewegingssysteem van een vlucht<strong>simulator</strong>, is het centrale onderzoeksthema van<br />

dit onderzoek geweest.<br />

Om hierop een antwoord te geven moest het volledige regelaarontwerpproces worden<br />

gedefinieerd, gestructureerd en geëvalueerd. Veel van de tussenstappen toegepast in dit proces<br />

kunnen al in de literatuur worden gevonden maar een integrale aanpak, waarin ook de<br />

specifieke eisen ten aanzien van een vlucht<strong>simulator</strong> werden meegenomen, ontbrak nog. Bij<br />

de keuze van bestaande dan wel nieuw voorgestelde bouwstenen in dit proces van systeemen<br />

regelaarontwerp, vormden integreerbaarheid en praktische toepasbaarheid de voornaamste<br />

argumenten.<br />

Door middel van modelvorming op basis van fysische balansvergelijkingen is eerst een<br />

relevante modelstructuur ten behoeve van systeemanalyse en regeling afgeleid. Daarna is<br />

deze structuur geëvalueerd met behulp van experimenten. Daarbij vond identificatie van de<br />

modelparameters en het in kaart brengen van het geldigheidsgebied van de gekozen structuur<br />

plaats. Vervolgens is de meest geschikte modelgebaseerde regelstrategie gekozen en<br />

aangepast voor toepassing op dit bewegingssysteem met zijn specifieke parallelle struc-<br />

247


248 Samenvatting en CV<br />

tuur. Ontwerp en implementatie van deze regeling op het bewegingssysteem van de Simona<br />

vlucht<strong>simulator</strong> was de volgende stap. Tenslotte is een testprocedure gedefinieerd om de<br />

prestatie van dit geregelde systeem te kunnen kwantificeren en te kunnen afzetten tegen de<br />

prestatie van een conventioneel geregeld systeem. Gegeven de specifieke toepassing van<br />

vluchtsimulatie moesten verschillende stappen in dit generieke ontwerpproces nader worden<br />

uitgewerkt.<br />

Opbouwen van vergaand inzicht in de kinematica en dynamica van het systeem was<br />

nodig om zowel de robuustheid van een conventionele regeling te behouden als ook de<br />

prestatie en de voorspelbaarheid te verhogen. Met fysische modelvorming kon kwalitatief<br />

inzicht worden opgebouwd en gelijkertijd gaf deze structuur de mogelijkheid om de parameters<br />

van het zich niet-lineair gedragende bewegingssysteem experimenteel te identificeren<br />

en de te implementeren regelaarstructuur te modulariseren. Met deze identificatie en experimentele<br />

evaluatie kon de vereiste nauwkeurigheid voor het model worden bereikt die<br />

nodig was om de prestatieverhoging met de modelgebaseerde regelaar ook in de praktijk te<br />

kunnen behalen.<br />

Fysische modelvorming van bewegings<strong>system</strong>en begint met de analyse van de systeemkinematica.<br />

In dit onderzoek wordt een nieuwe methode gepresenteerd waarmee het<br />

voor het eerst mogelijk werd om te garanderen dat singuliere punten buiten het werkgebied<br />

van het systeem vallen. Hierop verder bouwend kan ook nagegaan worden <strong>of</strong> een iteratief<br />

schema om de platformcoördinaten te berekenen, gegeven de gemeten lengte van de actuatoren,<br />

convergeert en snel genoeg convergeert voor implementatie in de werkelijke regeling.<br />

Hiermee kon voor het Simona bewegingsysteem de garantie worden geboden dat dit deel<br />

van de regelstructuur veilig toegepast kon worden. Inzicht in de kinematica gaf verder de<br />

mogelijkheid om door middel van calibratie de positioneringsnauwkeurigheid van het al<br />

nauwkeurig gespecificeerde systeem met een factor tien te verhogen waarmee direct ook<br />

het kinematisch model geëvalueerd was.<br />

Van het bewegingssysteem in de vorm van een Stewart platform is afgeleid dat deze tot<br />

de klasse van volledig parallelle <strong>system</strong>en behoort waardoor de dynamica als een stelsel expliciete<br />

differentiaalvergelijkingen beschreven kan worden vanuit de platformcoördinaten.<br />

Zowel in theorie als in de praktijk kon aangetoond worden dat de modes van de stijve<br />

lichamen volgend uit de interactie tussen hydraulica en mechanica de meest relevante systeemdynamica<br />

vormen bij <strong>simulator</strong>toepassingen. De relatie tussen platform- en actuatorcoördinaten<br />

beschreven door een jacobiaan vormt hierbij de centrale operator. Volgend<br />

uit deze analyse kan het systeem, met op het eerste gezicht sterke interactie, in elke stand<br />

lokaal vanuit de juiste coördinaten gezien worden als zes hydraulische actuatoren die elk<br />

vrijwel onafhankelijk een massa aandrijven.<br />

Met de voorgestelde modelgebaseerde regeling kunnen een aantal zaken bereikt worden<br />

die conventioneel niet mogelijk zijn. Ten eerste kunnen de krachten benodigd voor de<br />

gewenste versnelling aangepast worden voor de standsafhankelijke richtingen waarin ontkoppeld<br />

de sterk verschillende massa’s gevoeld worden. Vervolgens kunnen de toch nog<br />

optredende positioneringsfouten ook langs deze richtingen zonder interactie teruggekoppeld<br />

worden. Tenslotte wordt de interactie tussen de hydraulica en de mechanica geminimaliseerd<br />

door te compenseren voor de oliestroom die nodig is om de gewenste snelheid<br />

aan te houden. Implementatie van deze structuur heeft aangetoond dat hiermee een hogere


Samenvatting en CV 249<br />

bandbreedte, minder opslingering en een meer over de vrijheidsgraden geëgaliseerde respons<br />

kan worden bereikt dan met de minder gestructureerde conventionele regeling. Verder<br />

is aangegeven hoe nog meer bereikt kan worden als ook voorspellende informatie vanuit het<br />

voertuigsimulatiemodel aan de regeling wordt aangeleverd.<br />

De te behalen prestaties van de voorgestelde regelaarstructuur worden eerst begrensd<br />

door het optreden van parasitaire effecten ten gevolge van mechanische vervormingen, flexibiliteiten.<br />

Ook met de combinatie van de altijd iets vertraagde olietoevoer en zingende<br />

leidingen moet op nog wat hogere frequenties rekening gehouden worden. Door in het<br />

ontwerp van een <strong>simulator</strong> een laag gewicht en hoge stijfheid na te streven is echter aangetoond<br />

dat met de voorgestelde regelaarstuctuur over een breed spectrum kritische simulatieexperimenten<br />

zeer realistisch uitgevoerd kunnen worden.


250 Samenvatting en CV<br />

Curriculum Vitae<br />

Naam Sjirk Holger Koekebakker<br />

25 oktober 1967 Geboren te Vierhouten (Gemeente Ermelo).<br />

1980 - 1986 VWO aan het Agnes College te Leiden.<br />

1986 - 1993 Werktuigbouwkunde aan de Technische Universiteit Delft.<br />

Afgestudeerd vanuit de vakgroep Systeem- en Regeltechniek.<br />

Titel afstudeerwerk: ”Decoupling by means <strong>of</strong> H 1�F ;<br />

application to a wafer stepper”. Afstudeerwerk uitgevoerd<br />

op het Philips Natuurkundig Laboratorium te Eindhoven.<br />

1993 - 1994 Onderzoeker bij het TNO Productcentrum i.h.k.v.<br />

de vervangende dienst.<br />

1994 - 1998 Promovendus bij de vakgroep Systeem- en Regeltechniek<br />

van de faculteit Werktuigbouwkunde<br />

aan de Technische Universiteit Delft.<br />

1999 - Onderzoeker bij de Group Research Technology<br />

te R & D Venlo van Océ Technologies B.V.

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

Saved successfully!

Ooh no, something went wrong!